(Translated by https://www.hiragana.jp/)
GRAVITATIONAL LENS MODELS BASED ON SUBMILLIMETER ARRAY IMAGING OF HERSCHEL-SELECTED STRONGLY LENSED SUB-MILLIMETER GALAXIES AT z > 1.5 - IOPscience

A publishing partnership

Articles

GRAVITATIONAL LENS MODELS BASED ON SUBMILLIMETER ARRAY* IMAGING OF HERSCHEL-SELECTED STRONGLY LENSED SUB-MILLIMETER GALAXIES AT z > 1.5

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

Published 2013 November 22 © 2013. The American Astronomical Society. All rights reserved.
, , Citation R. S. Bussmann et al 2013 ApJ 779 25 DOI 10.1088/0004-637X/779/1/25

0004-637X/779/1/25

ABSTRACT

Strong gravitational lenses are now being routinely discovered in wide-field surveys at (sub-)millimeter wavelengths. We present Submillimeter Array (SMA) high-spatial resolution imaging and Gemini-South and Multiple Mirror Telescope optical spectroscopy of strong lens candidates discovered in the two widest extragalactic surveys conducted by the Herschel Space Observatory: the Herschel-Astrophysical Terahertz Large Area Survey (H-ATLAS) and the Herschel Multi-tiered Extragalactic Survey (HerMES). From a sample of 30 Herschel sources with S500 > 100 mJy, 21 are strongly lensed (i.e., multiply imaged), 4 are moderately lensed (i.e., singly imaged), and the remainder require additional data to determine their lensing status. We apply a visibility-plane lens modeling technique to the SMA data to recover information about the masses of the lenses as well as the intrinsic (i.e., unlensed) sizes (rhalf) and far-infrared luminosities (LFIR) of the lensed submillimeter galaxies (SMGs). The sample of lenses comprises primarily isolated massive galaxies, but includes some groups and clusters as well. Several of the lenses are located at zlens > 0.7, a redshift regime that is inaccessible to lens searches based on Sloan Digital Sky Survey spectroscopy. The lensed SMGs are amplified by factors that are significantly below statistical model predictions given the 500 μみゅーm flux densities of our sample. We speculate that this may reflect a deficiency in our understanding of the intrinsic sizes and luminosities of the brightest SMGs. The lensed SMGs span nearly one decade in LFIR (median LFIR = 7.9 × 1012L) and two decades in FIR luminosity surface density (median ΣしぐまFIR = 6.0 × 1011L kpc−2). The strong lenses in this sample and others identified via (sub-)mm surveys will provide a wealth of information regarding the astrophysics of galaxy formation and evolution over a wide range in redshift.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Strong gravitational lensing by massive galaxies provides one of the most striking visual confirmations of Einstein's theory of General Relativity. In the case of galaxy–galaxy lensing, the chance alignment of two galaxies along the line of sight provides information about both the lens and the source that cannot be obtained in any other way. The angular separation of multiple images of a lensed galaxy is typically parameterized in terms of the angular Einstein radius (here, denoted θしーたE) and provides an unambiguous measurement of the total mass of the lens (baryonic plus non-baryonic) inside θしーたE, as long as the distances to the lens and source are known (Schneider et al. 1992). At the same time, lensing increases the apparent size of the background source and conserves surface brightness in the process. A spatially unresolved measurement of the flux density from a lensed source is therefore a factor of order μみゅー (the magnification factor, defined in detail in Section 3.1) brighter than for an unlensed source, while spatially resolved measurements can provide a factor of ${\sim} \sqrt{\mu }$ higher resolution (Schneider et al. 1992).

Given the benefits of studying strong lenses, it is no surprise that significant efforts have been devoted to the assembly of large samples of them. The earliest of these efforts focused on surveys in the radio with the Jodrell Bank Very Large Array gravitational lens survey (JVAS; King & Browne 1996) and the Cosmic Lens All-Sky Survey (CLASS; Myers et al. 2003) or on Hubble Space Telescope (HST) follow-up of known strong lenses as part of the Center for Astrophysics Arizona Space Telescope Lens Survey (CASTLeS; Muñoz et al. 1998). Together these surveys have created a sample of ≈80 strong lenses (Schneider et al. 2006). More recently, surveys based on HST Advanced Camera for Surveys (ACS) follow-up of candidates selected from Sloan Digital Sky Survey (SDSS) spectroscopy as part of the Sloan Lens ACS Survey (SLACS; Bolton et al. 2008) and the SDSS Quasar Lens Search (SQLS; Inada et al. 2012), as well as candidates selected from the Canada France Hawaii Telescope Legacy Survey as part of the Strong Lensing in the Legacy Survey (SL2S) project (e.g., Sonnenfeld et al. 2013) have more than doubled this number. More recent upgrades associated with the SDSS-III Baryon Oscillation Spectroscopic Survey (BOSS; Eisenstein et al. 2011) promise to increase the sample size of SDSS-selected lenses by a factor of several (the BOSS Emission-Line Lens Survey, or BELLS; Brownstein et al. 2012). Finally, a new method of finding lenses has come to sudden prominence with the launch of the Herschel Space Observatory (Herschel; Pilbratt et al. 2010) and the advent of the South Pole Telescope (SPT; Carlstrom et al. 2011) and the Atacama Cosmology Telescope (Swetz et al. 2011): wide-field surveys at submillimeter (sub-mm) and millimeter (mm) wavelengths.

Surveys at (sub-)mm wavelengths are ideal tools for discovering lenses, in part because the observed-frame (sub-)mm flux density of a dusty galaxy at a given luminosity is approximately independent of redshift for z > 1 (Blain & Longair 1993) and in part because the number counts of unlensed submm sources (SMGs) falls off very steeply at high flux densities compared to optically selected galaxies (e.g., Barger et al. 1999; Coppin et al. 2006; Oliver et al. 2010; Clements et al. 2010). Strong lensing events are rare, so the key requirement for identifying them is wide-area coverage. This is now being provided by the Herschel Astrophysical Terahertz Large Area Survey (H-ATLAS; Eales et al. 2010), the Herschel Multi-Tiered Extragalactic Survey (HerMES; Oliver et al. 2012),33 the SPT (Vieira et al. 2010; Mocanu et al. 2013), and ACT (Marsden et al. 2013). In this paper, we focus on strong lens candidates selected from the two Herschel surveys.

Studies based on the H-ATLAS Science Demonstration Phase field (covering 14.4 deg2) and on HerMES (covering 94.8 deg2) have found that a simple selection at 500 μみゅーm of S500 > 100 mJy finds lenses with an efficiency of 70%–100% (Negrello et al. 2010; Wardlow et al. 2013). This single selection criterion also yields low-z spiral galaxies (Serjeant & Harrison 2005) and blazars with synchrotron emission spectra in the Herschel bands (de Zotti et al. 2005), but these are easily identified and removed using shallow optical imaging (SDSS is sufficient) and shallow radio imaging from the National Radio Astronomy Observatory Very Large Array Sky Survey (NVSS; Condon et al. 1998).

A world-wide, multi-wavelength follow-up effort is now underway to study strong lens candidates from H-ATLAS and HerMES that are selected to have S500 > 100 mJy. The main goals of this effort are to: (1) confirm that the candidates are indeed strong lenses; (2) use the lenses to study massive galaxy evolution over 0.2 < z < 1.3; and (3) use the lensed SMGs to learn about the nature of star-formation and galaxy evolution in luminous, dusty galaxies at z ≳ 1.5.

There are three key steps that must be taken to confirm the lensing hypothesis and study a member of the Herschel sample in detail: a redshift measurement for the lens (typically from optical spectroscopy); a distinct redshift for the background source (typically from radio or (sub-)mm wave spectroscopy); and spatially-resolved imaging of the source that is consistent with strong lensing. In this paper, we present data that mark significant progress on two of these three fronts. First, we give results from a large, multi-semester program with the Submillimeter Array (SMA; Ho et al. 2004) that comprises over 160 hr of on-source integration time and provides sub-arcsecond, spatially-resolved 880 μみゅーm images of 30 Herschel lens candidates (some of the SMA data have been published previously and are known lenses—we highlight where this is the case and provide references in Section 3.2). Second, we provide redshifts from optical spectroscopy of 8 lens candidates obtained with the Multiple Mirror Telescope (MMT), Gemini-South telescope (Gemini-S), William Herschel Telescope (WHT), and Very Large Telescope (VLT). Redshift measurements for the background sources for nearly all of the objects in this paper are already available from a wide range of facilities (see Section 2.1).

We use the SMA imaging, optical spectroscopy, and radio or (sub-)mm wave spectroscopy to determine which of the Herschel-selected lens candidates are indeed strongly lensed (μみゅー > 2, with multiple images of the lensed source), which are only moderately lensed (1 < μみゅー < 2, with only a single image of the lensed source), and which are inconclusive. We then develop and apply lens models in the visibility plane—as is appropriate for interferometers like the SMA—for all of the objects that show convincing evidence of moderate or strong lensing. This provides measurements of the total (baryonic and non-baryonic) masses within θしーたE (ME), the magnification factors of the background sources at 880 μみゅーm (μみゅー880) and the sizes of the background sources (rhalf). These are fundamental parameters needed to understand the physics of galaxy evolution at intermediate redshift (0.2 < z < 1.3) and high redshift (z > 1.5).

In Section 2, we describe our selection technique and present the SMA, MMT, Gemini-S, WHT, and VLT data (highlighting which datasets are new to this paper and which have been published previously). We also show HST or Keck adaptive optics (AO) imaging for comparison with the SMA imaging and reference the future papers that will present the HST and Keck data. Section 3 contains a description of our lens modeling methodology and a detailed description of each object in the sample. We discuss the implications of the lens modeling for the population of foreground galaxies discovered by Herschel and compare to existing surveys for lenses in Section 4. The lens model implications for the lensed SMGs are discussed in Section 5, with an emphasis on the size-scale of star-formation in SMGs at 1.5 < z < 4.5. Finally, we present our conclusions in Section 6.

Throughout this paper, we assume H0 = 71 km s−1 Mpc−1, $\Omega _{\rm m_0} = 0.27$, and $\Omega _{\Lambda _0} = 0.73$.

2. DATA

In this section, we describe how candidate strongly lensed SMGs are selected from wide-field Herschel surveys and present SMA high-spatial resolution imaging of the dust continuum emission from these candidate lensed SMGs as well as optical spectroscopy obtained with the MMT, Gemini-S, and WHT. We also highlight ancillary optical and near-IR imaging that is used to determine the position of the lensing galaxy or galaxies and reference the papers that fully present and analyze those data.

2.1. Selection of Candidate Lensed SMGs

The first suggestion that wide-field surveys (i.e., covering ≳ 100 deg2) at submm or mm wavelengths would efficiently identify strongly lensed galaxies was made nearly two decades ago (Blain 1996), but it is only in the past few years, with the advent of Herschel and the SPT, that such surveys have reached the requisite survey area and sensitivity to discover them in large numbers. We select candidate strongly lensed galaxies from the two widest Herschel extra-galactic surveys: H-ATLAS and HerMES. The total area considered for the candidate selection is ≈300 deg2 in H-ATLAS (comprising the full equatorial fields and ≈75% of the northern galactic pole field) and 94.8 deg2 in nine independent fields in HerMES (for details of the fields, see Oliver et al. 2012). The total area surveyed by Herschel as part of H-ATLAS and HerMES is ∼1000 deg2 (i.e., roughly a three-fold increase over the area considered for this paper).

An important aspect of candidate lens selection is source extraction and photometry for the Herschel Spectral and Photometric Imaging REceiver (SPIRE; Griffin et al. 2010) data. We summarize the relevant aspects of the methodology here and provide references where appropriate.

In the HerMES fields, source detection is achieved by applying the StarFinder code (Diolaiti et al. 2000) to the 250 μみゅーm images. Photometry is then computed using the HerMES XID pipeline (Roseboom et al. 2010), which allocates flux density based on the 250 μみゅーm position priors obtained with StarFinder.

In the H-ATLAS fields, sources are identified and flux densities are measured using the Multi-band Algorithm for source eXtraction (MADX; S. Maddox et al., in preparation). MADX first subtracts a smooth background, and then filters with the point-spread function (PSF) appropriate for each band. Next, >2.5σしぐま peaks are identified in the 250 μみゅーm map, and "first-pass" flux density estimates are obtained from the pixel values at these positions in each band. Sub-pixel positions are estimated by fitting to the 250 μみゅーm peaks, and more accurate flux-densities are estimated using bi-cubic interpolation to the accurate 250 μみゅーm position. In each band, the sources are sorted in order of decreasing flux density using the first-pass pixel values, and a scaled PSF is subtracted from the map before estimation of flux densities for any fainter sources. This step prevents faint source flux densities from being overestimated when they lie near brighter sources. Finally candidate sources are retained in the catalog if their flux densities are more than 5σしぐま in any of the three bands. The 5σしぐま flux density limits in H-ATLAS, including confusion noise (typically ≈6 mJy in all three SPIRE bands Nguyen et al. 2010), are 32 mJy at 250 μみゅーm, 36 mJy at 350 μみゅーm, and 45 mJy at 500 μみゅーm (Pascale et al. 2011; Rigby et al. 2011).

Although the HerMES and H-ATLAS teams use different methods to extract photometry, the sources that are the subject of this paper are all high signal-to-noise ratio (S/N), point sources as seen by Herschel. In this regime, we expect that the different methods should provide consistent flux density measurements.

The essence of the candidate lens selection technique is to identify objects that are bright at 500 μみゅーm. A complete description of the selection technique for the HerMES lens candidates is given by Wardlow et al. (2013). This paper includes the objects tabulated in Wardlow et al. (2013) as well as objects identified in H-ATLAS. We select objects that satisfy S500 > 100 mJy, a regime that has been shown from previous studies of smaller areas to have relatively little contamination from unlensed SMGs (Negrello et al. 2010; Wardlow et al. 2013). The primary contaminant is local universe galaxies (z < 0.1). These galaxies are spatially resolved in SDSS imaging and therefore trivial to remove. There is also a small contamination from blazars, which are non-thermal emitters and are easily removed using data from the NVSS or the Very Large Array Faint Images of the Radio Sky at Twenty-Centimeters survey (FIRST; Becker et al. 1995).

A total of 13 objects in HerMES satisfy S500 > 100 mJy and are not local galaxies or blazars (Wardlow et al. 2013). In the H-ATLAS fields, there are 91 objects that satisfy these criteria. Considering the combination of the two surveys, this results in a surface density on the sky of ≈0.26 deg−2. This value lies between the values of 0.32 deg−2 from Negrello et al. (2010) and 0.14 ± 0.04 deg−2 from Wardlow et al. (2013), as expected since it represents a combination and extension of these previous efforts. Cosmic variance likely explains the difference in the surface densities of lenses between HerMES and H-ATLAS. A detailed calculation of this effect requires taking into account the cosmic variance of both the lensed SMGs and the lenses themselves and is beyond the scope of this paper.

Efforts are on-going to obtain a complete database of follow-up observations for this sample of 104 candidate lensed SMGs. The present paper focuses on a subset of 30 candidates with superb existing follow-up observations (hereafter, we refer to this as the "SMA subsample"). These targets were initially selected on the basis of strong 1.2 mm detections from the Max Planck Millimeter Bolometer (MAMBO) array (Kreysa et al. 1998) at the Institut de Radioastronomie Millimétrique (IRAM) 30 m telescope (H. Dannerbauer et al. in preparation). Subsequent follow-up efforts have now provided high-spatial resolution 880 μみゅーm imaging with the SMA, spectroscopic redshifts of the lensed SMGs obtained with GBT, CSO, CARMA, PdBI, and Herschel (Cox et al. 2011; Harris et al. 2012; Lupu et al. 2012, D. A. Riechers et al., in preparation; M. Krips et al., in preparation, George et al. 2013), and spectroscopic redshifts to the lenses obtained with the MMT, Gemini-S, or WHT. In addition, Keck-II Near InfraRed Camera 2 (NIRC2) laser guide star adaptive optics (LGSAO) imaging has been obtained for nearly half of the candidate lensed SMG sample (Wardlow et al. 2013; J. Calanog et al., in preparation). These datasets provide the information needed to confirm the lensing hypothesis and begin analysis of the source and lens properties. Table 1 provides basic positional data for the SMA subsample, including the International Astronomical Union names, short names to aid comparison with previous publications, positions measured from the SMA 880 μみゅーm image (see Section 2.2), and redshift measurements for the lens(es) (see Sections 2.32.42.52.6, and references in the table) and background sources (references given in the table), where available.

