Brought to you by:

Articles

CO-EVOLUTION OF GALAXIES AND CENTRAL BLACK HOLES: OBSERVATIONAL EVIDENCE ON THE TRIGGER OF AGN FEEDBACK

Published 2012 April 13 © 2012. The American Astronomical Society. All rights reserved.
, , Citation Y. Matsuoka 2012 ApJ 750 54 DOI 10.1088/0004-637X/750/1/54

0004-637X/750/1/54

ABSTRACT

A comprehensive analysis of the extended emission-line region (EELR) around quasars is presented. A new Subaru/Suprime-Cam observation is combined with a literature search, resulting in a compilation of 81 EELR measurements for type-1 and type-2 quasars with an associated active galactic nucleus (AGN) and host galaxy properties. It is found that the EELR phenomenon shows clear correlation with the Eddington ratio, which links EELR to the constituents of principal component 1, or eigenvector 1, of the AGN emission correlations. We also find that EELR is preferentially associated with gas-rich, massive blue galaxies. This supports the idea that the primary determinant of EELR creation is gas availability and that the gas may be brought in by galaxy merger, triggering the current star formation as well as AGN activity, and also gives an explanation for the fact that most luminous EELRs are found around radio-loud sources with low Eddington ratio. By combining all the observations, it is suggested that EELR quasars occupy the massive blue corner of the green valley, the AGN realm, on the galaxy color–stellar mass diagram. Once a galaxy is pushed to this corner, an activated AGN would create an EELR by energy injection into the interstellar gas and eventually blow it away, leading to star formation quenching. The results presented here provide a piece of evidence for the presence of such an AGN feedback process, which may play a leading role in the co-evolution of galaxies and central super-massive black holes.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Some active galactic nuclei (AGNs) are surrounded by massive ionized nebulae extending to several tens of kiloparsec from the central engines. The first systematic studies of such extended emission-line regions (EELRs) around quasars in combination with nuclear spectral properties were presented by Boroson & Oke (1984) and Boroson et al. (1985). They found that nearly half of their 24 objects were EELR quasars (quasars with a detectable EELR, following Fu & Stockton 2007) and that the EELR quasars have larger [O iii] λ5007 equivalent widths (EWs), broader and bumpier Balmer lines, and weaker Fe ii emission compared to non-EELR quasars. In addition, the EELR quasars are predominantly associated with steep-spectrum radio emission with extended, lobe-dominant morphology, in contrast to the non-EELR quasars associated with flat-spectrum radio emission with compact, core-dominant morphology or without strong radio emission.

The correlation between EELR phenomenon and radio emission was further confirmed by Stockton & MacKenty (1987), who obtained narrowband [O iii] line images of 47 quasars. Above the limiting luminosity of extended [O iii] emission $L_{\rm E[{\rm O\,\mathsc{iii}}]} = 5 \times 10^{41}$ erg s−1, 10 out of 26 (10/26) of their steep-spectrum radio quasars (SSRQs) have EELRs, while the majority of the flat-spectrum radio quasars (FSRQs) and the radio-quiet quasars (RQQs; 7/7 of the FSRQs and 13/14 of the RQQs) are non-EELR sources (Fu & Stockton 2009). They also found a correlation between nuclear [O iii] luminosity and the presence of EELRs. Fu & Stockton (2007) analyzed the rest-frame ultraviolet spectra of 12 SSRQs in the Stockton & MacKenty (1987) sample and found that N v λ1240 emission is suppressed in EELR quasars, which they claimed points to the lower metallicity of the broad-line region (BLR) compared to non-EELR quasars. They did not find any systematic difference in other AGN properties such as continuum luminosity and mass of super-massive black holes (SMBHs) between the EELR and non-EELR SSRQs.

On the other hand, Husemann et al. (2008) observed 20 quasars dominated by RQQs and found EELRs around 8 of them. The correlation among nuclear Fe ii EW, Hβ line width, and EELR was also confirmed in their sample too. In addition, they found a dependence of EELRs on SMBH mass (MBH) and reported that MBH = 108.5M separates EELR quasars with larger MBH from non-EELR quasars with smaller MBH. Searches for EELRs around RQQs were also presented by Humphrey et al. (2010) and Villar-Martín et al. (2010, 2011), focusing on obscured type-2 quasars. They found EELRs around 10 out of 20 objects and claimed that a luminous EELR is not necessarily associated with powerful radio sources.

Observational constraints on the emergence of an EELR can possibly be a key to understanding the co-evolution of galaxies and SMBHs. It is now widely accepted that the tight correlation between SMBH mass and the bulge mass of their host galaxies (e.g., Magorrian et al. 1998; Häring & Rix 2004) indicates the intimately connected evolution of the two systems. The most compelling process of interaction between galaxies and SMBHs is AGN feedback, in which the energy emitted by AGNs regulates the mass accretion to SMBHs and drives galactic wind that can expel cold gas from and quench star formation in host galaxies. The AGN feedback could also give solutions to the two major problems in the current galaxy formation models based on the Λ cold dark matter (CDM) theory: the first is the "overcooling" problem in which much more massive galaxies are formed in the numerical simulations than are observed due to too-efficient gas cooling, and the second is the "inverted color–mass–morphology relation" problem in which massive galaxies are predicted to be blue and disk dominant rather than red and spheroid dominant as observed (e.g., Somerville et al. 2008). However, observational evidence is still lacking as to whether or not the AGN feedback is actually present and how it works if it is. While measurements of galaxy evolution imprinted on the stellar mass (luminosity) function (e.g., Bell et al. 2004; Faber et al. 2007; Brown et al. 2007, 2008) especially at its massive end (Conselice et al. 2007; Matsuoka & Kawara 2010; Matsuoka et al. 2011a) and its integration (the extragalactic background light; e.g., Matsuoka et al. 2011b) or analyses of the nature of the AGN phenomenon (e.g., Laor 2000; Boroson 2002; Matsuoka et al. 2007, 2008; Vestergaard et al. 2008) can put indirect constraints on the co-evolution, EELRs possibly provide more direct clues about the interrelation between galaxies and SMBHs. The previous measurements indicate that EELR gas is largely ionized by AGNs rather than by stellar population (Fu & Stockton 2009; Husemann et al. 2010; Keel et al. 2012); hence, we may be witnessing AGN feedback in action in this galaxy-wide ionization process. Indeed, Greene et al. (2011) recently showed that AGNs photoionize and kinematically disturb the interstellar medium throughout the entire host galaxies of luminous type-2 quasars.