Table 1. Positions and Redshifts of SMA Candidate Strong Gravitational Lens Sample

IAU Name Short Name R.A.880 Decl.880 zlens Ref. zsource Ref. Lens
(J2000) (J2000) Grade
1HerMES S250 J021830.5−053124 HXMM02 02:18:30.679 −05:31:31.60 1.35 ± 0.01 W13 3.39 ± 0.01 R13 A
1HerMES S250 J022016.5−060143 HXMM01 02:20:16.603 −06:01:43.20 0.6540.502a W13 2.307 ± 0.001 F13 C
H-ATLAS J083051.0+013224 G09v1.97 08:30:51.156 +01:32:24.35 0.6261.002a New 3.634 ± 0.001 R13 A
H-ATLAS J084933.4+021443 G09v1.124 08:49:33.362 +02:14:42.30 0.3478 ± 0.0001 I13 2.410 ± 0.003 H12 C
H-ATLAS J085358.9+015537 G09v1.40 08:53:58.862 +01:55:37.70 ... ... 2.0894 ± 0.0009 L13 B
H-ATLAS J090302.9−014127 SDP17 09:03:03.031 −01:41:27.11 0.9435 ± 0.0009 N10 2.3051 ± 0.0002 L12 A
H-ATLAS J090311.6+003906 SDP81 09:03:11.568 +00:39:06.43 0.2999 ± 0.0002 SDSS 3.042 ± 0.001 F11 A
H-ATLAS J090740.0−004200 SDP9 09:07:40.022 −00:41:59.80 0.6129 ± 0.0005 New 1.577 ± 0.008 L12 A
H-ATLAS J091043.1−000321 SDP11 09:10:43.061 −00:03:22.76 0.7932 ± 0.0012 N10 1.786 ± 0.005 L12 A
H-ATLAS J091305.0−005343 SDP130 09:13:05.107 −00:53:43.05 0.220 ± 0.002 N10 2.6256 ± 0.0005 F11 A
H-ATLAS J091840.8+023047 G09v1.326 09:18:40.927 +02:30:45.90 ... ... 2.5811 ± 0.0012 H12 X
1HerMES S250 J103826.6+581542 HLock04 10:38:26.611 +58:15:42.47 0.61 ± 0.02 W13 ... ... B
1HerMES S250 J105712.2+565457 HLock03 10:57:12.262 +56:54:58.70 ... W13 2.771 ± 0.001 R13 X
1HerMES S250 J105750.9+573026 HLock01 10:57:51.022 +57:30:26.80 0.60 ± 0.04 W13 2.957 ± 0.001 S11 A
H-ATLAS J113526.3−014605 G12v2.43 11:35:26.273 −01:46:06.55 ... ... 3.1276 ± 0.0005 H12 X
H-ATLAS J114637.9−001132 G12v2.30 11:46:37.980 −00:11:31.80 1.2247 ± 0.0001 New 3.2592 ± 0.0010 H12 A
H-ATLAS J125135.4+261457 NCv1.268 12:51:35.412 +26:14:58.63 ... ... 3.675 ± 0.001 K13 B
H-ATLAS J125632.7+233625 NCv1.143 12:56:32.544 +23:36:27.63 0.2551 ± 0.0001 New 3.565 ± 0.001 R13 A
H-ATLAS J132427.0+284452 NBv1.43 13:24:27.206 +28:44:49.40 0.997 ± 0.017 G05 1.676 ± 0.001 G13 C
H-ATLAS J132630.1+334410 NAv1.195 13:26:30.216 +33:44:07.60 0.7856 ± 0.0003 New 2.951 ± 0.001 H13 A
H-ATLAS J132859.3+292317 NAv1.177 13:28:59.246 +29:23:26.13 ... ... 2.778 ± 001 K13 X
H-ATLAS J133008.4+245900 NBv1.78 13:30:08.520 +24:58:59.17 0.4276 ± 0.0003 New 3.1112 ± 0.0001 R13 A
H-ATLAS J133649.9+291801 NAv1.144 13:36:49.985 +29:17:59.77 ... ... 2.2024 ± 0.0002 H12 B
H-ATLAS J134429.4+303036 NAv1.56 13:44:29.518 +30:30:34.05 0.6721 ± 0.0004 New 2.3010 ± 0.0009 H12 A
H-ATLAS J141351.9−000026 G15v2.235 14:13:52.092 −00:00:24.43 0.5470 ± 0.0003 New 2.4782 ± 0.0005 H12 C
H-ATLAS J142413.9+022303 G15v2.779 14:24:13.975 +02:23:03.60 0.595 ± 0.005 B12 4.243 ± 0.001 C11 A
1HerMES S250 J142823.9+352619 HBootes03 14:28:24.074 +35:26:19.35 1.034 ± 0.001 B06 1.325 ± 0.001 B06 C
1HerMES S250 J142825.5+345547 HBootes02 14:28:25.476 +34:55:47.10 0.414 ± 0.001 W13 2.804 ± 0.001 R13 A
1HerMES S250 J143330.8+345439 HBootes01 14:33:30.826 +34:54:39.75 0.59 ± 0.08 W13 3.274 ± 0.001 R13 A
H-ATLAS J144556.1−004853 G15v2.481 14:45:56.297 −00:48:51.70 ... ... ... ... X

Notes. Definition of lens grades: A = Obvious strong lensing morphology in SMA map and distinct lens and source redshifts; B = Obvious strong lensing morphology in SMA map, but only a single redshift measurement (either lens or source); C = Evidence for moderate lensing from SMA map and distinct lens and source redshifts; X = SMA imaging and spectroscopic redshifts do not provide conclusive evidence of lensing. Reference key: G05 = Gladders & Yee (2005); B06 = Borys et al. (2006); N10 = Negrello et al. (2010); S11 = Scott et al. (2011); F11 = Frayer et al. (2011); C11 = Cox et al. (2011); H12 = Harris et al. (2012); B12 = Bussmann et al. (2012); L12 = Lupu et al. (2012); W13 = Wardlow et al. (2013); I13 = Ivison et al. (2013); G13 = R. D. George et al. (in preparation); R13 = D. A. Riechers et al. (in preparation); K13 = M. Krips et al. (in preparation); L13 = R. E. Lupu et al. (in preparation); H13 = A. I. Harris et al. (in preparation). aMultiple lens redshifts have been measured for these targets. The redshift uncertainty is 0.001 in all cases.

Download table as:  ASCIITypeset image

Figure 1 shows the S350/S500 SPIRE colors as a function of S500 flux density for all galaxies in the H-ATLAS phase I catalog with S/N >3 in all SPIRE bands (grayscale, logarithmic scaling), the full sample of 104 candidate Herschel lensed SMGs in HerMES and H-ATLAS (cyan squares), the SMA subsample with superb follow-up data that is the focus of this paper, and a sample of objects selected from the SPT survey with published lens models (Hezaveh et al. 2013). The SMA subsample is biased to higher S500 values than the full sample. We therefore expect the lensing rate to be higher than in the full sample. The SPIRE colors (S350/S500 and S250/S350) are comparable between the full sample and the SMA subsample. The SMA subsample contains nearly all galaxies in the parent sample with S500 > 170 mJy (the lone exception is H-ATLAS J1429−002, which has the highest S350/S500 ratio in the full lens candidate sample and is the subject of a study based on data from the Atacama Large Millimeter/submillimeter Array (ALMA) and other facilities; H. Messias et al., in preparation).

Figure 1.

Figure 1. Herschel/SPIRE photometry of all galaxies in the H-ATLAS phase I catalog with S/N >3 at 250 μみゅーm, 350 μみゅーm, and 500 μみゅーm (grayscale). The sample of H-ATLAS and HerMES sources that satisfy the selection criteria used to select lens candidates are overplotted with yellow filled squares. The targets presented in this paper are represented by red filled circles ("SMA subsample"), and a comparison sample of lensed SMGs discovered by the SPT that have published lens models are represented by cyan stars. Representative error bars are shown in the lower right corner. The SMA subsample is biased toward higher 500 μみゅーm flux densities but has similar S350/S500 and S250/S350 colors (not shown).

Standard image High-resolution image

2.2. SMA Imaging

SMA data were obtained over a period of multiple semesters from 2010 March through 2013 May with a total of 162 hr of on-source integration time. Each target was typically observed in multiple array configurations with tint = 1–2 hr on-source per configuration. We used track-sharing of multiple targets per track to ensure the best possible uv coverage.

Observations took place in a range of conditions from superb (atmospheric opacities of τたう225 GHz = 0.04, phase errors of Δでるたrms) = 10°) to good (atmospheric opacities of τたう225 GHz = 0.1, phase errors of Δでるたrms) = 40°). Phase errors are estimated from a fixed monitoring system on a variety of baselines. For some of the observations, 1 or more antennas were unavailable, so in those cases the total number of antennas (Nant) used was less than 8. Most notably, in early 2011, the lower sideband of the SMA 345 GHz receiver in antenna 1 was flagged due to significant instrumental noise. In general, we optimized the SMA single-polarization 345 GHz receivers for continuum detection by tuning to a frequency of νにゅーLO ≈ 340 GHz. In some cases νにゅーLO varied from this value by up to 10 GHz to avoid retuning in the middle of the night when the SMA switched from another program. Table 2 presents details regarding the SMA observations, including the date, array configuration, local oscillator frequency (νにゅーLO), τたう225 GHz, ϕrms, Nant, tint, and original reference (some of the data used in this paper were originally presented elsewhere).

Table 2. SMA Observations

IAU Name Reference UT Date Array νにゅーLO τたう225 GHz Δでるたrms) Nant tint
Configurationa (GHz) (deg) (hr)
J021830.5−053124 I11 2009 Dec 10 COM 339.925 0.10 20 7 3.7
... W13 2010 Sep 25 EXT 342.003 0.10 30 7 2.0
... New 2011 Jan 26 VEX 343.160 0.08 10 6.5b 1.9
J022016.5−060143 W13 2010 Aug 14 SUB 342.017 0.07 10 8 0.9
... W13 2010 Sep 26 EXT 342.001 0.07 20 7 2.5
... F13 2011 Jan 4 VEX 340.226 0.06 25 6.5b 3.2
J083051.0+013224 New 2011 Dec 8 COM 339.564 0.09 20 8 1.4
... New 2012 Feb 4 EXT 339.946 0.04 20 8 2.4
J084933.4+021443 I13 2011 Dec 8 COM 339.564 0.09 20 8 1.4
... I13 2012 Jan 31 EXT 339.537 0.05 30 7 4.9
... I13 2012 Mar 31 VEX 339.917 0.05 25 7 1.7
J085358.9+015537 New 2011 Dec 8 COM 339.564 0.09 20 8 1.4
... New 2012 Feb 4 EXT 339.946 0.04 20 8 2.4
... New 2012 Mar 31 VEX 339.917 0.05 25 7 1.7
... New 2012 Apr 9 VEX 339.915 0.05 10 7 1.3
J090302.9−014127 New 2010 Dec 16 COM 342.410 0.13 20 7.5b 1.5
... New 2011 Jan 30 EXT 340.244 0.05 20 6.5b 1.7
... New 2011 Jan 26 VEX 343.160 0.08 35 6.5b 2.0
... New 2012 Apr 9 VEX 339.915 0.05 10 7 1.3
J090311.6+003906 N10 2010 Mar 16 SUB 340.725 0.05 10 5 2.5
... N10 2010 Apr 9 COM 341.609 0.07 25 6 2.6
... N10 2010 Apr 20 COM 340.714 0.06 30 7 2.4
... N10 2010 Feb 25 VEX 340.735 0.10 40 8 5.0
J090740.0−004200 New 2010 Dec 16 COM 342.410 0.13 20 7.5b 1.5
... New 2011 Jan 30 EXT 340.244 0.05 20 6.5b 1.7
... New 2012 Feb 6 EXT 339.989 0.05 30 8 0.9
J091043.1−000321 New 2010 Dec 16 COM 342.410 0.13 20 7.5b 1.5
... New 2011 Jan 30 EXT 340.244 0.05 20 6.5b 1.7
... New 2012 Feb 7 EXT 339.993 0.04 15 8 2.0
... New 2011 Jan 26 VEX 343.160 0.08 35 6.5b 2.0
J091305.0−005343 N10 2010 Mar 16 SUB 340.725 0.05 10 5 2.5
... N10 2010 Apr 9 COM 341.609 0.07 25 6 2.6
... N10 2010 Apr 20 COM 340.714 0.06 30 7 2.4
... N10 2010 Feb 28 VEX 340.735 0.07 5 7 5.0
J091840.8+023047 New 2012 Feb 6 EXT 339.989 0.05 30 8 0.9
... New 2012 Feb 7 EXT 339.993 0.04 15 8 2.0
J103826.6+581542 W13 2010 May 16 COM 341.981 0.06 35 7 3.8
J105712.2+565457 W13 2010 Dec 6 COM 338.148 0.05 10 8 1.1
... New 2011 Jan 4 VEX 340.226 0.08 20 6.5b 2.1
J105750.9+573026 C11 2010 May 14 COM 340.742 0.06 10 7 4.2
... New 2011 Jan 4 VEX 340.226 0.08 20 6.5b 2.1
J113526.3−014605 New 2012 Feb 4 EXT 339.946 0.04 20 8 1.3
J114637.9−001132 F12 2012 Jan 14 SUB 336.929 0.15 20 7 2.0
... F12 2011 May 22 COM 339.579 0.08 25 7 1.0
... New 2012 Feb 4 EXT 339.946 0.04 20 8 1.3
J125135.4+261457 New 2012 May 22 COM 339.579 0.09 35 7 1.7
... New 2012 Feb 6 EXT 339.989 0.05 30 8 1.2
J125632.7+233625 New 2012 May 25 COM 340.045 0.08 20 7 0.8
... New 2012 Feb 6 EXT 339.989 0.05 30 8 1.2
J132427.0+284452 F13a 2011 Dec 15 COM 339.561 0.05 10 8 3.0
... F13a 2012 Jan 31 EXT 339.537 0.05 30 7 4.4
... F13a 2012 Apr 9 VEX 339.915 0.05 10 7 2.3
J132630.1+334410 New 2012 Feb 7 EXT 339.993 0.04 15 8 1.4
J132859.3+292317 New 2013 May 3 SUB 340.757 0.12 25 6 0.8
... New 2012 Feb 7 EXT 339.993 0.04 15 8 1.4
... New 2012 Apr 24 VEX 339.960 0.06 20 7 2.5
J133008.4+245900 New 2012 May 25 COM 340.045 0.08 20 7 0.8
... New 2012 Feb 7 EXT 339.993 0.04 15 8 1.4
J133649.9+291801 New 2013 May 3 SUB 340.757 0.12 25 6 0.8
... New 2012 Feb 6 EXT 339.989 0.05 30 8 1.2
... New 2012 Apr 24 VEX 339.960 0.06 20 7 2.5
J134429.4+303036 New 2011 May 22 COM 339.579 0.09 35 7 1.7
... New 2011 Jul 26 EXT 341.037 0.06 15 8 1.7
... New 2012 Mar 17 EXT 339.949 0.04 35 7 3.2
... New 2012 Apr 9 VEX 339.915 0.05 10 7 2.3
J141351.9−000026 New 2011 May 23 COM 339.544 0.06 30 7 2.9
... New 2011 Jan 30 EXT 340.244 0.05 20 6.5b 1.1
... New 2011 Jan 24 VEX 341.449 0.04 10 5.5b 1.6
... New 2011 Jan 26 VEX 343.160 0.09 30 6.5b 1.3
J142413.9+022303 B12 2010 Jun 16 COM 342.100 0.10 30 8 3.0
... B12 2011 Jan 28 EXT 340.711 0.08 30 6.5b 1.3
... B12 2011 Jan 4 VEX 340.226 0.10 20 6.5b 1.8
... B12 2011 Jan 6 VEX 340.225 0.04 10 6.5b 1.7
J142823.9+352619 New 2011 Jan 28 EXT 340.711 0.08 10 6.5b 1.0
... New 2011 Feb 4 EXT 350.086 0.12 35 7 1.3
J142825.5+345547 W13a 2012 May 25 COM 340.045 0.08 20 7 0.8
... W13a 2011 Jul 26 EXT 341.037 0.06 15 8 1.7
J143330.8+345439 W13 2010 Dec 16 COM 342.410 0.11 20 7.5b 0.8
... W13 2011 Jan 28 EXT 340.711 0.08 10 6.5b 1.0
... W13 2011 Feb 4 EXT 350.086 0.12 35 7 1.3
J144556.1−004853 New 2011 May 23 COM 339.544 0.06 30 7 2.9
... New 2011 Jan 30 EXT 340.244 0.05 20 6.5b 1.1
... New 2011 Jan 26 VEX 343.160 0.09 30 6.5b 1.3

Notes. References key: N10 = Negrello et al. (2010); C11 = Conley et al. (2011); I11 = Ikarashi et al. (2011); F12 = Fu et al. (2012); B12 = Bussmann et al. (2012); W13 = Wardlow et al. (2013); F13 = (Fu et al. 2013); I13 = Ivison et al. (2013); W13a = J. L. Wardlow et al. (in preparation); F13a = H. Fu et al. (in preparation). aSUB = subcompact (longest baseline length ≈25 m); COM = compact (longest baseline length ≈75 m); EXT = extended (longest baseline length ≈220 m); VEX = very extended (longest baseline length ≈510 m). bThe lower sideband of one antenna was flagged for these observations.

Download table as:  ASCIITypeset images: 1 2

The SMA receivers make use of an intermediate frequency coverage of 4 GHz, providing a total of 8 GHz bandwidth (considering both sidebands), with a center-to-center sideband separation of 12 GHz. The primary goal of the observations is to detect the continuum emission at the highest possible significance. When a spectroscopic redshift for a Herschel source was available, we tuned the receivers to the closest CO rotational transition, as long as the S/N of the continuum data would not be compromised by doing so. This was possible for 8 of our targets. Since the background galaxies lie at 1.5 < z < 4.5, our observations probe emission from the J = 10 (or higher) levels in lines that are typically faint in SMGs and therefore difficult to detect. We defer a discussion of the molecular line measurements based on the SMA data to a future publication.

Bandpass calibrators were chosen primarily based on their 880 μみゅーm flux densities (where possible, we used calibrators with S880 > 5 Jy) and their observability at the beginning or end of each night. For absolute flux density calibration, Titan was used whenever possible, followed by Callisto, Neptune (in subcompact array only), and MWC349A (when no planets or moons were available). Amplitude and phase gain calibration was achieved by monitoring nearby (angular separation from science target of <15°), bright (S880 > 0.5 Jy) quasars. Whenever possible, we used multiple quasars for gain calibration, including a fainter quasar (S880 > 0.1 Jy) much closer to the science target (angular separation <5°) to provide an independent check of the reliability of the calibration, particularly phase transfer. We used the Interactive Data Language (IDL) MIR package to calibrate the uv visibilities.

We used the Multichannel Image Reconstruction, Image Analysis, and Display (MIRIAD) software package (Sault et al. 1995) to reconstruct and deconvolve the image from the visibilities. We used natural weighting to achieve maximum sensitivity for all targets. We combine visibility data from all available configurations for each target. The beam size and shape in the resulting images vary greatly from target to target. This is primarily because not every target was observed in all array configurations, but there is also some dependence on the declination of the target, because the SMA uv coverage is less complete at declinations near 0°. In general, we achieve spatial resolutions of ≈0farcs6 FWHM.

Figure 2 shows the SMA image of each object in the SMA subsample (red contours, beginning at ±3σしぐま and increasing by factors of $\sqrt{2}$) in comparison with the best available optical or near-IR image (grayscale; see Section 2.7). A detailed source-by-source description is deferred to Section 3.2. The position of the 880 μみゅーm emission centroid (estimated by-eye) for each source is presented in Table 1. There is no absolute significance to these centroid values, but they are necessary to undertake lens modeling.

Figure 2.

Figure 2. SMA 880 μみゅーm images (red contours, starting at ±3σしぐま and increasing by factors of $\sqrt{2}$) of candidate lensed SMGs from H-ATLAS and HerMES, overlaid on best available optical or near-IR images (logarithmic scaling; telescope and filter indicated in lower left corner of each panel). North is up and East is left, with axes having units of arcseconds relative to the 880 μみゅーm centroid as given in Table 1. The elliptical FWHM of the SMA's synthesized beam is shown in the lower right corner of each panel. The image separations are ≈1–2'', suggesting gravitational potential wells typical of isolated galaxies or small numbers of galaxies for the lenses (only two lensed sources are associated with galaxy clusters: J132427.0+284452 and J141351.9−000026).

Standard image High-resolution image

We use the SMA images in conjunction with knowledge of the redshifts of the lenses and sources (see Sections 2.32.42.5, and 2.6 for details) to characterize the nature of the lensing that is occurring in the SMA subsample. Those galaxies showing multiple images with a morphology typical of strong lensing and that have known distinct lens and source redshifts are given an A grade. Galaxies with obvious strong lensing morphology but with only a single known redshift measurement (either of the lens or the source) receive a B grade. We expect that all B grade systems are strong lenses, but without distinct redshift measurements we cannot be certain. Galaxies showing only a single image of the background source, but with known distinct lens and source redshifts, are given a C grade. Finally, an X grade is given to those objects where the SMA imaging and the available spectroscopic redshifts provide inconclusive evidence of lensing. Additional data are needed to determine whether lensing is occurring in these objects. Our grades are listed in Table 1.

We compute total flux densities at 880 μみゅーm within rectangular apertures customized to match the spatial extent of each object in the SMA images. Uncertainties on these measurements are derived by placing apertures of the same size and shape at random, non-overlapping locations within the SMA primary beam field of view (excluding regions containing flux density from the source) and computing the 1σしぐま root-mean-square variation (which is generally well-described by a Gaussian) in the distribution of aperture flux densities. The number of apertures varied from target to target, but was typically ≈100.

2.3. MMT Optical Spectroscopy

Long-slit spectroscopic observations using the MMT Red Channel Spectrograph (Schmidt et al. 1989) were conducted in the 2012A semester (PI: R. S. Bussmann) and provided data on H-ATLAS J125632.7+233625, H-ATLAS J132630.1+334410, H-ATLAS J133008.4+245900, and H-ATLAS J134429.4+303036. The total on-source integration times for these targets were 70–120 minutes apiece. The observations were obtained during dark time in near-photometric conditions and seeing was typically 1farcs0–1farcs5. Details of the observations for each target are given in Table 3.

Table 3. Optical Spectroscopy Observations

IAU Name Telescope UT Date Grating λらむだcentral Slit Width Δでるたλらむだ tint
(lines mm−1) (Å) ('') (Å) (min)
J083051.0+013224 WHT 2011 Apr 24 400 6500 1.0 15 60
J090740.0−004200 Gemini-S 2012 Feb 26 400 6710 1.5 10.5 120
J114637.9−001132 X-Shooter 2012 Sep 18/19 ... 16200 1.0 3.0 320
J125632.7+233625 MMT 2012 Feb 23 300 5504 1.5 17.9 80
J132630.1+334410 MMT 2012 Feb 23 270 6996 2.0 21.9 120
J133008.4+245900 MMT 2012 Feb 23 300 5504 1.5 16.4 70
J134429.4+303036 MMT 2012 Feb 22 270 6996 1.5 17.9 80
J141351.9−000026 Gemini-S 2012 Feb 26 150 6720 1.5 22.3 240

Download table as:  ASCIITypeset image

We obtained lamp flats for all targets at the beginning of each night for flat fields. We obtained a sequence of up to six consecutive 10 minute exposures on each target. For wavelength calibration, before and after each of these sequences we obtained comparison lamp observations using a He/Ar/Ne comparison lamp. We checked the focus periodically throughout the night.

We used either the 270 or the 300 lines per mm grating with a central wavelength (λらむだcentral) of 6996 Å or 5504 Å. The slit width was chosen to be as small as atmospheric seeing allowed and was either 1farcs5 or 2farcs0. The spectral resolution at λらむだcentral (Δでるたλらむだ) was 10–30 Å.

The long-slit data were reduced using standard IRAF one-dimensional spectroscopy routines. The primary aim of these observations for the purpose of this paper is to obtain a spectroscopic redshift for the putative lensing system. We accomplished this task with the xcsao routine in IRAF, using as a template a 5 Gyr old simple stellar population from Bruzual & Charlot (2003) with solar metallicity and a Chabrier initial mass function (IMF). This template does not perfectly match the lensing galaxy spectra, but it is sufficient to determine a robust redshift. The lens redshifts are presented in Table 1. A more detailed exploration of the optical spectra is deferred to a subsequent publication, so we do not show the spectra in this paper.

2.4. Gemini-South Optical Spectroscopy

Long-slit spectroscopic observations using the Gemini Multi-Object Spectrograph-South (GMOS-S; Hook et al. 2004) were conducted in queue mode during the 2012A semester as part of program GS-2012A-Q-52 (PI: R. S. Bussmann) and provided data on H-ATLAS J090740.0−004200 and H-ATLAS J141351.9−000026. The total on-source integration times for each of these targets were 1–4 hr. The observations were obtained during dark time in near-photometric conditions. Observational details are given in Table 3.

Some aspects of the observing strategy were common to both targets and followed the guidelines given by the Gemini observatory.34 Flat field observations were interspersed between the science exposures at each wavelength setting. CuAr arc lamp exposures were taken for the purpose of wavelength calibration, using the same instrumental setup as for the science exposures. A 5 Å spectral dither between exposures was used to cover the gap in the GMOS-S chip, and we binned the CCD pixels by a factor of 4 in both the spatial and spectral directions, providing a spatial scale of 0farcs288 pixel−1 and a spectral pixel scale of 2.69 Å. We used the GG455 blocking filter. Aspects of the observing strategy that varied from target to target are given in Table 3.

The long-slit data were reduced using the IRAF Gemini GMOS reduction routines, following the standard GMOS-S reduction steps in the example taken from the Gemini observatory webpage.35 We used the same procedure as outlined in Section 2.3 to measure a spectroscopic redshift using IRAF's xcsao task. The lens redshifts are presented in Table 1. We plan a detailed exploration of the optical spectra in future work, so we do not show the optical spectra in this paper.

2.5. WHT Optical Spectroscopy

Long-slit spectroscopic observations using the auxiliary port camera (ACAM; Benn et al. 2008) at the WHT were conducted in the 2011A semester (PIs: I. Pérez-Fournon and A. Verma) and provided three 1200 s exposures of H-ATLAS J083051.0+013224 on 2011 April 24. ACAM provides fixed-format spectroscopy, covering a spectral range of 3500–9400 Å (spectral resolution of ≈3.3 Å pix−1) using a 400 lines mm−1 transmission Volume Phase Holographic grating. The observing conditions were photometric and the seeing was 0farcs75. We have included the observational details for this target in Table 3 for completeness.

We obtained Tungsten lamp flats at the beginning of each night for flat fields. For wavelength calibration, we obtained comparison lamp observations using a CuNe lamp. We checked the focus periodically throughout the night using the ACAM imaging mode.

A position angle of 114 deg east of north was chosen to include several objects visible on the optical imaging close to the candidate lensing galaxy. The slit width was 1farcs0 and the corresponding spectral resolution provided by the grating was R ≈ 450, or 15 Å FWHM at 6750 Å.

The long-slit spectroscopic data were reduced using standard IRAF two-dimensional spectroscopy programs to correct for the distortions in the data and apply the wavelength and flux density calibration.

The ACAM spectroscopy, based on G-band and Mg absorption features, shows that the primary lens (i.e., the galaxy that is closest to the SMA emission centroid) is located at zlens = 0.626 ± 0.001. A nearby disk-like galaxy was detected in the acquisition images and spatially resolved from the primary lens (the existence of this additional galaxy is confirmed by ancillary high-spatial resolution imaging from Keck-II adaptive optics imaging; see Section 2.7). In the ACAM long-slit spectra an emission line is detected at 7462 Å at the spatial location of the disk-like galaxy. We cannot associate this line with emission lines at the same redshift as the primary lens. The most plausible identification is [O ii]3727 at z = 1.002 ± 0.001, in which case the two galaxies are at different redshifts.

2.6. VLT X-Shooter Spectroscopy

The optical and near-IR spectra of the southeastern lens in J114637−001132 (i.e., the galaxy denoted as "G2" by Fu et al. 2012) were obtained with the X-shooter spectrograph (Vernet et al. 2011) at the VLT. X-shooter provides simultaneous spectral coverage from 300 nm to 2.5 μみゅーm. The 1'' × 11'' slit was used in the "UVB" arm (300–560 nm), while the 0farcs9 × 11'' slits were used in the "VIS" arm (550–1020 nm) and "NIR" arm (1020–2480 nm), yielding a resolving power of λらむだ/Δでるたλらむだ of 4350, 7450 and 5300 in the three bands, respectively. The observations were obtained on 2012 September 18 and 19. The total integration time was 320 minutes and comprised individual exposures of 20 minutes each. The object was nodded along the slit by a few arcsec between one exposure and the next.

Data reduction was performed by following the standard steps of the public X-shooter pipeline (Goldoni et al. 2006). Sky emission lines were subtracted by exploiting temporally contiguous exposures in which the objects was nodded in a different position of the slit. After flat-fielding, the pipeline extracts the different orders of the echelle spectrum, which are then rectified, wavelength calibrated and merged. Then the spectra were calibrated in flux by using the observation of a spectrophotometric standard. The final mono-dimensional spectrum was extracted from an aperture of 1''.

The lensing galaxy G2 is clearly detected with several emission lines ([O iii],Hb,Ha,[N ii],[S ii], [O ii]) that imply a redshift of zlens = 1.2247.

2.7. Ancillary Optical and Near-IR Imaging

In all cases where the SMA has clearly resolved multiple images of the background source, there is no evidence for submm emission from the lens. This means that detection of the foreground lens requires observations at optical or near-IR wavelengths. This paper makes use of the best available optical or near-IR imaging to pinpoint the location of the lens and determine whether it comprises multiple galaxies. This imaging is shown in grayscale in Figure 2, and the telescope and filter used are given in the lower left corner of each panel. Fifteen objects use HST Snapshot imaging (marked as "HST F110W" in Figure 2), five use Keck-II/NIRC2-LGSAO imaging (marked as "Keck-II_NIRC2 Ks" in Figure 2), five use full-orbit HST imaging (marked as "HST F160W" in Figure 2), four use SDSS i-band imaging (marked as "SDSS i" in Figure 2), and two use WHT Ks-band imaging (marked as "WHT Ks" in Figure 2).

The focus of this paper is lens modeling of the SMA data. A detailed analysis of the optical and near-IR imaging will appear in a set of papers specific to the HST Snapshot imaging (S. Amber et al., in preparation; J. Calanog et al. in preparation), the full-orbit HST imaging (M. Negrello et al., in preparation, S. Dye et al., in preparation), and the Keck-II/NIRC2-LGSAO imaging (J. Calanog et al., in preparation) and the WHT Ks imaging (P. Martínez-Navajas et al., in preparation).

To facilitate comparison with existing surveys for lenses based on SDSS spectroscopy (e.g., Bolton et al. 2008; Brownstein et al. 2012), we compute i-band photometry using SDSS Data Release 9 (DR9) for all of the objects in the SMA subsample. The imaging aspect of DR9 provides five optical bands: u, g, r, i, and z. The 95% completeness levels for point sources are u = 22.0, g = 22.2, r = 22.2, i = 21.3, and z = 20.5 (AB mag), corresponding to flux densities of 5.7 μみゅーJy, 4.8 μみゅーJy, 4.8 μみゅーJy, 11.0 μみゅーJy, and 23.0 μみゅーJy, respectively. The median seeing in the images at r-band is typically 1farcs3.

We searched for counterparts in the DR9 catalog within a 2'' radius of each expected lens position based on the best available optical or near-IR imaging. If a counterpart was found, then it was assigned photometry directly from the DR9 catalog. If no counterpart was found, we used our own custom aperture photometry code to measure the 2σしぐま limiting flux density at the position of the target (note that at these wavelengths, the lens is typically much brighter than the source). We used a 4'' diameter circular aperture and computed the sky background in an annulus with an inner radius of 2'' and an outer radius of 5''. We measured the uncertainties by placing N random apertures (where N ≈ 300) of the same size and shape within 3' of the lens candidate (taking care to avoid any objects found in the DR9 catalog) and computing the 68% confidence interval of the dispersion in the measured flux densities. The i-band AB magnitudes are reported in Table 4 (limits indicate 2σしぐま values), along with the Herschel/SPIRE and SMA 880 μみゅーm measurements (the values reported in the table do not include absolute flux density calibration uncertainty of 7%).