In this paper, a new observation of an EELR around five quasars at z ∼ 0.3 is first presented. It is then combined with previous observations of EELRs and properties of associated AGNs and host galaxies obtained mainly from the Sloan Digital Sky Survey (SDSS; York et al. 2000) archive. By using this compiled sample as well as the control samples of SDSS quasars and galaxies, we aim to perform a comprehensive analysis of EELRs in the context of galaxy evolution. A cosmology with H0 = 70 km s−1 Mpc−1, ΩM = 0.3, and ΩΛ = 0.7 is assumed throughout this work. Luminosity and mass are given in units of erg s−1 and the solar mass (M) unless otherwise noted. All the presented magnitudes are on the AB system.

2. DATA

2.1. New Observation

The observed quasars were selected from the SDSS Data Release 7 (DR7) quasar catalog (Schneider et al. 2010; Shen et al. 2011). The simple selection criteria were set based on redshift (z ≃ 0.3 in order for [O iii] λ5007 to fall in the narrowband filter transmission described below) and nuclear [O iii] luminosity ($L_{\rm N[{\rm O\,\mathsc{iii}}]} > 10^{42}$ erg s−1). Among the objects meeting the criteria, those with various SMBH mass and radio loudness and with the highest [O iii] luminosity were picked up. The properties of the observed quasars are summarized in Table 1. They effectively complement the existing measurements on the AGN principal-component (PC) plane explored below.

Table 1. Targets of Subaru/Suprime-Cam Observation

Quasar r-band Redshift $\log L_{{\rm N[{\rm O\,\mathsc{iii}}]}}$ log MBH Rb $\log L_{\rm E[{\rm O\,\mathsc{iii}}]}$
  Mag.a   (erg s−1) (M)   (erg s−1)
SDSS 091401.75+050750.6 17.32 0.301 42.85 ± 0.01 9.36 ± 0.10 L 42.17 ± 0.08
SDSS 095456.89+092955.7 17.92 0.298 42.93 ± 0.01 8.56 ± 0.08 L <41.47
SDSS 113949.47+402048.5 17.52 0.314 42.97 ± 0.01 8.21 ± 0.08 Q <41.48
SDSS 150740.92+445331.5 18.16 0.314 42.69 ± 0.01 7.34 ± 0.14 L <41.53
SDSS 150752.66+133844.5 17.79 0.322 42.53 ± 0.01 9.67 ± 0.02 Q 41.83 ± 0.08

Notes. aSDSS PSF magnitude. The measurement error is <0.02 mag. bRadio loudness. Q: radio-quiet, L: radio-loud.

Download table as:  ASCIITypeset image

The targets were observed with the prime-focus imaging facility Suprime-Cam (Miyazaki et al. 2002) mounted on the Subaru telescope.1 The instrument covers a wide field of view of 34 × 27 arcmin2 with 10 2k × 4k CCDs, with a pixel scale of 0.20 arcsec. The narrowband NA656 (effective wavelength λeff = 6570 Å, effective bandpass Δλeff = 140 Å) and the broadband Rceff = 6530 Å, Δλeff = 1110 Å) filters were used to trace the redshifted [O iii] line and underlying continuum, respectively. The observation was carried out on 2011 May 3–4 under the open-use program S11A-028. The sky condition was sometimes non-photometric, but the measured flux of celestial objects was mostly stable within 10% variation. The seeing was 0.8–1.1 arcsec. For each target, nine 300 s or three 900 s exposures in NA656 and nine 100 s or three 300 s exposures in Rc were obtained, depending on the source brightness. In addition, a spectrophotometric standard star SA 104-335 was observed with the same instrument configuration as the target observation.

Data reduction was performed with the Suprime-Cam Deep Field Reduction (SDFRED; Ouchi et al. 2004) software in a standard manner including bias subtraction, flat fielding, distortion correction, sky subtraction, masking bad pixels, and stacking. Photometric calibration in NA656 was achieved by referring to the observed flux of SA 104-335, whose intrinsic magnitude was derived by convolving its spectrophotometry data (Stone 1996) with the filter transmission function. Then, the Rc images were calibrated relative to NA656 by requiring the mean (RcNA656) color to be zero for galaxies detected with high signal-to-noise ratios (point sources were not used because stellar Hα absorption could affect this calibration). The scatter of ∼0.05 mag was found around the mean between Rc and NA656 magnitudes in the calibration.

The reduction process then proceeded to the continuum subtraction from the NA656 images. For this purpose, the nuclear spectra of the quasars were retrieved from the SDSS DR7 archive2 (an example is shown in Figure 1). They were measured within SDSS spectroscopic fibers of 3 arcsec aperture. For each spectrum, the continuum levels beneath [O iii] λ4959 and λ5007 were estimated by interpolation between continuum windows at both (shorter and longer wavelength) sides of the lines, and then the spectrum was decomposed into [O iii] and continuum. Convolving them with the filter transmissions gave the relative fluxes of [O iii] and continuum in NA656 ($f_{{\it NA}_{656}}^{\rm [{\rm O\,\mathsc{iii}}]}$ and $f_{{\it NA}_{656}}^{\rm cont}$) and in Rc ($f_{R_{\rm c}}^{\rm [{\rm O\,\mathsc{iii}}]}$ and $f_{R_{\rm c}}^{\rm cont}$) for each object. Using these measures, accurate continuum subtraction was achieved as follows: The radial profiles of all the quasar nuclei on the obtained images are consistent with the point-spread functions (PSFs) derived from nearby stars; hence, they are dominated by the radiation from central unresolved sources. The NA656 and Rc images were decomposed into nuclear (NAnuc656, Rnucc) and extended (NAext656, Rextc) components by the best-fit Moffat profiles, fitted within a 10 pixel (2 arcsec) aperture on top of the quasar peak positions. Then, the nuclear [O iii] distributions are given by