Table 4. Spatially Integrated Flux Densities of Strong Lens Samplea

IAU Name ib S250 S350 S500 S880
(AB mag) (mJy) (mJy) (mJy) (mJy)
J021830.5−053124 >22.6 92 ± 7 122 ± 8 113 ± 7 66.0 ± 5.4
J022016.5−060143 20.32 ± 0.06 180 ± 7 192 ± 8 132 ± 7 28.3 ± 3.4
J083051.0+013224 20.85 ± 0.09 260 ± 7 321 ± 8 269 ± 9 85.5 ± 4.0
J084933.4+021443 19.01 ± 0.02 242 ± 7 293 ± 8 231 ± 9 50.0 ± 3.5
J085358.9+015537 >22.3 389 ± 7 381 ± 8 241 ± 9 61.4 ± 2.9
J090302.9−014127 20.92 ± 0.11 347 ± 7 339 ± 8 219 ± 9 54.7 ± 3.1
J090311.6+003906 18.17 ± 0.01 138 ± 7 199 ± 8 174 ± 9 78.4 ± 8.2
J090740.0−004200 20.94 ± 0.07 471 ± 7 343 ± 8 181 ± 9 24.8 ± 3.3
J091043.1−000321 21.41 ± 0.09 417 ± 6 378 ± 7 232 ± 8 30.6 ± 2.4
J091305.0−005343 18.74 ± 0.02 116 ± 6 140 ± 7 108 ± 8 36.7 ± 3.9
J091840.8+023047 >22.4 142 ± 7 175 ± 8 138 ± 9 18.8 ± 1.6
J103826.6+581542 18.71 ± 0.02 191 ± 7 157 ± 10 101 ± 7 30.2 ± 2.2
J105712.2+565457 22.0 ± 0.4 114 ± 7 147 ± 10 114 ± 7 50.3 ± 5.9
J105750.9+573026 20.15 ± 0.04 403 ± 7 377 ± 10 249 ± 7 55.7 ± 5.8
J113526.3−014605 >22.5 290 ± 7 295 ± 8 216 ± 9 48.6 ± 2.3
J114637.9−001132 21.44 ± 0.10 290 ± 6 356 ± 7 295 ± 8 86.0 ± 4.9
J125135.4+261457 >22.2 145 ± 7 201 ± 8 212 ± 9 78.9 ± 4.4
J125632.7+233625 18.70 ± 0.02 214 ± 7 291 ± 8 261 ± 9 97.2 ± 6.5
J132427.0+284452 >22.6 347 ± 7 377 ± 8 268 ± 9 30.2 ± 2.2
J132630.1+334410 20.25 ± 0.07 179 ± 7 279 ± 8 265 ± 9 65.2 ± 2.3
J132859.3+292327 >22.6 264 ± 9 310 ± 10 261 ± 10 50.1 ± 2.1
J133008.4+245900 20.00 ± 0.03 273 ± 7 282 ± 8 214 ± 9 59.2 ± 4.3
J133649.9+291801 >22.7 295 ± 8 294 ± 9 191 ± 10 36.8 ± 2.9
J134429.4+303036 20.88 ± 0.06 481 ± 9 484 ± 13 344 ± 11 73.1 ± 2.4
J141351.9−000026 22.0 ± 0.2 190 ± 7 240 ± 8 200 ± 9 33.3 ± 2.6
J142413.9+022303 21.62 ± 0.12 115 ± 7 192 ± 8 203 ± 9 90.0 ± 5.0
J142823.9+352619 22.2 ± 0.4 323 ± 6 244 ± 7 140 ± 33 18.4 ± 2.5
J142825.5+345547 19.89 ± 0.04 159 ± 6 196 ± 7 157 ± 33 42.3 ± 4.7
J143330.8+345439 >22.3 158 ± 6 191 ± 7 160 ± 33 59.6 ± 3.9
J144556.1−004853 >22.7 141 ± 7 157 ± 8 130 ± 9 9.0 ± 2.1

Notes. aMeasurement uncertainties for Herschel photometry do not include absolute flux density calibration uncertainty of 7%. bi-band magnitudes obtained from SDSS DR9.

Download table as:  ASCIITypeset image

3. LENS MODELS

The SMA data provide sufficient sensitivity and spatial resolution to permit tight constraints on parameters of the lens models for a total of 25 lensed SMGs out of the SMA subsample of 30 (those labeled with grade A, B, or C in Table 1). For some of these objects, deep HST or Keck-II/NIRC2-LGSAO data exist that permit the assembly of lens models which take into account simultaneously the optical, near-IR, and submm data. However, because this sample of lensed SMGs are at z > 1.5 and are heavily obscured by dust, the lensed emission is frequently detected only in the SMA data and not in the optical or near-IR. Therefore, for the current analysis we focus our efforts on lens models based solely on SMA data (for the handful of exceptions, see Section 3.1 for details) and defer full spectral energy distribution (SED) lens modeling to subsequent publications. We describe the methodology behind the lens modeling in Section 3.1 and give a detailed discussion of each object in Section 3.2. We defer an examination of the ensemble properties of the lenses and lensed sources to Sections 4 and 5, respectively.

3.1. Methodology

The SMA is an interferometer, so the surface brightness map of each lensed SMG is obtained with incomplete sampling of the uv plane. This means that surface brightness is not necessarily conserved and that the pixel-to-pixel errors in the surface brightness map are correlated. For these reasons, it is important to compare model and data visibilities rather than surface brightness maps. We follow the methodology used in Bussmann et al. (2012), who presented the first lens model derived from a visibility-plane analysis of interferometric imaging of a strongly lensed SMG discovered in wide-field submm surveys. We summarize important details here and refer the interested reader to Bussmann et al. (2012) for further information.

We use the publicly available Gravlens software (Keeton 2001) to map emission from the source plane to the image plane for a given lensing mass distribution. To represent the lens mass profile, we use Nlens singular isothermal ellipsoid (SIE) profiles, where Nlens is the number of lensing galaxies found from the best available optical or near-IR imaging (a multitude of evidence supports the SIE as a reasonable choice; for a recent review, see Treu 2010).

The source(s) are assumed to have Sérsic profile morphologies (Sersic 1968). We always use a single Sérsic profile in our fits, with the exception of objects that are clearly only moderately lensed (i.e., singly imaged with μみゅー < 2) and that show evidence of multiple source-plane components in the SMA imaging. This is true for J022016.5−060143 and J084933.4+021443.

Each SIE is fully described by the following five free parameters: the position of the lens relative to the SMA emission centroid (Δでるたαあるふぁlens = αあるふぁlensαあるふぁ880 and Δでるたδでるたlens = δでるたlensδでるた880; these can be compared with the position of the optical or near-IR counterpart relative to the SMA emission centroid: ΔでるたαあるふぁNIR = αあるふぁNIRαあるふぁ880 and ΔでるたδでるたNIR = δでるたNIRδでるた880), the mass of the lens (parameterized in terms of the angular Einstein radius, θしーたE), the ellipticity of the lens (epsilonlens; defined as 1 − b/a), and the position angle of the lens (ϕlens; degrees east of north). When there is evidence for additional lenses from optical or near-IR imaging (see Figure 2), we estimate by-eye centroids for each lens (carrying an uncertainty of order 1 pixel, or 0farcs04 and 0farcs12 in the Keck-II/NIRC2-LGSAO and HST images, respectively) and fix the positions of the additional lenses with respect to the primary lens. Therefore, each additional lens has only 3 free parameters: θしーたE, epsilonlens, and ϕlens. We assume secondary, tertiary, etc., lenses are located at the same redshift as the primary lens, unless there is evidence against that assumption (e.g., J083051.0+013224).

Each Sérsic profile is fully described by the following seven free parameters: the position of the source relative to the primary lens (Δでるたαあるふぁs = αあるふぁsαあるふぁlens and Δでるたδでるたs = δでるたsδでるたlens), the intrinsic flux density (Sin), the Sérsic index (ns), the half-light semi-major axis length (as), the ellipticity (epsilons, defined as 1 − b/a), and the position angle (ϕs, degrees east of north).

The total number of free parameters for any given system is Nfree = 5 + 3 × (Nlens − 1) + 7*Nsource, where Nsource is the number of Sérsic profiles used.

We adopt loose, uniform priors for all model parameters. The 1-σしぐま absolute astrometric solution between the SMA and optical/near-IR images is generally 0farcs2, so in our modeling efforts the prior on the position of the lens covers ±0farcs6 in both R.A. and Decl. (i.e., 3σしぐま in each direction). In Section 3.2, we discuss the level of agreement between the astrometry from the images and the astrometry from the lens modeling. For θしーたE, the prior covers 0farcs1–6''. The ellipticities of the lens and source are always restricted to be <0.7. No prior is placed on the position angle of the lens or source. The intrinsic flux density is allowed to vary from 0.1 mJy to the total flux density observed by the SMA. The source position is allowed to vary by ±1'' relative to the position of the primary lens. The Sérsic index varies from 0.1 to 4.0. The half-light radius varies from 0farcs05 to 1farcs5.

For a given set of model parameters, Gravlens generates a surface brightness map of the lensed emission (note that no model of the emission from the lens is needed because the lenses are undetected in the SMA imaging). This surface brightness map can then be used as input to MIRIAD's uvmodel task, which produces a "simulated visibility" dataset (Vmodel) by computing the Fourier transform of the model lensed image and sampling the resulting visibilities to match the sampling of the actual observed SMA visibility dataset (VSMA). The quality of fit for a given set of model parameters is determined from the chi-squared value (χかい2) according to the following equations:

Equation (1)

Equation (2)

Equation (3)

where σしぐまreal(u, v) = σしぐまimag(u, v) is the 1σしぐま uncertainty level for each visibility and is determined from the system temperatures (this corresponds to a natural weighting scheme).

To sample the posterior probability density function (PDF) of our model parameters, we use emcee (Foreman-Mackey et al. 2013), a Markov chain Monte Carlo (MCMC) code that uses an affine-invariant ensemble sampler to obtain significant performance advantages over standard MCMC sampling methods (Goodman & Weare 2010).

We employ a "burn-in" phase with 250 walkers and 1000 iterations (i.e., 250,000 samplings of the posterior PDF) to identify the best-fit model parameters. This position is then used to initialize the "final" phase with 250 walkers and 20 iterations (i.e., 5,000 samplings of the posterior PDF) to determine uncertainties on the best-fit model parameters. The autocorrelation time for each parameter in a given ensemble of walkers and is of order unity for each parameter, implying that we have 5,000 independent samplings of the posterior PDF, more than enough to obtain a robust measurement of the mean and uncertainty on each parameter of the model.

During each MCMC iteration, we also measure the magnification factor at 880 μみゅーm, μみゅー880 (we follow the nomenclature in the SMG literature here and use μみゅー to refer to the total magnification obtained by summing over all individual lensed components). Here, we describe how we measure μみゅー880.

First, we take the unlensed, intrinsic source model and measure the total flux density (Sin) within an elliptical aperture (Ain) centered on the source with ellipticity and position angle equal to that of the source model and with a semi-major axis length of 2as. Second, we take the lensed image of the best-fit model and measure the total flux density (Sout) within the aperture Aout, where Aout is determined by using Gravlens to map Ain in the source plane to Aout in the image plane (using the lens parameters which correspond to the best-fit model). The magnification is then computed simply as μみゅー880 = Sout/Sin. The best-fit value and 1σしぐま uncertainty are drawn from the posterior PDF, as with the other parameters of the model.

The choice of Ain has important implications for magnification measurements. For multiply imaged systems, apertures that are too large include in the source plane too much flux density that is far away from the caustic and relatively unmagnified. This is a particularly important issue for the models used here because the Sérsic index of the background source is a free parameter. The Sérsic index is partially degenerate with the half-light radius of the source, in the sense that good fits to the data can be obtained with a combination of small source size and small Sérsic index or large source size and large Sérsic index. In accordance with this, our model fits sometimes include relatively large sources where significant fractions of the unlensed flux density (≈10%–20%) arise from regions in the source plane far away from the caustic and hence contribute nothing to the observed lensed emission. This situation biases the estimate of μみゅー880 below the true value. Conversely, apertures that are too small will omit flux density in the source plane that is detected at high significance in the SMA imaging, thus biasing the estimate of μみゅー880 above the true value. Our choice—double the semi-major axis length of the source—represents a compromise between these two extremes.

3.2. Descriptions of Individual Objects

In this section, we describe the basic characteristics of each object in the SMA subsample, including the position of the lens relative to the SMA 880 μみゅーm emission centroid, the lensing configuration (where applicable), and any unique notes for each object. Figure 3 shows the best-fit model in comparison with the SMA data for every lensed SMG with a robust lens model. Tables 5 and 6 present, for lenses and sources respectively, the model parameter mean values and 1σしぐま uncertainties as drawn from the posterior PDF for each parameter. Note that in some cases the posterior PDFs are non-Gaussian and therefore the best-fit model shown in Figure 3 does not always correspond perfectly to the model parameter mean values presented in Tables 5 and 6.

Figure 3.
Standard image High-resolution image
Figure 3.
Standard image High-resolution image
Figure 3.

Figure 3. Comparison of best-fit lens models with SMA data. Odd columns show the surface brightness maps of the best-fit model lensed emission (grayscale) in comparison with the observed surface brightness maps from the SMA (red contours, beginning at ±2σしぐま and increasing by factors of $\sqrt{2}$). Even columns show the residual surface brightness maps obtained by subtracting the best-fit lens model visibilities with the observed visibilities by the SMA (contours drawn at same levels as odd panels). For reference, all panels show the critical curves (orange line), caustics (cyan lines), position of the lens(es) (black circles), the half-light area of the background source(s) (magenta filled ellipses), and the FWHM of the SMA synthesized beam (black hatched ellipses).

Standard image High-resolution image

Table 5. Gravitational Lens Model Results: Lens Properties

Object ΔでるたαあるふぁNIR ΔでるたδでるたNIR Δでるたαあるふぁlens Δでるたδでるたlens θしーたE epsilonlens ϕlens χかい2 NDOF
('') ('') ('') ('') ('') (deg)
J021830.5−053124 −0.17 −0.02 0.09 ± 0.04 −0.13 ± 0.05 0.44 ± 0.02 0.35 ± 0.10 156 ± 18 106027.4 114284
J022016.5−060143 2.1 0.6 2.07 ± 0.01 0.63 ± 0.01 0.27 ± 0.13 0.29 ± 0.13 114 ± 52 131464.2 133819
... −4.76 0.3 −4.76 0.3 0.34 ± 0.14 0.30 ± 0.18 53 ± 54 ... ...
J083051.0+013224 0.18 0.10 0.20 ± 0.01 0.00 ± 0.02 0.39 ± 0.02 0.43 ± 0.05 123 ± 3 97761.3 100561
... 0.54 0.595 0.54 0.595 0.43 ± 0.02 0.25 ± 0.07 47 ± 9 ... ...
J084933.4+021443 −7.90 −0.60 −7.90 −0.60 1.41 ± 0.04 0.11 ± 0.06 62 ± 22 188965.5 168376
J085358.9+015537 0.04 0.02 0.09 ± 0.03 −0.03 ± 0.01 0.553 ± 0.004 0.06 ± 0.02 70 ± 12 160445.4 161200
J090302.9−014127 0.20 0.16 0.092 ± 0.006 0.03 ± 0.02 0.33 ± 0.02 0.39 ± 0.07 83 ± 5 124529.4 124378
J090311.6+003906 0.07 0.10 0.13 ± 0.02 0.03 ± 0.05 1.52 ± 0.03 0.34 ± 0.05 179 ± 4 198305.2 188164
J090740.0−004200 −0.03 0.17 0.01 ± 0.04 0.03 ± 0.04 0.59 ± 0.04 0.50 ± 0.08 44 ± 5 98244.2 91314
J091043.1−000321 0.22 −0.14 0.05 ± 0.02 −0.18 ± 0.01 0.95 ± 0.02 0.08 ± 0.03 152 ± 24 171966.9 159664
J091305.0−005343 −0.32 0.35 −0.45 ± 0.05 0.38 ± 0.08 0.43 ± 0.07 0.51 ± 0.13 43 ± 14 211928.5 200412
J103826.6+581542 −1.80 0.25 −1.1 ± 0.2 0.27 ± 0.09 2.0 ± 0.2 0.54 ± 0.11 18 ± 3 78239.2 76704
J105750.9+573026 −0.65 −1.19 −0.11 ± 0.05 −0.80 ± 0.08 3.86 ± 0.01 0.52 ± 0.01 14 ± 1 142740.4 120764
... 1.7 3.4 1.7 3.4 0.12 ± 0.01 0.54 ± 0.11 170 ± 9 ... ...
J114637.9−001132 −0.40 0.02 −0.59 ± 0.02 0.02 ± 0.02 0.65 ± 0.02 0.26 ± 0.04 114 ± 19 90621.1 94863
... 1.154 1.270 1.154 1.270 0.67 ± 0.01 0.50 ± 0.04 68 ± 2 ... ...
... −3.076 −1.688 −3.076 −1.688 0.61 ± 0.03 0.33 ± 0.07 95 ± 12 ... ...
... −4.340 0.632 −4.340 0.632 0.51 ± 0.04 0.50 ± 0.10 77 ± 6 ... ...
J125135.4+261457 0.04 0.13 0.05 ± 0.04 −0.25 ± 0.09 1.02 ± 0.03 0.46 ± 0.06 122 ± 1 72453.5 62912
J125632.7+233625 0.11 0.25 0.00 ± 0.01 0.25 ± 0.01 0.68 ± 0.01 0.31 ± 0.03 123 ± 1 62465.3 45880
J132427.0+284452 3.09 5.22 3.09 5.22 1.7 ± 0.4 0.34 ± 0.14 81 ± 16 228617.0 213957
... −7.51 −9.66 −7.51 −9.66 2.2 ± 0.3 0.14 ± 0.09 88 ± 15 ... ...
J132630.1+334410 0.68 −0.78 0.53 ± 0.05 −0.54 ± 0.05 1.80 ± 0.02 0.26 ± 0.04 66 ± 4 43605.2 38964
J133008.4+245900 −0.24 −0.02 −0.14 ± 0.03 −0.13 ± 0.02 0.88 ± 0.02 0.52 ± 0.03 81 ± 1 71948.3 52744
J133649.9+291801 ... ... 0.00 ± 0.11 0.03 ± 0.16 0.40 ± 0.03 0.38 ± 0.14 120 ± 13 100742.9 95844
J134429.4+303036 −0.03 −0.10 −0.03 ± 0.02 −0.01 ± 0.02 0.92 ± 0.02 0.39 ± 0.06 172 ± 14 182936.2 171864
J141351.9−000026 −0.32 −2.65 −0.32 ± 0.03 −2.50 ± 0.03 1.13 ± 0.09 0.31 ± 0.12 118 ± 33 132668.6 123420
J142413.9+022303 −0.11 0.79 −0.27 ± 0.03 0.63 ± 0.03 0.57 ± 0.01 0.29 ± 0.01 62 ± 1 108899.5 144191
... 0.025 −0.327 0.025 −0.327 0.40 ± 0.01 0.06 ± 0.02 133 ± 14 ... ...
J142823.9+352619 −0.14 −0.02 −0.01 ± 0.10 0.00 ± 0.11 0.10 ± 0.03 0.36 ± 0.18 87 ± 40 17284.0 17036
J142825.5+345547 0.25 −0.05 −0.20 ± 0.03 −0.04 ± 0.03 0.77 ± 0.03 0.46 ± 0.06 56 ± 5 93163.7 75448
J143330.8+345439 0.00 0.05 −0.07 ± 0.03 0.07 ± 0.02 0.28 ± 0.02 0.59 ± 0.08 104 ± 7 91554.9 117732

Download table as:  ASCIITypeset image

Table 6. Gravitational Lens Model Results: Source Properties

IAU Name Δでるたαあるふぁs Δでるたδでるたs ns as epsilons ϕs μみゅー880 μみゅーm
('') ('') ('') (deg)
J021830.5−053124 −0.08 ± 0.03 0.22 ± 0.04 2.1 ± 0.9 0.33 ± 0.12 0.29 ± 0.13 82 ± 22 4.4 ± 1.0
J022016.5−060143 −1.54 ± 0.02 0.65 ± 0.02 2.1 ± 0.8 0.16 ± 0.05 0.25 ± 0.10 68 ± 34 1.5 ± 0.3
... −2.56 ± 0.02 −0.82 ± 0.07 1.7 ± 0.8 0.30 ± 0.07 0.40 ± 0.13 45 ± 53 1.2 ± 0.1
... −2.61 ± 0.02 −2.05 ± 0.02 1.4 ± 0.7 0.28 ± 0.06 0.19 ± 0.09 72 ± 44 1.2 ± 0.0
J083051.0+013224 −0.25 ± 0.01 0.08 ± 0.02 1.8 ± 0.3 0.14 ± 0.01 0.32 ± 0.06 20 ± 7 6.9 ± 0.6
J084933.4+021443 0.67 ± 0.03 −0.73 ± 0.03 2.1 ± 0.7 0.15 ± 0.03 0.20 ± 0.12 103 ± 35 2.8 ± 0.2
... 11.3 3.0 0.4 ± 0.3 0.25 ± 0.02 0.33 ± 0.10 24 ± 18 1.0
... 14.3 4.0 2.0 ± 0.9 0.17 ± 0.07 0.43 ± 0.17 93 ± 34 1.0
J085358.9+015537 −0.09 ± 0.03 0.001 ± 0.002 2.0 ± 0.7 0.06 ± 0.01 0.33 ± 0.14 83 ± 17 15.3 ± 3.5
J090302.9−014127 −0.06 ± 0.01 −0.01 ± 0.01 2.5 ± 0.6 0.42 ± 0.12 0.22 ± 0.12 116 ± 23 4.9 ± 0.7
J090311.6+003906 −0.26 ± 0.03 0.04 ± 0.05 2.2 ± 0.4 0.52 ± 0.10 0.35 ± 0.06 94 ± 11 11.1 ± 1.1
J090740.0−004200 −0.15 ± 0.03 0.14 ± 0.03 2.0 ± 1.1 0.16 ± 0.10 0.31 ± 0.14 73 ± 57 8.8 ± 2.2
J091043.1−000321 −0.017 ± 0.008 0.23 ± 0.02 1.5 ± 0.5 0.13 ± 0.02 0.35 ± 0.12 5 ± 20 10.9 ± 1.3
J091305.0−005343 0.42 ± 0.05 −0.37 ± 0.06 3.0 ± 0.6 0.76 ± 0.12 0.56 ± 0.13 129 ± 10 2.1 ± 0.3
J103826.6+581542 1.1 ± 0.2 −0.33 ± 0.06 3.0 ± 0.7 0.45 ± 0.18 0.44 ± 0.19 129 ± 32 7.1 ± 1.5
J105750.9+573026 −0.07 ± 0.04 0.76 ± 0.06 2.4 ± 0.8 0.57 ± 0.08 0.58 ± 0.11 122 ± 6 9.2 ± 0.4
J114637.9−001132 −0.50 ± 0.06 −0.28 ± 0.04 0.5a 0.38 ± 0.03 0.77 ± 0.03 107 ± 4 9.5 ± 0.6
J125135.4+261457 0.02 ± 0.04 0.22 ± 0.09 1.9 ± 0.6 0.15 ± 0.03 0.22 ± 0.10 109 ± 32 11.0 ± 1.0
J125632.7+233625 0.014 ± 0.006 −0.12 ± 0.01 1.1 ± 0.6 0.07 ± 0.01 0.37 ± 0.14 140 ± 21 11.3 ± 1.7
J132427.0+284452 −3.8 ± 0.4 −5.1 ± 0.6 2.4 ± 0.4 0.72 ± 0.09 0.69 ± 0.01 169 ± 17 2.8 ± 0.4
J132630.1+334410 −0.60 ± 0.03 0.60 ± 0.03 1.1 ± 0.3 0.22 ± 0.02 0.18 ± 0.09 150 ± 15 4.1 ± 0.3
J133008.4+245900 0.05 ± 0.01 0.23 ± 0.02 1.6 ± 0.7 0.09 ± 0.03 0.37 ± 0.10 129 ± 29 13.0 ± 1.5
J133649.9+291801 −0.04 ± 0.09 −0.05 ± 0.15 1.2 ± 0.5 0.19 ± 0.03 0.43 ± 0.12 125 ± 13 4.4 ± 0.8
J134429.4+303036 0.22 ± 0.02 0.04 ± 0.02 1.9 ± 0.5 0.24 ± 0.06 0.39 ± 0.07 100 ± 15 11.7 ± 0.9
J141351.9−000026 0.13 ± 0.13 1.40 ± 0.09 1.5 ± 0.5 0.30 ± 0.04 0.48 ± 0.12 62 ± 7 1.8 ± 0.3
J142413.9+022303 −0.24 ± 0.03 −0.53 ± 0.03 2.9 ± 0.3 0.64 ± 0.07 0.27 ± 0.09 77 ± 12 4.6 ± 0.5
J142823.9+352619 −0.00 ± 0.07 0.03 ± 0.08 1.9 ± 1.2 0.10 ± 0.07 0.33 ± 0.18 64 ± 62 3.0 ± 1.5
J142825.5+345547 0.05 ± 0.01 0.15 ± 0.03 0.7 ± 0.4 0.16 ± 0.04 0.50 ± 0.09 49 ± 10 10.3 ± 1.7
J143330.8+345439 −0.08 ± 0.02 −0.07 ± 0.02 1.9 ± 0.7 0.31 ± 0.06 0.55 ± 0.14 116 ± 13 4.5 ± 0.4

Note. ans = 0.5 was assumed for this source.

Download table as:  ASCIITypeset image

J021830.5−053124. SMA data from the compact and extended array configurations were originally presented in Ikarashi et al. (2011) and Wardlow et al. (2013). This paper presents new data obtained the in the very extended array, permitting the first resolved measurement of the 880 μみゅーm emission from this object. The best-fit lens model finds a position for the lens that is offset relative to that indicated in the HST image by 0farcs26 in R.A. and −0farcs11 in Decl. These values are near the expected level of 0farcs2–0farcs3 absolute astrometric uncertainty between the SMA and HST reference frames. The lensed emission is barely resolved by the SMA due to the small Einstein radius of the lens. The bulk of the source-plane emission originates outside the tangential caustic, favoring a two-image rather than four-image configuration (excluding the central demagnified image, which is never detected at the sensitivity levels probed in our data). This is consistent with the best-fit magnification factor of μみゅー880 = 4.4 ± 1.0.

J022016.5−060143. This object is the subject of a detailed study by Fu et al. (2013). We present it here mostly for completeness, but also to test the validity of the model in Fu et al. (2013) using the visibility-plane lens modeling technique described in this paper. We use a two-component lens model and a three-component source model to reproduce the observed 880 μみゅーm emission. We enforce a minimum Einstein radius of θしーたE > 0farcs2 for both lenses (corresponding to a minimum mass of Mlens > 1.6 × 1010M). This is intended to reproduce the prior on the lens masses used by Fu et al. (2013) based on the lens stellar masses and an assumed relation between the stellar mass and dark matter halo mass. The best-fit parameters of this model are statistically consistent with those found by Fu et al. (2013), despite a very different approach in modeling the data. In particular, we find a modest magnification factor is appropriate for each of the three components, with an average total magnification factor of 〈μみゅー880 μみゅーm〉 = 1.3 ± 0.1.

J083051.0+013224. The lensed emission has an unusual configuration that reflects the complexity of the foreground mass distribution due to the presence of a secondary lensing galaxy <1'' northwest of the primary lensing galaxy. WHT/ACAM spectroscopy shows that the primary lens is located at zlens = 0.626 based on G-band and Mg absorption features, while the secondary lens has very faint continuum and an emission line at observed frame 7462 Å. If this feature is [O ii]3727, then its redshift is zlens2 = 1.002. The best-fit lens model for this system assigns each of the foreground galaxies approximately equal Einstein radii (≈0farcs4). Given the stated lens and source redshifts, this implies lens masses of ≈0.7 × 1011M and 1.6 × 1011M, respectively.

J084933.4+021443. This object is the subject of a detailed study by Ivison et al. (2013). It is presented here for completeness and to ensure that the use of a visibility-plane lens model provides the same results as given in Ivison et al. (2013). Note that the center and extent of the image cutout shown in Figure 3 has been adjusted from the SMA emission centroid used in Figure 2 to include only the lens and the lensed source (dubbed "T" in Ivison et al. 2013). This is done simply to facilitate the comparison of model and data. An additional two sources used in the fitting process are not shown in this diagram. These two unlensed sources are modeled using the same visibility method but assuming no magnification by foreground objects. We find evidence for a larger magnification factor for "T" than Ivison et al. (2013): μみゅー880 = 2.8 ± 0.2 instead of μみゅー880 = 1.5 ± 0.2. A 2σしぐま emission peak located just northwest of the lens can be seen in Figure 3. If real, this emission peak supports the notion of a higher magnification factor for this object. Overall, however, our results are in broad agreement with those of Ivison et al. (2013).

J085358.9+015537. Although the image separations are small for this object (θしーたE = 0farcs553 ± 0farcs004), the S/N is high. The source appears to lie very close to the caustic (which is itself small due to the low ellipticity of the lens), implying a high magnification factor μみゅー880 = 15.3 ± 3.5.

J090302.9−014127. The image separations are close to the smallest values found in the SMA subsample (θしーたE = 0farcs35 ± 0farcs02). There has been tentative evidence (3σしぐま) of CO(J = 5–4) emission from the lens at zlens = 0.942 ± 0.004 (Lupu et al. 2012), but this has not been confirmed with subsequent, more sensitive observations that rule out lens redshifts from 0.922 to 0.944 (Omont et al. 2011). In the lens model presented here, the lens is assumed to be an insignificant submm emitter.

J090311.6+003906. Because of the high S/N and well-separated images of the background source (θしーたE = 1farcs52 ± 0farcs03), the lens model is well-constrained for this object. However, the image of the residual visibilities shows emission at the ±3σしぐま level, possibly an indication of complexity in the source structure that is not captured by a single Sérsic profile.

J090740.0−004200. The counter image to the northwest of the lens is detected at the 4σしぐま level and provides good constraints on the lens model for this object.

J091043.1−000321. This is one of a handful of objects with emission at the >±2σしぐま level that can be seen in the surface brightness map made from the residual visibilities. An edge-on galaxy located 4farcs4 to the northwest with a position angle of 135° east of north could be responsible for an external shear that has not been included in the lens model. Alternatively, the residual flux density may reflect a more complicated source structure than can be represented by our choice of a single Sérsic profile.

J091305.0−005343. The best-fit lens models for this object are obtained when the source is relatively large and most of it is located outside the region in the source plane that produces multiple images. For this reason, the best-fit magnification factor is relatively low (μみゅー880 = 2.1 ± 0.3).

J091840.8+023047. This object is unique in having no counterpart within 1'' of the 880 μみゅーm centroid in the HST/F110W Snapshot imaging. There is no obvious morphological signature of lensing based on the SMA data, so this may be a rare unlensed SMG. It is not included in Figure 3 since no lens model is available. Further investigation is needed to determine the nature of this object.

J103826.6+581542. This object was originally presented by Wardlow et al. (2013). We present it here for completeness and to test the validity of the model in Wardlow et al. (2013) using the visibility-plane lens modeling technique described in this paper. Our results for the size (0farcs45 ± 0farcs18 versus <0farcs5) and magnification factor of the background source (7.1 ± 1.5 versus $5.32^{+1.28}_{-1.06}$) are consistent with those reported in Wardlow et al. (2013). Since a redshift measurement for the background source remains unavailable, no further analysis is possible for this object.

J105712.2+565457. This object was originally presented by Wardlow et al. (2013). We present a slightly modified reduction of this object here. Instead of MIRIAD's mossdi task, which was used in Wardlow et al. (2013) to image this object, here we have shifted all of the visibility datasets to have the same phase center and then used MIRIAD's clean task. This resulted in slightly improved spatial resolution (0farcs68 × 0farcs57 versus 1farcs18 × 0farcs97), that is still not sufficient to distinguish cleanly separated images from a single lensed source. It is not included in Figure 3 since no lens model is available. Further investigation is needed to determine the nature of this object.

J105750.9+573026. This object was originally analyzed in a series of papers reporting its discovery, interstellar medium (ISM) properties, gas dynamics, and (on the basis of Kp-band imaging) lensing geometry (Conley et al. 2011; Scott et al. 2011; Riechers et al. 2011b; Gavazzi et al. 2011). We present new very extended array data here and compare the SMA and Keck lens models. The very extended array data are not as sensitive as the compact array data, so natural weighting provides a beam size of 0farcs99 × 0farcs94. This is insufficient to resolve the individual images of the lensed source seen in the image from the compact array only data and is further indication that the lens model is primarily constrained by the compact array data. Because the S/N and spatial resolution in the Keck image are superior to those of the SMA image, we fix the parameters of the lens to match those of the model found by Gavazzi et al. (2011). We find a similar magnification factor at 880 μみゅーm compared to Kp (9.2 ± 0.4 versus 10.9 ± 0.7). The model has difficulty reproducing the locations of the images seen in the SMA data. In addition, we find an offset in the position of the lens of 0farcs54 in R.A. and 0farcs4 in Decl. This represents a ≈2–3σしぐま discrepancy in R.A. and may be an indication that some of the assumptions in our model are over-simplifications.

J113526.3−014605. There is no counterpart detected in a 15-minute Ks integration with the WHT. The SMA 880 μみゅーm image is clearly resolved, but does not show individual, well-separated images of a lensed SMG. It is not included in Figure 3 since no lens model is available. Further investigation is required to determine whether strong lensing is occurring in this object. The main avenues for progress are higher-spatial resolution submm imaging (e.g., SMA very extended array) and deeper observations in the optical or near-IR to determine whether or not there is a lensing galaxy.

J114637.9−001132. This object was originally presented by Fu et al. (2012). Here, we present new SMA extended array data that resolve the lensed emission into five striking images of the SMG. The position and morphology of the lensed SMG in the source plane reported here are statistically consistent with those found by Fu et al. (2012). The presence of flux density at the 3σしぐま level in the map of the residual visibilities (see Figure 3) indicates our simple assumptions about the lens mass model (singular isothermal ellipsoids at the locations of galaxies identified in the Keck-II/NIRC2-LGSAO and HST/F110W imaging) and/or the background source Sérsic profile may be breaking down.