Equation (1)

while the extended [O iii] distributions are

Equation (2)

The second term in expression (2), which is equivalent in principle to the ratio $\Delta \lambda _{{\rm eff}, R_{\rm c}}$/($\Delta \lambda _{{\rm eff}, R_{\rm c}} - \Delta \lambda _{\rm eff, {\it NA}_{656}}$), corrects for the continuum oversubtraction of the first term due to the [O iii] themselves contained in Rextc. Note that expression (1), although it is most accurate, is not applicable to the extended components since the relative fluxes of [O iii] and the continuum are different from $f_{\it NA_{656}}^{\rm [{\rm O\,\mathsc{iii}}]}$ and $f_{\it NA_{656}}^{\rm cont}$ measured in the SDSS spectroscopic fibers. It is worth emphasizing that the spectral information is quite useful in the above process as an accurate cross-calibrator of the images taken in the different filters.

Figure 1.

Figure 1. Nuclear spectrum of SDSS 1139+4020 (solid line) with the arbitrary-scaled transmission functions of broadband Rc and narrowband NA656 filters (dotted lines).

Standard image High-resolution image

2.2. Data Compilation

Thanks to the massive spectroscopic observations conducted by the SDSS, nuclear properties are now known for many quasars with EELR measurements in the literature. We compile the EELR measurements of type-1 quasars from Boroson et al. (1985), Stockton & MacKenty (1987), and Husemann et al. (2008, 2010). The continuum and line luminosities given in Stockton & MacKenty (1987) are converted to the values in the cosmology adopted here. The measurements of Husemann et al. (2008) are corrected in accordance with their re-analysis (B. Husemann et al. 2012, in preparation). Then, we cross-match the quasars with the SDSS DR7 quasar catalog (Shen et al. 2011) and obtain nuclear properties such as continuum and line luminosity, SMBH mass, and radio characteristics based on the Faint Images of the Radio Sky at Twenty cm survey (FIRST; Becker et al. 1995). For some objects without counterparts in the SDSS catalog, the nuclear properties are taken from the above papers as well as from Fu & Stockton (2007). Combined with the new observation presented in this paper, we end up with 61 type-1 quasars with EELR measurements (27 detections and 34 non-detections) and with a variety of nuclear properties. The compiled sample is listed in Table 2. We assign radio flag 0 to RQQs, flag 1 to flat-spectrum (radio spectral index αν < 0.5) or core-dominant (based on the FIRST morphology; Jiang et al. 2007) radio quasars, and flag 2 to steep-spectrum (αν > 0.5) or lobe-dominant radio quasars. The SDSS quasars are separated into radio-quiet and radio-loud sources at the radio loudness parameter R = 5 (Shen et al. 2011). For simplicity, the objects with radio flags 0, 1, and 2 are hereafter referred to as RQQs, FSRQs, and SSRQs, respectively, since there is a well-known, strong correlation between the radio morphology and spectral index.

Table 2. Nuclear Properties of Type-1 Quasars with EELR Measurements

Quasar Redshift log λLλ5100 log MBH Ra $\log L_{\rm N[{\rm O\,\mathsc{iii}}]}$ $\log L_{\rm E[{\rm O\,\mathsc{iii}}]}$ Ref.
    (erg s−1) (M)   (erg s−1) (erg s−1)  
4C 15.01 0.450 45.61 9.24 2 43.58 <42.56 2
PG 0026+129 0.142 44.96 ± 0.11 8.98 0 42.53 ± 0.09 41.80 ± 0.14 3
PG 0050+124 0.061 44.31 ± 0.11 8.00 0 42.26 ± 0.09 <43.0 3
PG 0052+251 0.155 44.98 ± 0.10 8.55 0 42.70 ± 0.09 41.75 ± 0.13 1, 2, 3
SDSS 0057+1446 0.172 44.93 ± 0.01 9.38 ± 0.02 0 42.23 ± 0.03 <43.0 3
HE 0132−0441 0.154 44.69 ± 0.10 7.73 0 42.15 ± 0.16 <43.0 3
NAB 0137−01 0.334 45.17 ± 0.01 9.01 ± 0.02 0 42.62 ± 0.02 <41.53 2
SDSS 0155−0857 0.165 44.43 ± 0.02 8.82 ± 0.04 0 42.15 ± 0.02 41.43 ± 0.19 3
HE 0157−0406 0.218 44.72 ± 0.09 8.52 0 42.09 ± 0.10 <43.0 3
Mrk 1014 0.163 44.78 ± 0.01 8.06 ± 0.05 1 42.74 ± 0.02 42.22 ± 0.14 2, 3
OI 287 0.444 45.08 ± 0.01 9.09 ± 0.21 2 43.51 ± 0.01 <42.47 2
SDSS 0836+4426 0.254 45.29 ± 0.01 8.73 ± 0.06 0 43.28 ± 0.01 43.08 ± 0.09 3
SDSS 0914+0507 0.301 44.84 ± 0.01 9.36 ± 0.10 2 42.85 ± 0.01 42.30 ± 0.08 5
SDSS 0948+4335 0.226 44.76 ± 0.01 8.43 ± 0.03 0 42.44 ± 0.01 <43.0 3
SDSS 0954+0929 0.298 44.54 ± 0.01 8.56 ± 0.08 2 42.93 ± 0.01 <41.60 5
Ton 28 0.329 45.44 ± 0.01 8.03 ± 0.43 0 42.42 ± 0.02 <42.34 2
4C 13.41 0.241 45.41 ± 0.01 9.54 ± 0.05 2 42.40 ± 0.02 <41.41 2
HE 1029−1401 0.086 44.97 ± 0.05 8.70 0  ⋅⋅⋅ 41.78 ± 0.12 4
PG 1049−005 0.359 45.59 ± 0.01 9.17 ± 0.03 0 43.47 ± 0.01 <42.31 2
4C 61.20 0.421 45.26 ± 0.01 9.10 ± 0.09 1 43.34 ± 0.01 <42.55 2
4C 10.30 0.422 44.78 ± 0.01 9.33 ± 0.05 2 42.75 ± 0.01 <41.62 2
3C 249.1 0.313 45.42 8.96 2 43.46 42.85 ± 0.02 1, 2
SDSS 1131+2632 0.244 44.56 ± 0.02 9.08 ± 0.04 0 42.81 ± 0.01 42.11 ± 0.14 3
SDSS 1139+4020 0.314 44.63 ± 0.01 8.21 ± 0.08 0 42.97 ± 0.01 <41.61 5
4C 49.22 0.334 44.70 ± 0.01 8.50 ± 0.01 1 42.45 ± 0.01 <41.81 2
GQ Comae 0.165 44.15 ± 0.01 8.36 ± 0.03 0 42.58 ± 0.01 40.19 ± 0.30 1, 2
PKS 1217+023 0.240 45.13 ± 0.01 8.87 ± 0.08 2 42.78 ± 0.01 41.82 ± 0.16 2, 3
4C 25.40 0.268 44.73 ± 0.01 8.86 ± 0.03 2 42.66 ± 0.01 42.61 ± 0.02 2
3C 273 0.158 46.04 ± 0.09 9.06 1 42.87 ± 0.10 <41.73 1, 2, 3
SDSS 1230+6621 0.184 44.53 ± 0.01 8.26 ± 0.03 0 42.66 ± 0.01 41.90 ± 0.15 3
HE 1228+0131 0.117 44.95 ± 0.09 8.29 0 42.23 ± 0.11 <43.0 3
SDSS 1230+1100 0.236 44.73 ± 0.01 8.60 ± 0.05 0 42.64 ± 0.01 <43.0 3
PG 1307+085 0.154 44.80 ± 0.01 8.59 ± 0.05 0 42.69 ± 0.01 <41.06 2
B2 1425+267 0.364 45.14 ± 0.02 9.66 ± 0.03 2 42.99 ± 0.01 42.37 ± 0.04 1, 2
PG 1427+480 0.221 44.62 ± 0.01 7.95 ± 0.03 0 42.52 ± 0.01 <43.0 3
SDSS 1444+0633 0.208 44.48 ± 0.01 8.33 ± 0.03 0 42.44 ± 0.01 <43.0 3
HE 1453−0303 0.206 45.31 ± 0.12 8.59 0 42.54 ± 0.12 <43.0 3
SDSS 1507+4453 0.314 44.39 ± 0.01 7.34 ± 0.14 1 42.69 ± 0.01 <41.66 5
SDSS 1507+1338 0.322 44.79 ± 0.01 9.67 ± 0.02 0 42.53 ± 0.01 41.96 ± 0.08 5
4C 37.43 0.371 45.21 ± 0.01 9.37 ± 0.11 2 43.18 ± 0.01 43.02 ± 0.01 1,2
PG 1519+226 0.136 44.40 ± 0.01 7.90 ± 0.07 0 41.34 ± 0.09 <43.0 1
OR 241 0.254 44.86 ± 0.01 8.44 ± 0.05 1 42.39 ± 0.01 <41.85 2
3CR 323.1 0.265 45.20 ± 0.01 9.07 ± 0.03 2 42.99 ± 0.01 42.35 ± 0.13 1, 2, 3
4C 11.50 0.436 44.99 ± 0.01 8.87 ± 0.08 2 43.23 ± 0.01 42.68 ± 0.03 2
Ton 256 0.131 44.51 ± 0.01 7.99 ± 0.02 1 43.00 ± 0.01 42.80 ± 0.12 1, 3
SDSS 1655+2146 0.154 44.67 ± 0.01 8.28 ± 0.04 0 43.03 ± 0.01 43.03 ± 0.09 3
PG 1700+518 0.292 45.43 ± 0.09 8.98 0 42.63 ± 0.10 41.81 ± 0.18 2, 3
3CR 351 0.372 45.79 ± 0.01 9.81 ± 0.04 2 43.50 ± 0.01 <42.37 2
II Zw 136 0.061 44.40 ± 0.10 8.27 0 41.95 ± 0.09 40.90 ± 0.32 1, 3
PKS 2135−147 0.200 44.97 9.15 2 43.16 41.85 ± 0.04 2
HE 2152−0936 0.192 45.59 ± 0.10 8.76 0 42.17 ± 0.11 <43.0 3
HE 2158−0107 0.213 44.89 ± 0.11 8.60 0 42.59 ± 0.09 42.32 ± 0.13 3
HE 2158+0115 0.160 44.21 ± 0.10 7.94 0 42.01 ± 0.09 <43.0 3
4C 31.63 0.298 45.99 8.54 1 42.92 <42.08 1, 2
PG 2214+139 0.067 44.44 ± 0.11 8.64 0 41.81 ± 0.10 <43.0 3
PG 2233+134 0.326 45.19 ± 0.01 8.57 ± 0.31 0 42.50 ± 0.07 <40.58 2
PKS 2251+113 0.325 45.07 9.15 2 42.98 42.51 ± 0.04 1, 2
HE 2307−0254 0.221 44.89 ± 0.10 8.39 0 42.00 ± 0.11 <43.0 3
4C 09.72 0.433 45.17 9.30 2 42.80 <42.71 2
PKS 2349−014 0.174 44.77 ± 0.02 8.81 ± 0.03 2 42.45 ± 0.01 42.11 ± 0.11 3
HE 2353−0420 0.229 44.42 ± 0.10 8.12 0 42.58 ± 0.09 42.47 ± 0.12 3

Note. aRadio flag. 0: RQQ, 1: FSRQ (flat-spectrum or core-dominant radio quasar), 2: SSRQ (steep-spectrum or lobe-dominant radio quasar). References. (1) Boroson et al. 1985; (2) Stockton & MacKenty 1987; (3) Husemann et al. 2008 (including the re-analyzed data given in B. Husemann et al. 2012, in preparation); (4) Husemann et al. 2010; (5) This work.