J125135.4+261457. The offset between the position of the lens from the lens model and from the WHT astrometry is 0farcs01 in R.A. and 0farcs38 in Decl. compared to the best-fit parameters from the lens model. The offset in Decl. is larger than expected given the astrometric uncertainty in aligning the SMA and WHT reference frames. On the other hand, the lens model correctly predicts the ellipticity and position angle of the lens potential (no priors were assumed for the shape of the lens potential), a strong indication that the lens model is robust.

J125632.7+233625. The S/N is very high in this object and the images of the lensed SMG are well-separated, making the parameters of the lens model very robust.

J132427.0+284452. This object is the subject of detailed studies by (George et al. 2013) and H. Fu et al. (in preparation) to explore its dust, gas, and stellar properties. In this paper, we use the Keck-II/NIRC2-LGSAO image to constrain the positions of the lenses and apply our visibility-plane lens modeling technique to the SMA data. We find that a two-lens mass model is needed to reproduce the observed data. These correspond directly to two galaxies detected at high significance in the Keck imaging. However, it should be noted that there are two nearby clusters detected in the Red-Sequence Cluster (RSC) survey (RCS J132427+2845.2 at z = 0.997 ± 0.017 and RCS J132419+2844.7 at z = 0.802 ± 0.018 Gladders & Yee 2005). The centers of these clusters are uncertain due to a lack of X-ray data and no clear brightest-cluster galaxy, but the RCS surface density map suggests that RCS J132427+2845.2 lies only 10'' away from lensed SMG. Our lens modeling here does not account for the presence of this cluster. Furthermore, the background source is not multiply imaged in the SMA data, so the constraints on the lens parameters are weak. The possibility of counter images falling outside of the SMA primary beam (FWHM of 36'' at 880 μみゅーm) is low due to the lack of such a counterpart in the SPIRE maps.

J132630.1+334410. Two well-separated images of the lensed SMG are obvious in the SMA data, indicating the source is strongly lensed but is not among the highest magnification sources. The lens modeling result is consistent with this (μみゅー880 = 4.1 ± 0.3), suggesting a robust model fit.

J132859.3+292317. The nature of this object is unclear based on existing data. It is well-detected and clearly spatially resolved by the SMA at 880 μみゅーm. The problem is the lack of optical or near-IR imaging at a depth beyond what is achieved in SDSS, where the object is undetected in all bands. One plausible interpretation is that the object is lensed by a relatively low-mass foreground galaxy at intermediate redshift so that it is undetected in the SDSS images. Alternatively, it is conceivable that the object is not lensed and has an intrinsic submm flux density of S880 = 51.8 ± 2.0 mJy. We consider this latter option unlikely since there are no known SMGs with intrinsic submm flux densities that high. However, without additional data to confirm this intuition, we do not consider this object further in our analysis. It is not included in Figure 3 since no lens model is available.

J133008.4+245900. Multiple, well-separated images are detected in the SMA data, consistent with the relatively large inferred magnification factor from the lens model (μみゅー880 = 13.0 ± 1.5), suggesting a robust lens model has been obtained.

J133649.9+291801. This object is similar to J132859.3+292317 in that there is no significant detection in any of the SDSS optical bands, nor is deeper optical or near-IR imaging available. However, the SMA morphology provides evidence typical of strong lensing. Here, we assume that the object is strongly lensed.

J134429.4+303036. The lensed images are well-separated and well-detected in the SMA data. In fact, the S/N is so high that the map of the residual visibilities reveals emission at the ±3σしぐま level, likely indicating that some of our model assumptions are over-simplifications. Nevertheless, the model captures the vast majority of the SMA emission and therefore provides a fair representation of intrinsic source size and luminosity. This object is similar to J125135.4+261457 in that the lens model successfully predicts the ellipticity and position angle of the lens potential without any non-standard priors placed on these parameters.

J141351.9−000026. No counter image of this target is detected in the SMA data, an immediate indication that this object is not strongly lensed. The possibility of counter images falling outside of the SMA primary beam (FWHM of 36'' at 880 μみゅーm) is low due to the lack of such a counterpart in the SPIRE maps. The lens modeling confirms this, with μみゅー880 = 1.8 ± 0.3. The lens is located 7farcs7 northeast of a brightest cluster galaxy (BCG). The lens redshift used here is from the BCG, and in order to derive the mass of the lens from its θしーたE value, we have assumed that it is located at the same redshift as the BCG.

J142413.9+022303. This object is the subject of a detailed study by Bussmann et al. (2012). We do not reproduce the results of the previous work here because it used the same visibility-plane lens modeling technique outlined in this paper.

J142823.9+352619. This object was originally discovered in Spitzer mid-IR imaging of the Boötes Field (Borys et al. 2006) and has since been the subject of a great many follow-up observations, a thorough summary of which may be found in Wardlow et al. (2013). We present new SMA extended array data which do not resolve the source, further corroborating the idea that this object is very small and is not strongly lensed (i.e., μみゅー < 2). Since we know that there is an intervening galaxy along the line of sight to the Herschel source, we impose a minimum Einstein radius for the lens of θしーたE > 0farcs1 (corresponding to a minimum lens mass of Mlens > 1010M). The constraints on the lens model are weak. One of the few robust claims we can make regarding this source is that it is very small (as < 0farcs2).

J142825.5+345547. This object is the subject of a detailed study by J. L. Wardlow et al. (in preparation). The lens model suggests the lens is located 0farcs20 west of the SMA emission centroid, but the peak of the edge-on spiral seen in the astrometrically-aligned HST image occurs 0farcs25 east of the SMA emission centroid. This is a difference of 0farcs45 in R.A. (the difference in Decl. is insignificant) and is larger than the expected 1σしぐま astrometric uncertainty of 0farcs2. However, estimating the peak position of the lensing galaxy is difficult because it is nearly exactly edge-on with a prominent dust lane and very little bulge. Moreover, the lensed images are well-resolved and well-detected in the SMA data, and the lens model correctly predicts the position angle and approximate ellipticity of the lens without any priors on these parameters. For these reasons, we consider the lens model for this object to be robust.

J143330.8+345439. This object was originally presented in Wardlow et al. (2013). We present a slightly modified reduction of this object here in which we shift all of the visibility datasets to have the same phase center and then use MIRIAD's clean task. This provides slightly improved spatial resolution that is still insufficient to identify clearly separated images of a single lensed source. Nevertheless, we use the knowledge that spectroscopic redshifts are available for both the lens and background source to infer that strong lensing is occurring. We then model this object using the same set of model parameters as applied to the other objects in the SMA subsample. Note that absolute astrometric calibration based on alignment to existing ground-based imaging is not available in the HST image for this object, but the lens model still finds a position for the lens that is in reasonable agreement (within 0farcs4) with that in the HST image.

J144556.1−004853. No lens or source redshift is available for this object. The SMA data resolve the 880 μみゅーm emission, but detect the source at relatively low S/N (S/N <6). Keck-II/NIRC2-LGSAO Ks-band imaging detects a counterpart that is located between the resolved components identified in the SMA data. This could be a detection of the lens, or it could simply be an intrinsic component of an unlensed SMG. Unlike J132859.3+292317, the total 880 μみゅーm flux density (S880 = 9.0 ± 2.1 mJy) is not unprecedented for unlensed SMGs. It is not included in Figure 3 since no lens model is available. Further investigation is required to determine if strong lensing is occurring in this object.

4. PROPERTIES OF LENSES DISCOVERED BY HERSCHEL

Two of the most basic properties of strong gravitational lensing galaxies are their i-band magnitudes (tracing the stellar light emitted by the lens, since the background sources detected by Herschel are dust-obscured and high-redshift, and therefore faint in the optical) and Einstein radii. The left panel of Figure 4 shows these values for the objects in the SMA subsample with robust lens models and compares them to other strong lenses found in SLACS (Bolton et al. 2008), BELLS (Brownstein et al. 2012), CASTLeS (Muñoz et al. 1998), and CLASS (Myers et al. 2003; Browne et al. 2003).

Figure 4.

Figure 4. Properties of lenses discovered by Herschel (red circles) and SPT (cyan stars), compared with a compilation of lenses from BELLS and SLACS (gray squares), SL2S (magenta squares), CASTLES (blue squares), CLASS (green squares), and SQLS (yellow squares). Left: Einstein radius (θしーたE) as a function of i-band AB magnitude. The Herschel sample is fainter and shows a wider range in θしーたE values than any of the previous samples of lenses. Right: mass enclosed within θしーたE as a function of lens redshift. Herschel has identified lenses that are lower in mass or higher in redshift than any of the optically-based searches (SLACS, BELLS, and SL2S). The range in parameter space occupied by the Herschel data points is comparable to that of CASTLES, CLASS, and SQLS, but Herschel promises to provide a sample size that is over an order of magnitude larger (González-Nuevo et al. 2012).

Standard image High-resolution image

In this figure, the i-band magnitudes for the SMA subsample come from SDSS DR9 (see Section 2.7 for details). We account for multiple lens systems by assigning a fraction of the total SDSS i-band flux density to each lens. The appropriate fraction is determined from the ratio of the θしーたE for that lens to the sum of the θしーたE values for that system. This explains why a few objects appear in the left panel of Figure 4 to have i-band magnitudes below the SDSS limit of i ≈ 22.5 (AB mag).

Figure 4 helps clarify the observational distinctions between lenses discovered via wide-field (sub)mm surveys (i.e., Herschel and SPT) and optically-selected surveys like SLACS, BELLS, SL2S, and SQLS. Note that the CASTLeS and CLASS samples of lenses have similar properties to the Herschel and SPT lenses, but will not grow further in size. Although BELLS and SL2S go much deeper in the i-band than SLACS, they are still biased toward brighter lenses than the source-selected surveys by the need for detections in SDSS optical spectroscopy. Even SQLS, which does not require optical spectroscopic detections from SDSS, has lenses with brighter i-band magnitudes than the source-selected samples. Indeed, 10 out of 30 objects in the SMA subsample are undetected in SDSS imaging, likely indicating that the SMA subsample probes higher redshifts or lower lens masses than any of the previous surveys. There is very little overlap in the observational properties of Herschel-selected lenses and SLACS or BELLS lenses, indicating that the two techniques are highly complementary.

With only a modest investment of observing time (≈1 hr on-source per target) with 4–6 m class optical telescopes, it is possible to measure spectroscopic redshifts for most of the lenses in the SMA subsample (see Sections 2.3 and 2.4 for details). We use the standard equations from Schneider et al. (1992) to compute the mass inside θしーたE for the lensing galaxies, ME. These values are shown for the SMA subsample as a function of lens redshift, zlens, in the right panel of Figure 4. This plot emphasizes the new range in parameter space that is probed by Herschel-selected lenses compared to SDSS-based lens searches: high redshift (zlens > 0.6) and low mass (ME < 1011M). The distinction in redshift confirms the evidence based on photometric redshifts presented by González-Nuevo et al. (2012). The objects from the literature with the most overlap with the SMA subsample are those from CASTLeS, CLASS, and SQLS—as expected, since these are source-selected samples of lenses. It should be noted that this comparison is not entirely fair to the Herschel-selected lenses because the SMA subsample is missing lens redshifts for optically-faint targets that are likely to have lower mass or lie at higher redshift (zlens ≳ 1). A handful of lenses identified by the SPT (Hezaveh et al. 2013) are also shown in this diagram, and are likely to have ensemble properties similar to those of the Herschel-selected lenses once statistically significant sample sizes are available.

Lenses discovered by Herschel have relatively high ellipticities compared to lenses selected from optical surveys. We measure a median ellipticity of epsilonlens = 0.35 ± 0.15. In comparison, Brewer et al. (2012) study a subset of the SLACS sample where the foreground deflector has an inclined disk and measure a median ellipticity of epsilonlens = 0.39 ± 0.07. There is a theoretical basis for why such an effect could occur: optical surveys for lenses might miss a large segment of highly inclined lenses due to dust obscuration by the foreground deflector (e.g., Bartelmann & Loeb 1998; Blain et al. 1999). A submm survey for lenses (like ours) is not affected by this limitation. However, our models do not include the effect of shear, which has a well-known degeneracy with lens ellipticity (e.g., Keeton et al. 1997). We therefore view this as an interesting line of research for further study and urge caution when readers consider this result.

As a final note on the properties of the lensing galaxies discovered by Herschel, it is worth emphasizing the sample size at present and how large it might grow in the future. The SMA subsample consists of a subset of 30 candidate lensed SMGs selected from 104 objects with S500 > 100 mJy within ≈400 deg2 of wide-field Herschel surveys. When the Herschel catalogs are complete, a total of ≈1000 deg2 of sky will be surveyed and should provide a sample of ≈250 lens candidates. This is comparable to the expected number of lensed SMGs found by the full SPT survey (Vieira et al. 2013), but is already a factor of ≈5 larger than other source-selected or heterogeneous surveys such as CLASS or CASTLeS. It is also comparable in size to SLACS and the initial release of strong lenses from BELLS.

5. INTRINSIC PROPERTIES OF LENSED SMGs DISCOVERED BY HERSCHEL

In this section, we focus on the intrinsic properties of the SMA subsample of lensed SMGs. We begin by discussing the size bias inherent to samples of strongly lensed galaxies. We then compare our magnification measurements with statistical predictions. Next, we describe our methodology for measuring dust temperatures, intrinsic (i.e., unlensed) FIR luminosities and FIR luminosity surface densities for the SMA subsample by combining modified blackbody fitting of the SPIRE and SMA photometry with the magnification factors and intrinsic source sizes predicted by our lens models. Throughout, we define the LFIR as integrated over 40–120 μみゅーm in the rest-frame (studies indicate a bolometric correction factor from LFIR to LIR of 1.91 is typical; e.g., Dale et al. 2001).

5.1. Size Bias in the SMA Subsample of Lensed SMGs

Strong gravitational lensing permits the study of SMGs at higher spatial resolutions than would otherwise be possible (e.g., a highly magnified SMG at z = 2 has been studied at 100 pc resolution with the SMA; Swinbank et al. 2010). While this is a highly attractive feature of strong lensing, it does necessitate certain unique considerations when transferring conclusions regarding lensed SMGs to the unlensed population. Chief amongst these considerations is the size bias inherent in any flux-limited sample of lensed galaxies. This bias has been investigated in a quantitative manner by a number of authors (Serjeant 2012; Hezaveh et al. 2012; Wardlow et al. 2013), who find that those objects with the brightest apparent flux densities should also have higher magnification factors and smaller sizes, on average.

Objects with high magnification factors are preferentially selected to have small sizes by flux-limited surveys like those of Herschel and SPT. This is because the degree of magnification depends primarily on the fraction of the source that is close to the caustic. A source that is extended relative to the size of the caustic will inevitably have a significant fraction of its flux density emitted in a region that is not highly magnified, so the total magnification factor summed over the entire source is not critically dependent on the exact location of the source relative to the caustic. Conversely, a population of lensed sources which are intrinsically compact will have a bimodal distribution of magnification factors that depends primarily on how far from the caustic the source is located. If a source is not highly magnified, it will likely not be bright enough to appear in our sample.

Figure 5 demonstrates this degeneracy between size (shown as half-light radius or rhalf, where $r_{\rm half} = a_{\rm s} \sqrt{(1-e_{\rm s})}$) and magnification factor (μみゅー880) for the SMA subsample and a handful of objects from SPT (Hezaveh et al. 2013). Our determinations of these values for the SMA subsample are presented in Table 6. Nearly every lensed SMG with rhalf < 0farcs1 is associated with μみゅー880 > 10. However, there are a surprising number of sources with 0farcs1 < rhalf < 0farcs2 and relatively modest magnification factors of 3 < μみゅー880 < 10. We will return to the implications of the large number of low-magnification objects in Section 5.2.

Figure 5.

Figure 5. Magnification factor at 880 μみゅーm as a function of half-light radius for lensed SMGs discovered by Herschel (red circles) and SPT (cyan stars). The most highly magnified sources are also the smallest, consistent with expectations from theoretical models (Serjeant 2012; Hezaveh et al. 2012). Vertical dashed lines represent maximum achievable spatial resolution at 880 μみゅーm for unlensed sources with (from right to left) SMA, ALMA Cycle 1, and full ALMA. In some cases, strong lensing permits the SMA to resolve sources that would otherwise require baseline lengths of >10 km (i.e., full ALMA).

Standard image High-resolution image

For a source at z = 2, these sizes correspond to a physical scale of ≈1 kpc, which is at the low end of sizes measured for unlensed SMGs (Tacconi et al. 2006). It must be noted, however, that there is a subtle bias when comparing the sizes of lensed and unlensed SMGs: if intrinsically low surface brightness regions are preferentially located farther from the caustic than high surface brightness regions (as expected for the reasons outlined above), then we expect the lensed SMGs to show smaller sizes than unlensed SMGs, because differential lensing makes detecting the low surface brightness regions more difficult than in the unlensed scenario.