Download table as:  ASCIITypeset image

We also collect the EELR measurements of 20 type-2 quasars (10 detections and 10 non-detections) presented by Humphrey et al. (2010) and Villar-Martín et al. (2010, 2011). Since they were selected from the SDSS type-2 quasars identified by Zakamska et al. (2003), luminosity and EW of emission lines arising from the narrow-line region (NLR) as well as optical ugriz magnitudes of the host galaxies are available in the archive. We derive the rest-frame gi color and stellar mass of the host galaxies from the SDSS model magnitudes and spectroscopic redshifts using a k-correction code developed by Blanton & Roweis (2007). The contributions of relatively strong [O ii] λ3727 and [O iii] λ4959, λ5007 lines from quasars are removed before the calculation by subtracting ΔmX = −2.5log (1 + fX, Y[EWY/wX]) from the X (u, g, r, i, or z) band magnitude if line Y is included in the band, where fX, Y is the filter transmission at line Y relative to the band average and wX is the bandwidth. We summarize the derived properties of the type-2 quasars in Table 3. All but two sources, SDSS 1228+0050 and SDSS 1625+3106, are radio-quiet (SDSS 1625+3106 is in the transition range of radio-quiet and radio-loud objects; Villar-Martín et al. 2011). In addition, we use all the type-2 quasars in Zakamska et al. (2003) and ∼60,000 galaxies extracted from the SDSS galaxy sample at z = 0.3–0.6, the redshift range of the main sample (the type-2 quasars in Table 3), as control objects. Their rest-frame color and stellar mass are calculated in the same way the objects in the main sample are calculated.

Table 3. Nuclear and Host Galaxy Properties of Type-2 Quasars with EELR Measurements

Quasar Redshift $\log L_{\rm N[{\rm O\,\mathsc{iii}}]}$ Rest-frame Stellar Mass EELRc Ref.
    (erg s−1)a gi (mag)b log M(M)    
SDSS 0025−1040 0.303 42.32 0.75 11.0 +1 3
SDSS 0123+0044 0.399 42.72 1.14 10.9 +1 2
SDSS 0217−0013 0.344 42.34 0.93 11.1 +1 3
SDSS 0234−0745 0.310 42.36 0.91 10.5 −1 3
SDSS 0840+3838 0.313 42.21 1.13 11.4 +1 1
SDSS 0849+0150 0.376 41.65 1.01 10.9 −1 3
SDSS 0920+4531 0.402 42.63 0.91 11.1 −1 1
SDSS 0955+0346 0.421 42.19 1.24 11.2 +1 3
SDSS 1153+0326 0.575 43.20 0.62 11.2 +1 3
SDSS 1228+0050 0.575 42.87 0.85 11.0 −1 3
SDSS 1307−0214 0.425 42.51 1.09 11.4 +1 3
SDSS 1337−0128 0.329 42.31 0.90 11.3 +1 3
SDSS 1407+0217 0.309 42.49 0.86 10.5 −1 3
SDSS 1413−0142 0.380 42.84 1.18 11.0 −1 3
SDSS 1546−0005 0.383 41.77 1.18 11.0 −1 3
SDSS 1550+3950 0.347 43.05 0.60 11.3 +1 1
SDSS 1625+3106 0.379 42.60 1.35 11.1 −1 1
SDSS 1726+6021 0.333 42.16 0.96 10.5 −1 1
SDSS 1739+5442 0.384 42.01 1.24 11.1 −1 1
SDSS 2358−0009 0.402 42.91 0.88 11.1 +1 3

Notes. aThe measurement error is <0.01. bTypical error of a few tens of magnitude is expected, including the uncertainties in the photometry and k-correction. cEELR detection flag (+1: detected, −1: non-detected). References. (1) Humphrey et al. 2010; (2) Villar-Martín et al. 2010; (3) Villar-Martín et al. 2011.

Download table as:  ASCIITypeset image

While the above samples of type-1 and type-2 objects are used below to probe the different aspects of quasars, we should keep in mind that the two populations can be intrinsically different. Elitzur (2012) clarified this point by demonstrating that the two types of AGNs are drawn preferentially from the distribution of dust-torus covering factors. Since type-2 AGNs have dustier environments around them than do type-1 counterparts, their host galaxies might tend to have higher star formation rates since (1) the presence of ample dust indicates active star formation producing them and (2) the negative feedback (suppression of star formation) by the AGN radiation might be weaker due to the dust obscuration.

3. EMERGENCE OF EELR

3.1. Notes on the New Sample

The Suprime-Cam [O iii] line images of the newly observed quasars are presented in Figure 2. To aid visual inspection, the gray scale consistently represents linear flux levels between zero and quasar peak values, while the white crosses indicate the PSF sizes (twice the full widths at half-maximum (FWHMs)). The most spectacular feature is found around SDSS 1507+1338. The emission feature, extending across ∼50 kpc along the NE–SW direction, is reminiscent of the ionization cones created by AGN radiation. SDSS 0914+0507 also has apparently extended emission, with a knot ∼35 kpc away from the center in the NW direction. On the other hand, the radial profiles of SDSS 0954+0929, SDSS 1139+4020, and SDSS 1507+4453 are consistent with the PSFs and show no signs of EELRs. We measure the EELR luminosity in an annulus of inner radius 10 kpc and outer radius 30 kpc. The wing components of the nuclear radiations are estimated from the PSFs and subtracted. The incomplete transmission of the [O iii] lines through the NA656 filter is also corrected using the filter transmission function. We include in the error budget of the derived luminosity the uncertainties in the continuum subtraction and in the correction for the wing contribution of the nuclear radiation, background sky noise, and the possible flux variation assumed to be 10% due to the non-photometric sky condition. The results are listed in the last column of Table 1. The above measurements support our visual inspection finding detectable EELRs only in SDSS 0914+0507 and SDSS 1507+1338. The nuclear [O iii] luminosity ($L_{{\rm N[{\rm O\,\mathsc{iii}}]}}$) is also measured within a 3 arcsec aperture and is found to agree with the SDSS spectroscopic measurements within $\Delta \log L_{{\rm N[{\rm O\,\mathsc{iii}}]}} = \pm 0.1$.

Figure 2.

Figure 2. Suprime-Cam [O iii] line images of (a) SDSS 0914+0507, (b) SDSS 0954+0929, (c) SDSS 1139+4020, (d) SDSS 1507+4453, and (e) SDSS 1507+1338. The gray scale represents linear flux levels between zero and the quasar peak values, while the white crosses indicate twice the FWHMs of the PSFs. The physical scale of the panels is 50 kpc on a side. North is up, east is to the right.

Standard image High-resolution image

3.2. Link to AGNs

As described in Section 1, the presence of EELRs correlates with some AGN properties. Quasars with larger [O iii] EW, broader and bumpier Balmer lines, fainter Fe ii, and SSRQ-like radio emission have a higher possibility of accompanying EELRs around them. The correlation is fairly strong, as Fu & Stockton (2009) showed that if certain quasars are pre-selected based on their radio and nuclear [O iii] emissions, ∼50% of them would turn out to have EELRs as opposed to ∼20% with no a priori assumption. Note that the above emissions arise from different parts of AGNs: broad Balmer lines and the Fe ii lines from parsec-scale BLR, the [O iii] lines from kiloparsec-scale NLR, and radio and EELR emissions from more extended, galaxy-wide scales. The correlation among them, except for EELRs, has been well known since the seminal work by Boroson & Green (1992). With the PC analysis technique, they showed that most of the variance in AGN spectral properties is contained in two sets of correlations. A strong anti-correlation between the measures of [O iii] and Fe ii is PC 1, or eigenvector 1, with which the Balmer line profile is also associated. The general consensus is that PC 1 is driven mainly by Eddington ratio (e.g., Boroson 2002; Sameshima et al. 2011). In Figure 3, we plot Hβ line width versus the relative Fe ii strength $R_{{\rm Fe\,\mathsc{ii}}}$ ≡ EW(Fe ii)/EW(Hβ) of the SDSS quasars in Shen et al. (2011). Both measures constitute the four-dimensional eigenvector 1 (4DE1; other measures are the soft X-ray photon index and C iv λ1549 line profile) space proposed by Sulentic et al. (2000). The anti-correlation between them is evident in Figure 3, which forms a sequence of Eddington ratios. The data scatter reflects the orientation effects as well as the intrinsic variety of quasars, since one would observe larger line FWHM and $R_{{\rm Fe\,\mathsc{ii}}}$ (Fe ii is emitted from the outermost part of BLR; e.g., Matsuoka et al. 2008) as the viewing angle relative to the dust-torus plane decreases. On the other hand, PC 2 links optical luminosity and He ii λ4686 line emission. Its physical origin is thought to be mass accretion rate to SMBHs because of its strong correlation with optical luminosity. Therefore, one would expect that the ratio of PC 1 and PC 2 is proportional to SMBH mass, which was actually proved by Boroson (2002). Furthermore, Laor (2000) found that radio loudness is strongly related to SMBH mass as radio-loud/radio-quiet quasars are associated with most/least massive SMBHs. The above arguments suggest that most of the AGN properties can be understood on the PC 1–PC 2 plane.

Figure 3.

Figure 3. Hβ line width (FWHM) vs. the relative Fe ii strength $R_{{\rm Fe\,\mathsc{ii}}}$ = EW(Fe ii)/EW(Hβ) of the SDSS quasars. The red, green, and black dots represent the objects with high (−0.8 < log [Lbol/LEdd]), middle (−1.2 < log [Lbol/LEdd] < −0.8), and low (log [Lbol/LEdd] < −1.8) Eddington ratio values, respectively. The typical error of each data point is indicated at the top right corner.

Standard image High-resolution image

We plot the Eddington ratio and continuum luminosity of the compiled type-1 sample, as well as all of the SDSS quasars in the Shen et al. (2011) catalog, in Figure 4. First of all, we remark that the objects with EELR measurements cover the distribution of all of the SDSS quasars fairly uniformly on this plane. This was not anticipated since the existing EELR observations are subject to strong selection biases individually. We see the clear separation of radio-quiet and radio-loud sources as pointed out by Laor (2000), the latter being associated with most massive SMBHs. In addition, SSRQs are observed to have higher values of Eddington ratio or SMBH mass than FSRQs. We also find some correlations containing EELRs. They are shown more clearly in Figure 5 where the fraction of EELR detection is plotted as a function of the nuclear properties. The EELR fraction is strongly anti-correlated with the Eddington ratio, and its weak correlation with SMBH mass may also be present. The former relation is most likely the underlying basis of the correlation between EELR and PC 1. While no clear trend is found with respect to radio emission, we do find that EELR luminosity is systematically higher in SSRQs than in the rest of the sample; the mean luminosity is $L_{\rm E[{\rm O\,\mathsc{iii}}]}$ = 1042.4 and 1041.9 erg s−1 for the SSRQs and RQQs, respectively. The difference will be even larger if we consider the fact that the EELR luminosity of most SSRQs measured by Stockton & MacKenty (1987) is underestimated since the emission in a central 10 kpc aperture is excluded. This explains why the EELR fraction around SSRQs is much higher than those around the other populations in the pioneering work of Stockton & MacKenty (1987).

Figure 4.

Figure 4. Eddington ratio vs. continuum luminosity of the compiled type-1 sample (filled circles) and the whole of the SDSS quasars (dots). RQQs, FSRQs, and SSRQs in the compiled sample are represented by black, green, and red colors, respectively. The quasars with detectable EELRs are marked with the larger circles whose sizes are proportional to $\log L_{\rm E[{\rm O\,\mathsc{iii}}]}$. PC 1 and PC 2 increase along the horizontal and vertical axes on this plane, while SMBH mass increases from the top left to the bottom right corner. The lines of constant SMBH mass, log MBH = 8.0, 8.5, 9.0, 9.5, are shown by the dotted lines.

Standard image High-resolution image
Figure 5.

Figure 5. Fraction of EELR detection as a function of Eddington ratio (a), SMBH mass (b), continuum luminosity (c), and radio emission property (d). The fractions are calculated in the bins represented by the horizontal bars. The number of objects in each bin is also indicated.