Definitive measurements of the relative bias in the size measurements of lensed and unlensed SMGs will require spatially resolved observations of unlensed SMGs. The dashed vertical lines in Figure 5 illustrate both the advantages offered by lensing and the difficulties that must be overcome to assemble a statistically significant sample of unlensed SMGs with spatially resolved imaging. From right to left, the three lines indicate the maximum spatial resolutions available with the SMA, Cycle 1 ALMA, and full ALMA. It is only with the full ALMA and baselines >10 km that spatial resolution better than 0farcs1 can be achieved at these wavelengths—i.e., matching the best the SMA can do today for lensed SMGs discovered by Herschel.

5.2. Testing Predictions Derived from Lens Statistics

The number counts of unlensed SMGs fall off dramatically at the bright end of the luminosity function (e.g., Barger et al. 1999; Coppin et al. 2006; Oliver et al. 2010; Clements et al. 2010). This is the central reason why wide-field surveys at (sub-)mm wavelengths are useful tools for discovering strongly lensed galaxies (e.g., Blain 1996). There are several key elements of astrophysical interest in models which predict the magnification factor as a function of (sub-)mm flux density for strongly lensed galaxies found in wide-field (sub-)mm surveys. These are discussed in detail elsewhere (Perrotta et al. 2002; Negrello et al. 2007; Paciga et al. 2009; Hezaveh & Holder 2011; Wardlow et al. 2013), so we provide only the briefest of summaries here. In short, they are the lens mass profile (typically assumed to match the analytical form of Navarro–Frenk–White (NFW; Navarro et al. 1997) or a singular isothermal sphere) and the number densities of lenses and (unlensed) sources as functions of mass and redshift.

Figure 6 shows the magnification factor as a function of the 500 μみゅーm flux density for each strongly lensed SMG in the SMA subsample. Recall that we are complete for S500 > 170 mJy (see Section 2.1), except for one object which is the subject of a paper by H. Messias et al., (in preparation). The blue line is taken from Wardlow et al. (2013) and represents the predicted mean μみゅー as a function of S500 for a complete sample of strongly lensed SMGs (μみゅー > 2). There are far more low-μみゅー880 objects than expected based on the model. In fact, only two objects (J085358.9+015537 and J142825.5+345547) have μみゅー880 values that are consistent at the 1σしぐま level with the model predictions.

Figure 6.

Figure 6. Magnification factor from the SMA lens model as a function of 500 μみゅーm flux density. The prediction for these values from Wardlow et al. (2013) is shown by the dotted blue line, whereas the dashed purple line traces the same model, but with parameters tuned to jointly match the observed number counts and magnification factors shown in this diagram. The solid green lines show the effects of different maximum magnifications and are taken from Lapi et al. (2012).

Standard image High-resolution image

It is not presently clear why the model over-predicts the magnification factors at a given S500 value. One possibility is that our assumption of a single, smooth Sérsic profile for the background source leads to underestimates in some cases of the magnification factor. For example, one might imagine that a multi-component, clumpy model for the source morphology could reproduce the observed data while yielding larger magnification factors on average. Testing such models is beyond the scope of this paper, but we acknowledge that this possibility exists.

If the discrepancy between model and data is not simply a product of limited data quality, there still exist several possible explanations. Investigation into the dependence of the predicted mean magnification on the various model parameters shows that this prediction is sensitive to a number of factors. The first of these is the shape of the intrinsic SMG number counts. The intrinsic counts are not well constrained at the bright end, primarily because of the contribution from lensed SMGs (Wardlow et al. 2013). Modest changes in the parameters of the Schechter function used to characterize the counts have a significant effect on the predicted mean magnification factor. This can be seen by a comparison of the dotted blue line in Figure 6, which traces the predicted μみゅー880(S500) curve from Wardlow et al. (2013) based on the observed number counts of SMGs, and the dashed purple line, which shows μみゅー880(S500) found from a satisfactory joint fit to the SMG number counts and the magnification measurements reported in Table 6. The Schechter function from the joint fit has a flatter slope (by 25%), a brighter characteristic flux density (by 15%), and a lower normalization (by 30%). It is also possible that a Schechter function provides an inadequate description of the bright end shape of the number counts (e.g., if blending is not properly taken into account; see Fu et al. 2013; Ivison et al. 2013; Hodge et al. 2013; Karim et al. 2013).

Another important parameter for predicting magnification factors is a fixed maximum magnification (μみゅーmax), that is intended to reflect both the intrinsic sizes of the background sources as well as the typical angular separation between the centers of the sources and the centers of the lenses. The green lines shown in Figure 6 are taken from Lapi et al. (2012) and show the effect of this parameter on μみゅー880(S500). These curves help to highlight the apparent bimodality in μみゅー880 values in the SMA subsample. One speculative explanation for this is a bimodal distribution of sizes in SMGs (this is supported by the sizes from the lens modeling; see Table 6 and Figure 5) and hence a bimodal distribution in μみゅーmax. Larger samples and higher-spatial resolution are needed to confirm this intriguing possibility, which could plausibly result from SMGs comprising a population of rotating disks as well as major mergers. Such a study is beyond the scope of this paper and is therefore deferred to future publications.

5.3. SED Fitting Methodology

All but two sources in this sample have short baseline SMA data (i.e., D < 50 m) that provide a robust total flux density at 880 μみゅーm. The exceptions are J132630.1+334410, which is resolved into two distinct images of the background source, and J142823.9+352619, which appears unresolved in extended array only data. We therefore expect that the SMA total flux density measurements are reliable for the entire sample. In conjunction with the Herschel/SPIRE photometry, these data can be used to measure the apparent (i.e., lensed) bolometric luminosity as well as the shape of the far-IR SED. In this section, we describe the methodology we use to undertake this task.

For galaxies at redshifts of 1.5 < z < 4.5, Herschel/SPIRE and SMA 880 μみゅーm photometry probe rest-frame 45 μみゅーm to 350 μみゅーm. At these wavelengths, the dominant contribution is thermal emission from dust heated by star-formation or an active galactic nucleus (AGN). We fit single-temperature, optically-thin, modified blackbody curves to the data, assuming an emissivity parameter of βべーた = 1.5 (Hildebrand 1983). Studies based on Infrared Astronomical Satellite data have shown that this simple model gives a useful measure of the heating conditions of the ISM in galaxies (Desert et al. 1990). For the highest redshift sources (zsource > 3.5), the 250 μみゅーm channel of SPIRE probes the Wien side of the modified blackbody curve. The observed 250 μみゅーm flux density is under-predicted by this model if there are alternative powering sources that drive mid-IR luminosity (e.g., hot dust from AGN or intense SF). If such powering sources exist, the best-fit dust temperatures will be artificially inflated to compensate for the stronger 250 μみゅーm emission. However, for the four objects in the SMA subsample at z > 3.5), we do not find evidence for significantly higher dust temperatures or poor fits to the data (see Table 7, indicating that a simple modified blackbody is a reasonable choice even for the high-redshift objects.

Table 7. Intrinsic (i.e., Unlensed) Properties of SMA Sample (Assuming βべーた = 1.5, Optically Thin Modified Blackbody, and LFIR Integrated over 40–120 μみゅーm)

IAU Name $\chi ^2_{\rm min}$ Tdust log(Mdust) log(LFIR) rhalf log(ΣしぐまFIR)
(K) (M) (L) (kpc) (L kpc−2)
J021830.5−053124 7.11 36 ± 1 9.49 ± 0.11 12.79 ± 0.09 2.03 ± 0.71 11.45 ± 0.41
J022016.5−060143 4.46 37 ± 1 9.10 ± 0.10 12.79 ± 0.05 1.14 ± 0.37 11.93 ± 0.34
... ... ... 9.07 ± 0.10 12.76 ± 0.05 2.09 ± 0.35 11.34 ± 0.15
... ... ... 9.00 ± 0.10 12.67 ± 0.05 2.09 ± 0.39 11.24 ± 0.17
J083051.0+013224 0.86 44 ± 1 9.16 ± 0.06 13.09 ± 0.05 0.85 ± 0.07 12.44 ± 0.07
J084933.4+021443 9.62 36 ± 1 8.92 ± 0.05 12.72 ± 0.04 1.10 ± 0.22 11.86 ± 0.18
... ... ... 9.31 ± 0.05 13.11 ± 0.04 1.69 ± 0.19 11.87 ± 0.10
... ... ... 8.76 ± 0.05 12.56 ± 0.04 1.04 ± 0.42 11.85 ± 0.51
J085358.9+015537 0.98 36 ± 1 8.91 ± 0.09 12.37 ± 0.09 0.41 ± 0.08 12.37 ± 0.17
J090302.9−014127 1.13 38 ± 1 9.29 ± 0.07 12.92 ± 0.05 3.05 ± 0.92 11.20 ± 0.30
J090311.6+003906 0.87 34 ± 1 9.18 ± 0.06 12.45 ± 0.04 3.30 ± 0.65 10.63 ± 0.18
J090740.0−004200 10.30 43 ± 2 8.73 ± 0.10 12.58 ± 0.11 1.09 ± 0.62 11.92 ± 0.74
J091043.1−000321 25.66 41 ± 1 8.69 ± 0.06 12.50 ± 0.06 0.89 ± 0.16 11.82 ± 0.16
J091305.0−005343 0.04 35 ± 1 9.56 ± 0.08 12.94 ± 0.07 4.14 ± 0.72 10.92 ± 0.15
J105750.9+573026 2.64 47 ± 1 8.83 ± 0.05 12.94 ± 0.03 1.83 ± 0.26 11.62 ± 0.13
J114637.9−001132 1.86 41 ± 1 9.12 ± 0.06 12.90 ± 0.04 1.59 ± 0.13 11.70 ± 0.07
J125135.4+261457 2.24 39 ± 1 9.02 ± 0.06 12.68 ± 0.05 0.93 ± 0.21 11.97 ± 0.20
J125632.7+233625 0.41 40 ± 1 9.10 ± 0.07 12.80 ± 0.06 0.40 ± 0.07 12.80 ± 0.16
J132427.0+284452 47.59 37 ± 1 9.39 ± 0.07 12.92 ± 0.07 3.44 ± 0.44 11.05 ± 0.11
J132630.1+334410 13.33 36 ± 1 9.48 ± 0.04 13.00 ± 0.04 1.57 ± 0.17 11.81 ± 0.09
J133008.4+245900 1.14 43 ± 1 8.78 ± 0.06 12.66 ± 0.05 0.55 ± 0.24 12.52 ± 0.56
J133649.9+291801 7.62 39 ± 1 9.15 ± 0.09 12.87 ± 0.08 0.87 ± 0.34 12.27 ± 0.41
J134429.4+303036 6.90 38 ± 1 9.06 ± 0.04 12.73 ± 0.05 1.50 ± 0.40 11.62 ± 0.26
J141351.9−000026 23.28 38 ± 1 9.50 ± 0.09 13.18 ± 0.08 1.73 ± 0.31 11.92 ± 0.16
J142413.9+022303 0.49 41 ± 1 9.36 ± 0.05 13.17 ± 0.03 3.79 ± 0.38 11.22 ± 0.09
J142823.9+352619 6.14 39 ± 2 9.19 ± 0.20 12.77 ± 0.20 0.71 ± 0.43 12.52 ± 0.84
J142825.5+345547 0.88 38 ± 1 8.85 ± 0.10 12.45 ± 0.09 0.89 ± 0.18 11.77 ± 0.19
J143330.8+345439 0.84 39 ± 1 9.34 ± 0.06 12.96 ± 0.05 1.59 ± 0.34 11.78 ± 0.19

Note. Error bars are pertinent only to our chosen model and therefore underestimate the true uncertainties.

Download table as:  ASCIITypeset image

Another consideration that is important at high redshift is the influence of the cosmic microwave background (CMB) radiation. The CMB acts as an additional source of heating of the dust that shifts the SED to warmer temperatures and boosts the observed flux densities. However, for the dust temperatures and redshifts of the SMA subsample, this effect is insignificant (da Cunha et al. 2013) and we therefore do not incorporate it into our model fitting process.

The modified blackbody curve used here has the following form for the flux density, Sνにゅー, given a dust temperature, Tdust:

Equation (4)

An analysis of more complicated models that incorporate additional temperature components (e.g., Dunne & Eales 2001; Planck Collaboration et al. 2011) or allow the frequency at which the thermal emission becomes optically thick to vary (e.g., Hayward et al. 2012) is beyond the scope of this paper. This is because we are chiefly concerned with the apparent far-IR luminosities (μみゅーLFIR) of the sources, which are relatively insensitive to the particular details of the chosen modified blackbody model. The dust temperature is also of interest here, but mainly for the purpose of comparison to existing samples of SMGs. Our choices here are well-matched to those that have been made previously, thereby facilitating direct comparisons to the literature (e.g., Magnelli et al. 2012).

For a given μみゅーLFIR and Tdust, we compute model flux densities at the Herschel/SPIRE and SMA bands by multiplying the modified blackbody curve with the appropriate filter function and integrating (for the SMA, we assume a top-hat filter shape with 8 GHz of bandwidth centered on νにゅーLO). Calibration uncertainties are accounted for by adding 7% uncertainty in quadrature to each photometric measurement. Confusion noise is also included, though it is largely insignificant at the flux densities of the lensed SMGs. We use the emcee software package to iterate over plausible ranges in μみゅーLFIR (1010–1015L) and Tdust (20–100 K) values for each lensed SMG (see Section 3.1 for a description of the behavior of emcee). We use 100 walkers and 200 iterations in the "burn-in" stage to converge on the best-fit model, keeping in mind the additional 7% absolute flux density calibration uncertainty in the Herschel/SPIRE and SMA photometry. In the final stage, we use 100 walkers and 10 iterations for a total of 1000 sets of model parameters. These provide the shape of the posterior probability density functions for μみゅーLFIR and Tdust, which are then used to compute the best-fit values and the 1σしぐま confidence intervals. We use our measurements of μみゅー880 from Section 5 to obtain the intrinsic, unlensed FIR luminosity, LFIR. Finally, we compute dust masses using the standard equation (Hildebrand 1983),

Equation (5)

where $\kappa _{\rm d}^{\rm rest}$ is the mass absorption coefficient and B(νにゅーrest, Tdust) is the value of the blackbody function for the given Tdust and computed at the rest-frame frequency νにゅーrest. We obtain $\kappa _{\rm d}^{\rm rest}$ by interpolation of the values in Draine (2003) over the appropriate range in rest-frame wavelength. The uncertainty in $\kappa _{\rm d}^{\rm rest}$ is a factor of a few and dominates the total error budget for our dust mass estimates.

We follow the procedure outlined here for all SMGs with Herschel/SPIRE and submm photometry. Besides the objects in this paper, this also includes SMGs from Hezaveh et al. (2013) (with photometry and redshifts coming from Weiß et al. 2013; Bothwell et al. 2013) and Magnelli et al. (2012).

5.4. The SEDs of Herschel-selected Lensed SMGs

The results of our SED fitting procedure are given for the SMA subsample in Table 7, including the best-fit reduced chi-squared value ($\chi ^2_{\rm min}$), the dust temperature (Tdust), the dust mass (Mdust; error bars do not include the factor of a few uncertainty in the mass opacity coefficient), the FIR luminosity (LFIR), the half-light radius (rhalf), and the FIR luminosity surface density (ΣしぐまFIR). The error bars are pertinent only to our chosen model of a single optically thin modified blackbody. In this table, the magnification factor and uncertainty inferred from the lens model have been used to compute intrinsic values and their uncertainties. For 13 objects, $\chi ^2_{\rm min} > 2$, suggesting that the single-component modified blackbody model is an over-simplification in nearly half of the SMA subsample.

The TdustLFIR diagram is useful for characterizing the shape and amplitude of the rest-frame far-IR SED of dusty galaxies. Figure 7 shows these parameters for the objects in the SMA subsample, a handful of SPT lensed SMGs (Hezaveh et al. 2013), and a sample of primarily unlensed SMGs (Magnelli et al. 2012). Some of the objects in this diagram are known to have multiple components in the source plane (e.g., J022016.5−060143). In these cases, we show the LFIR value integrated over all components in the source plane. This is primarily done because our measurements of Tdust are based on Herschel photometry in large part, which does not resolve the individual components in the source plane.

Figure 7.

Figure 7. Dust temperature as a function of intrinsic (i.e., unlensed) FIR luminosity for lensed SMGs discovered by Herschel (red circles) and SPT (cyan stars), as well as unlensed and lensed SMGs from Magnelli et al. (2012; gray squares). Error bars include uncertainty in magnification factors.