Standard image High-resolution image

One of the highest dependencies of EELRs is found on nuclear [O iii] luminosity, $L_{\rm N[{\rm O\,\mathsc{iii}}]}$. In Figure 6, we show Eddington ratio versus $L_{\rm N[{\rm O\,\mathsc{iii}}]}$ of the same samples. It is clearly seen that the quasars with EELR measurements are biased to high-$L_{\rm N[{\rm O\,\mathsc{iii}}]}$ compared to the SDSS quasars. Here, we find that low-$L_{\rm N[{\rm O\,\mathsc{iii}}]}$ quasars tend to lack EELRs: only 3 out of 20 objects with $L_{\rm N[{\rm O\,\mathsc{iii}}]} < 10^{42.5}$ erg s−1 have detectable, faint EELRs, while the EELR fraction is >50% at the higher $L_{\rm N[{\rm O\,\mathsc{iii}}]}$. This trend was reported previously for smaller samples (e.g., Stockton & MacKenty 1987). Figure 6 also tells us that $L_{\rm N[{\rm O\,\mathsc{iii}}]}$ does not strongly correlate with the Eddington ratio, although it correlates with PC 1 (Boroson & Green 1992). No clear correlation between the two measures indicates that PC 1 is not solely driven by the Eddington ratio—there should be another physical mechanism(s) that links nuclear [O iii] luminosity with other PC 1 constituents.

Figure 6.

Figure 6. Eddington ratio vs. nuclear [O iii] line luminosity of the compiled type-1 sample and all of the SDSS quasars. The symbols are as in Figure 4. The dotted line represents $L_{\rm N[{\rm O\,\mathsc{iii}}]}$ = 1042.5 erg s−1.

Standard image High-resolution image

What can we learn from these correlations? While the clear anti-correlation between EELRs and the Eddington ratio is observed, it does not seem to reflect their direct relationship through the AGN radiation since (1) the EELR fraction is highest for the AGNs with the most inefficient gas accretion to SMBHs, (2) nuclear [O iii] luminosity produced similarly by AGN radiation shows no clear correlation with Eddington ratio, and (3) a positive correlation between the EELR fraction and continuum luminosity is not observed. Therefore, it is more reasonable to consider the presence of an independent mechanism that leads to the high/low EELR fraction and low/high Eddington ratio at the same time. The dependence of EELRs on radio emission and (possibly) SMBH mass may also be related to this hidden mechanism. We will return to this issue below. The lack of a simple relation between EELRs and continuum luminosity (accretion rate) was also reported by Fu & Stockton (2009), who suggested that the correlation between nuclear and extended [O iii] emissions, which is also evident in our larger sample, may be the sequence of the available gas amount rather than AGN emission characteristics. The hypothesis that the gas availability is the primary determinant of EELR creation is supported by the probe of host galaxies in the following section.

3.3. Link to Host Galaxies

If EELRs represent a certain stage of galaxy evolution, as the AGN feedback scenario implies, their presence may be related to stellar population properties of host galaxies. Here, an investigation of such a relation is presented for the first time, thanks to the recent measurements of EELRs around SDSS type-2 quasars at z = 0.3–0.6 (Humphrey et al. 2010; Villar-Martín et al. 2010, 2011). In Figure 7, we plot stellar mass versus rest-frame gi color of the type-2 quasars as well as the control samples defined in Section 2.2. Most SDSS galaxies at these relatively high redshifts are luminous red galaxies (LRGs), forming a clump at the massive red end on this color–stellar mass diagram (we could see the red sequence rather than a clump if the SDSS observations were deep enough to detect fainter galaxies). The type-2 quasars are bluer than the LRGs (see also Zakamska et al. 2003) but they are redder than the star-forming galaxies that form the "blue cloud" around gi ∼ 0.6 in the local universe (e.g., Gavazzi et al. 2010). This is consistent with the result of Schawinski et al. (2010), who found that local AGNs reside in the "green valley" between the blue cloud and the red sequence on the color–stellar mass diagram. Interestingly, we find that the EELR quasars are associated with bluer and more massive galaxies compared to their non-EELR counterparts. In order to show this trend more clearly, we define a "massive-blueness" indicator, $\mathcal {I}_{\rm mb} \equiv \log M_{\star }$ − (gi), which increases diagonally toward the bottom right corner of Figure 7 as shown by the arrow. The number distribution of the type-2 quasars with EELR measurements as a function of $\mathcal {I}_{\rm mb}$ is plotted in Figure 8. It shows a clear, monotonic increase in the fraction of EELR quasars as $\mathcal {I}_{\rm mb}$ increases: only 1/8 of the lowest-$\mathcal {I}_{\rm mb}$ quasars have EELRs, while the fraction rises to 7/8 in the highest-$\mathcal {I}_{\rm mb}$ counterparts.

Figure 7.

Figure 7. Stellar mass vs. rest-frame gi color of the galaxies hosting type-2 quasars with EELR measurements (green circles). The objects with detectable EELRs are marked with the larger circles. The type-2 quasars of Zakamska et al. (2003) at z = 0.3–0.6 are represented by the black circles, while the distribution of ∼60,000 SDSS galaxies in the same redshift range is shown by the contours (the contour levels are logarithmic). The arrow represents the "massive blueness" indicator $\mathcal {I}_{\rm mb}$ used in Figure 8.

Standard image High-resolution image
Figure 8.

Figure 8. Number distribution of the type-2 quasars with EELR measurements as a function of the "massive blueness," $\mathcal {I}_{\rm mb} = \log M_{\star }$ − (gi). The hatched area is for the EELR quasars.