Standard image High-resolution image

Objects in the SMA subsample populate the high-LFIR, high-Tdust regime (median LFIR = 7.9 × 1012L, median Tdust = 39 K) compared to unlensed SMGs (median LFIR = 2.8 × 1012L, median Tdust = 32 K). The selection of the brightest objects found in wide-field surveys is the dominant reason for the high LFIR values reported here. However, thanks to lensing, we probe a relatively wide range in LFIR: 2.7–17.0 × 1012L, despite selecting the brightest objects at 500 μみゅーm. The top axis in Figure 7 shows the star formation rates (SFRs) that are inferred based on the Kennicutt (1998) prescription for converting LIR (computed from LFIR assuming a bolometric correction factor of 1.91; Dale et al. 2001) to SFR assuming a Salpeter IMF. Finally, the high dust temperatures in the SMA subsample reflect the fact that a greater portion of the IR luminosity is emitted at short wavelengths in the highest luminosity sources. This result is consistent with what has been found previously for lensed SMGs (Harris et al. 2012) as well as for unlensed SMGs (Magnelli et al. 2012) and is one of the first indications that Herschel-selected lensed SMGs are not greatly dissimilar in their physical properties from unlensed SMGs (the same conclusion is reached by Harris et al. 2012).

5.5. Size Scale of Star-formation in Herschel-selected Lensed SMGs

One of the central themes in understanding the evolution of the most luminous galaxies during the epoch of peak SFR density in the universe (i.e., 1 < z < 4; e.g., Burgarella et al. 2013) is the role played by major mergers. It has been clear for decades that the most luminous galaxies at z ∼ 0—commonly known as ultra-luminous infrared galaxies (ULIRGs) and defined to have LIR > 1012L— are powered by major mergers (e.g., Armus et al. 1987; Clements et al. 1996; Murphy et al. 1996). However, it is also clear that such systems contribute only trivially to the SFR density in the universe today because they are so rare (Soifer et al. 1986). Since ULIRGs contribute an increasing fraction of the total SFR density as a function of redshift (e.g., Le Floc'h et al. 2005; Magnelli et al. 2011), some theoretical efforts have suggested that an increase in the merger rate in conjunction with strong inflows of gas from the intergalactic medium (IGM) at high redshift could allow the merger paradigm to explain a large fraction of the luminous galaxies at these epochs (e.g., Hopkins et al. 2010). Providing support for this theoretical paradigm are sub-arcsecond observations of CO emission in a handful of SMGs which show that a significant fraction of SMGs that are spatially resolved have multiple, similar mass components—i.e., they are major mergers (Tacconi et al. 2008; Engel et al. 2010; Ivison et al. 2011; Riechers et al. 2011a).

On the other hand, some recent theoretical attempts to simulate the formation of galaxies on cosmological scales (i.e., cubes that are ≈200 Mpc on a side) have found evidence that favors a model of galaxy formation in which smooth flows of gas from the IGM feed large, extended reservoirs of gas in disk galaxies (e.g., Kereš et al. 2009; Davé et al. 2010). There is even observational evidence based on dynamical models that disk-like systems exist at high redshift (Hodge et al. 2012). Ultimately, detailed dynamical models of statistically significant samples can resolve the dispute between the merger and disk paradigms. However, assembling the requisite datasets is extremely expensive in terms of telescope time. A far more feasible goal is spatially resolved observations of the dust emission in SMGs at high redshift.

The dust emission in SMGs is critically important because it represents reprocessed emission from young massive stars which provide a reliable measure of the instantaneous SFR (within the past ≈10 Myr). This assumes no significant contribution from a cold, diffuse ISM component (supported by our measurements of Tdust) and no significant contribution from an AGN (our use of the FIR luminosity integrated over 40–120 μみゅーm rest-frame is intended to mitigate this possibility). Measuring the size-scale and the luminosity of this dust can therefore in principle be used to contrast extended galaxy morphologies expected from accretion-fuelled disks (e.g., Dekel et al. 2009) and compact morphologies expected from dissipational mergers of gas-rich disks (e.g., Mihos & Hernquist 1996).

The left panel of Figure 8 shows the physical sizes (circularized radii, rhalf) as a function of LFIR for the SMA subsample and the SPT sample (Hezaveh et al. 2013). In this figure, objects with multiple components in the source plane are plotted individually, unlike in Figure 7. In doing this, we have assumed that the FIR luminosity in each component is proportional to its intrinsic flux density at 880 μみゅーm from the lens model (i.e., μみゅーFIRμみゅー880. This assumption is consistent with our modeling efforts because the only time we use multiple components in the source plane is when it is clear that only moderate lensing is occurring, in which case μみゅーFIRμみゅー880 is a valid approximation. Two comparison samples of unlensed LIRGs and ULIRGs from Rujopakarn et al. (2011) are shown: one at 0.5 < z < 2.5 and one at z ∼ 0. The sample of intermediate and high redshift LIRGs and ULIRGs relies on size measurements based primarily on radio observations (Chapman et al. 2004; Muxlow et al. 2005; Biggs & Ivison 2008; Casey et al. 2009), but also includes a handful of with mm size measurements (Tacconi et al. 2006, 2010; Daddi et al. 2010) and the assumption that the radio and far-IR sizes are correlated. Finally, Figure 8 also shows the median and 1σしぐま range in these values for a handful of unlensed SMGs with high-spatial resolution imaging by Tacconi et al. (blue rectangle; 2006).

Figure 8.

Figure 8. Left: half-light radius as a function of FIR luminosity for lensed SMGs discovered by Herschel (red circles) and SPT (cyan stars), as well as galaxies from a compilation in Rujopakarn et al. (2011) at z > 0.5 (filled gray diamonds) and at z ∼ 0 (open squares). The blue shaded region represents the median and 1σしぐま range found for unlensed SMGs by Tacconi et al. (2006). Right: far-IR luminosity surface density as a function of FIR luminosity. The orange dashed line traces the theoretical limit of ΣしぐまSFR for an optically thick disk (Thompson et al. 2005). The SMA subsample spans nearly one decade in LFIR and two decades in ΣしぐまFIR. A handful of sources approach or exceed the highest values observed in local LIRGs and ULIRGs (ΣしぐまFIR = 1013L kpc−2).

Standard image High-resolution image

We find a wide range in sizes for the SMA subsample: the minimum, median, and maximum rhalf values are 0.41 kpc, 1.53 kpc, and 4.16 kpc. In comparison, Rujopakarn et al. (2011) find values of 0.9 kpc, 2.6 kpc, and 8.0 kpc for the intermediate and high redshift sample. Part of this difference is a result of lensing, which lets us access spatial resolutions that are otherwise inaccessible (the radio observations used to measure sizes in the comparison sample have a typical angular resolution of 0farcs15, corresponding to a physical scale of ≈1 kpc at the redshifts of interest). Local LIRGs and ULIRGs have smaller sizes than their higher redshift counterparts (Rujopakarn et al. 2011), but the sizes of objects in the SMA subsample begin to overlap with those of the local LIRGs and ULIRGs. Very strong lensing (e.g., μみゅー ≳ 30) can reach ≈100 pc scale spatial resolution (Swinbank et al. 2010), but these occurrences must be rare because no such object is found in the SMA subsample.

The right panel of Figure 8 shows the FIR luminosity surface density (ΣしぐまFIR) as a function of LFIR for the same set of objects as in the left panel. A wide range in ΣしぐまFIR is evident for the SMA subsample: the minimum, median, and maximum ΣしぐまFIR values are 0.05, 0.6, and 6.0 × 1012L kpc−2, respectively. A horizontal line drawn at 1000 M yr−1 kpc−2 represents the theoretical limit for SFR surface density in a sustainable radiation-pressure supported disk (Thompson et al. 2005). A handful of objects in the SMA subsample and one object in the SPT sample lie near to or in excess of this limit, possibly indicating that they are in a very short-lived phase of evolution (e.g., the coalescence stage of a major merger). There are also, however, a number of objects with ΣしぐまFIR values more than an order of magnitude below the limit, suggesting that multiple modes of star-formation are viable in high redshift SMGs.

A spherically symmetric dust source radiating as a blackbody obeys the Stefan–Boltzmann law relating emitted flux density and the temperature of the dust. We use this fact to infer an alternative measure of the size of the lensed SMG, which we denote as rSB:

Equation (6)

where σしぐまSB is the Stefan–Boltzmann constant. This quantity is similar in scope to the "effective radius" described in Greve et al. (2012) for lensed SMGs discovered by the SPT. Figure 9 shows the results of this analysis for the SMA and SPT samples. In this diagram, the error bars reflect the formal values obtained for our given set of model assumptions—e.g., the uncertainty introduced by the assumption of a single-temperature modified blackbody is not included. A dashed line traces the 1:1 relation between these two measures of the size (i.e., rSB = rhalf).

Figure 9.

Figure 9. Comparison of size measurements (rSB) computed from application of Stefan–Boltzmann law to intrinsic IR luminosities and dust temperatures with those obtained from direct measurements from lens models (circularized half-light radius, rhalf). A dashed blue line traces rSB = rhalf. Approximately 75% of the sample has rhalfrSB. For these sources, the FIR emission is likely to be optically thick.

Standard image High-resolution image

A total of 15 out of 23 sources satisfy the unphysical criterion of rSB > rhalf. The best explanation for the large number of sources that violate the blackbody limit is that our Tdust values are underestimated by the assumption of optically thin emission. In fact, rSBrhalf suggests optically thick FIR emission. Adopting an optically thick model for the SED fitting would lead to larger dust temperatures by 25%–50%, an increase that is nearly sufficient for all of the sources in our sample to satisfy rSB < rhalf. The exact geometry of the source is unknown and is therefore an additional complicating factor. Nevertheless, this crude line of analysis is one indication that the FIR emission is optically thick in lensed SMGs discovered by Herschel, similar to local ULIRGs, which have rhalf ≈ 2–3 × rSB (Solomon et al. 1997). Only a handful of sources have large rhalf values and low ΣしぐまFIR values typical of optically thin disks far from the Eddington limit (e.g., J091305.0−005343).

6. CONCLUSIONS

We present sub-arcsecond submm imaging from the SMA for 30 strong lens candidates discovered by Herschel in the H-ATLAS and HerMES wide-field surveys. The candidates are selected to have S500 > 100 mJy and the SMA subsample contains nearly all targets with S500 > 170 mJy. We also present optical spectroscopy from the MMT, Gemini-S, and WHT that provide new redshift measurements for 8 of the putative lenses. Nearly all candidates in the SMA subsample have existing spectroscopic redshifts for the putative lensed SMGs from blind CO searches with the GBT (Harris et al. 2012), CSO (Lupu et al. 2012), CARMA (D. A. Riechers et al., in preparation), and PdBI (Cox et al. 2011; M. Krips et al., in preparation).

Out of the SMA subsample of 30, there are 16 that have distinct lens and source redshifts and obvious lensed morphology in the submm ("grade A" lenses). Four objects have convincing morphological signatures of lensing, but only one spectroscopic redshift measurement—we consider these to be highly likely to be strongly lensed ("grade B" lenses). Another five objects have distinct redshift measurements for lens and source, but the SMA imaging reveals only one image of the background source, suggesting modest magnification factors: μみゅー880 < 2 ("grade C" lenses). Finally, there are five objects that lack distinct redshift measurements and do not show obvious morphological signatures of lensing—additional data are needed to determine whether or not these "grade X" systems are strongly lensed. In total, the strong lensing rate is 70%–87% (83%–100% if moderately lensed systems are included as well).

We use the SMA data to develop lens models in the visibility plane, as is appropriate for interferometers like the SMA. We derive lens models for the 25 objects with obvious signatures of lensing (either strong or moderate) in the submm. In conjunction with redshifts from optical and mm-wave spectroscopy, the lens models provide measurements of the mass of the lenses inside the Einstein radius, as well as the size and far-IR luminosity of the lensed SMGs.

We find that the lenses are at higher redshifts and have lower masses than lenses found in surveys based on SDSS optical spectroscopy, in agreement with expectations for a source-selected (rather than lens-selected) survey for lenses. The number of lenses that will be found from wide-field (sub-)mm surveys (González-Nuevo et al. 2012) promises to be comparable to that from SDSS-based searches, but the former provide access to a population of lenses with fundamentally different properties. For this reason, lenses found by Herschel and SPT are highly complementary to those found by SDSS and will remain so for the foreseeable future.

The lensed SMGs probe over a decade in sizes (median circularized half-light radii of 1.6 kpc) and intrinsic (i.e., unlensed) FIR luminosity (median LFIR of 7.9 × 1012L). Applying the Kennicutt (1998) prescription to convert LIR to SFR, we use the sizes and LFIR values to infer a nearly two-decade range in SFR surface density (median ΣしぐまSFR = 200 M yr−1 kpc−2). A handful of lensed SMGs lie near or above the theoretical limit of ΣしぐまSFR = 1000 M yr−1 kpc−2 for an optically thick disk, but there are also several objects with ΣしぐまSFR values over an order of magnitude below this limit, implying that multiple modes of star-formation may be required to explain SMGs at z ≳ 1.5.

The magnification factors we measure for the lensed SMGs are significantly lower than predicted from models based on number counts of unlensed SMGs. This may be an indication that the bright end of the SMG luminosity function or the intrinsic sizes of SMGs are currently poorly understood.

Finally, it is worth emphasizing that the advent of ALMA makes the future in this field looks very promising. Many of the unsolved questions from this work can be addressed in a direct manner by obtaining more sensitive submm observations at higher spatial resolution. For a given amount of integration time, ALMA (when fully operational) will provide factors of 10–100 improvement in these quantities compared to the SMA.

The results described in this paper are based on observations obtained with Herschel, an ESA space observatory with science instruments provided by European-led Principal Investigator consortia and with important participation from NASA. The Herschel-ATLAS is a project with Herschel. The H-ATLAS Web site is http://www.h-atlas.org/. US participants in H-ATLAS acknowledge support from NASA through a contract from JPL.

This research has made use of data from the HerMES project (http://hermes.sussex.ac.uk/). HerMES is a Herschel Key Programme utilizing Guaranteed Time from the SPIRE instrument team, ESAC scientists, and a mission scientist. HerMES is described in Oliver et al. (2012). The HerMES data presented in this paper will be released through the Herschel Database in Marseille HeDaM (36).

SPIRE has been developed by a consortium of institutes led by Cardiff Univ. (UK) and including: Univ. Lethbridge (Canada); NAOC (China); CEA, LAM (France); IFSI, Univ. Padua (Italy); IAC (Spain); Stockholm Observatory (Sweden); Imperial College London, RAL, UCL-MSSL, UKATC, Univ. Sussex (UK); and Caltech, JPL, NHSC, Univ. Colorado (USA). This development has been supported by national funding agencies: CSA (Canada); NAOC (China); CEA, CNES, CNRS (France); ASI (Italy); MCINN (Spain); SNSB (Sweden); STFC, UKSA (UK); and NASA (USA).

R.S.B. acknowledges support from the SMA Fellowship program. H.F, A.C., and J.L.W. acknowledge support from NSF CAREER AST-0645427. A portion of this work was completed at the Aspen Center for Physics during a 2013 summer workshop on dusty galaxies at high redshift. R.S.B. acknowledges the hospitality of the Aspen Center for Physics, which is supported by the National Science Foundation Grant No. PHY-1066293. S.J.O., L.W., and A.S. acknowledge support from the Science and Technology Facilities Council (grant No. ST/I000976/1). M.N. acknowledges financial support from PRIN INAF 2012 project "Looking into the dust-obscured phase of galaxy formation through cosmic zoom lenses in the Herschel Astrophysical Large Area Survey." A.O. and R.G. acknowledge support from the Programme National Cosmologie et Galaxies (PNCG). J.G.N. acknowledges financial support from Spanish CSIC for a JAE-DOC fellowship and partial financial support from the Spanish Ministerio de Ciencia e Innovacion project AYA2010-21766-C03-01. I.P.-F., P.M.-N., N.L. and A.S. acknowledge support from the Spanish grant AYA2010-21697-C05-04. We thank K. Rosenfeld for assistance in implementing the visibility-plane aspect of the lens modeling software used in this paper. We thank the anonymous referee for a timely review that provided useful comments and helped improved the clarity of the manuscript.

The SMA is a joint project between the Smithsonian Astrophysical Observatory and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smithsonian Institution and the Academia Sinica. The authors wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Mauna Kea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain.

Observations reported here were obtained at the MMT Observatory, a joint facility of the Smithsonian Institution and the University of Arizona.

Based on observations obtained at the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agreement with the NSF on behalf of the Gemini partnership: the National Science Foundation (United States), the National Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), Ministério da Ciência, Tecnologia e Inovação (Brazil) and Ministerio de Ciencia, Tecnología e Innovación Productiva (Argentina).

The William Herschel Telescope is operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofísica de Canarias.

Facilities: SMA - SubMillimeter Array, MMT - MMT at Fred Lawrence Whipple Observatory, Gemini:South - Gemini South Telescope, ING:Herschel - Isaac Newton Group's William Herschel Telescope, VLT:Melipal - Very Large Telescope (Melipal)

Footnotes

Please wait… references are loading.
10.1088/0004-637X/779/1/25