Standard image High-resolution image

What is the origin of this dichotomy? Blue colors point to current or recent star formation activity, which in turn indicates the presence of a certain amount of gas in the galaxies. Based on the velocity structure and mass of EELR gas around SSRQs, as well as the correlation between EELRs and radio emission, Fu & Stockton (2009) suggested that EELRs are formed in galaxy minor mergers accompanied by the production of radio jets associated with AGNs. A similar conclusion was reached by Husemann et al. (2010) for an RQQ. Villar-Martín et al. (2011) found that the majority of the type-2, radio-quiet EELR quasars show the signs of galaxy merger or interaction, while none of the non-EELR counterparts show such signs. These observations suggest that the gas may be brought in by the merging companions. The merger/interaction possibly activates the observed star formation as well as the AGN that ionizes the gas and creates the EELR, since the interstellar gas of merging galaxies can be effectively brought into the nuclear regions with rapid loss of the angular momentum (e.g., Barnes & Hernquist 1996). The fact that EELRs are preferentially found in gas-rich, massive blue galaxies is consistent with both the above scenario and the previous arguments that the presence of an EELR is regulated by the amount of available gas. Furthermore, massive galaxies are known to be the dominant hosts of radio-loud AGNs (e.g., Floyd et al. 2004; Best et al. 2005). It naturally explains the EELR–radio emission connection described above, and can also provide the "hidden" mechanism behind the anti-correlation between EELRs and the Eddington ratio since the majority of radio-loud AGNs are powered by inefficiently accreting, most massive SMBHs (see Figure 4).

The recent kinematic observations have revealed the presence of galaxy-wide, disturbed gas in quasar host galaxies (e.g., Fu & Stockton 2009; Greene et al. 2011). While only a small fraction of the gas is generally found to have high enough velocity to escape the galaxies, Greene et al. (2011) showed that these estimates are in fact lower limits and may be consistent with expectations of recent AGN feedback models. If the feedback scenario is correct, the gas responsible for EELR and star formation (and hence the blue colors of galaxies) would soon be swept away by the AGN-driven, galactic winds. In this regard, it may be noteworthy that none of the most luminous (log λLλ5100 ⩾ 45.5) type-1 quasars in our sample have EELRs, as shown in Figure 4. It is tempting to raise the possibility of the gas stripping, but we have to be aware that (1) the relatively high Eddington ratio for these objects may be responsible for the deficit of EELRs, and (2) the bright central radiation can prevent us from detecting EELRs in the most luminous sources. More data are needed for further inspection of this issue. While the process discussed above corresponds to the so-called "quasar-mode" feedback, it is interesting to point out that a clear interaction between radio jets and EELR clouds with high-velocity dispersion is observed in SSRQs (Fu & Stockton 2009), which invokes another "radio mode" of the feedback.

In the nearby universe, Keel et al. (2012) investigated the properties of Seyfert galaxies with EELRs dominated by type-2 objects. While their morphology supports the merger origin of EELRs, they have relatively quiescent kinematics without a sign of high-velocity gas leaving the galaxies seen in higher-luminosity AGNs. This may simply indicate that the creation of the powerful galactic winds requires quasar-class energy input into the interstellar medium; hence, the relation of EELR to AGNs may be different at high and low luminosities.

4. SUMMARY AND CONCLUSIONS

A comprehensive analysis of EELR phenomenon around quasars is presented in this paper. We compile the past EELR measurements for type-1 (Boroson et al. 1985; Stockton & MacKenty 1987; Husemann et al. 2008) and type-2 (Humphrey et al. 2010; Villar-Martín et al. 2010, 2011) quasars and combine them with the new observation of five quasars at z ∼ 0.3. The new observation was carried out using the Subaru/Suprime-Cam with the narrowband NA656 filter (corresponding to the redshifted [O iii] λ5007) and led to the discovery of EELR around two sources, SDSS 091401.75 + 050750.6 and SDSS 150752.66 + 133844.5. The observation effectively complements the existing measurements on the AGN PC plane. The properties of associated AGNs (SMBH mass, continuum and line luminosity, and radio characteristics) and host galaxies (stellar mass and rest-frame color) are collected and calculated from the SDSS and other observations.

We find that EELR anti-correlates with the Eddington ratio, which is most likely the underlying basis of the known correlations between EELR and the PC 1 (eigenvector 1) constituents. EELR and nuclear [O iii] emissions are also highly correlated. We argue that the primary determinant of EELR creation is the gas availability, rather than AGN emission characteristics. Recent observations (Fu & Stockton 2009; Husemann et al. 2010; Villar-Martín et al. 2011) revealed that a significant fraction of EELR galaxies show signs of a recent minor merger, which may have introduced the gas and activated star formation as well as the AGN that creates EELR. In line with the above arguments, we find that EELR is preferentially associated with gas-rich, massive blue galaxies. This may explain the whole set of observed correlations containing EELR, since massive galaxies are the dominant hosts of radio-loud AGNs characterized with low Eddington ratio and high SMBH mass.

The hierarchical structure formation models based on the ΛCDM theory predict that galaxies are assembled through a sequence of mergers of smaller building blocks. In Figure 9, we show a schematic view of the color–stellar mass diagram, or equivalently the color–magnitude diagram (CMD), of observed galaxies. Once a galaxy grows up to the massive tip of the blue cloud (represented by the seven-rayed star in the figure), it has two ways to evolve further. If no subsequent merger happens for a sufficient period of time, the galaxy would migrate into the low-mass end of the red sequence as the star formation fades out. It can then move up the red sequence toward the massive end through dry merger with gas-poor galaxies (see also Faber et al. 2007). On the other hand, a wet merger of the galaxy with gas-rich galaxies at any stage of the above evolution could push the galaxy into the green valley where the AGN fraction is the highest. If the galaxy (and hence the central SMBH, assuming the positive bulge mass–SMBH mass correlation) is massive enough and the merging events involve an ample supply of gas for triggering vigorous star formation, the galaxy would be driven to the massive blue realm at the extreme corner of the valley. This event would also activate the AGN, which ionizes the gas and creates luminous EELR. The galaxy-wide energy injection by the AGN radiation might then blow away the gas and quench the star formation, pushing the galaxy back to the red sequence. While this last argument is less certain due to the lack of evidence about the energy injection process, recent observations have started to reveal the presence of AGN-driven, galaxy-scale outflows that may be quenching the star formation there (e.g., Greene et al. 2011; Rupke & Veilleux 2011; Cano-Diaz et al. 2012).

Figure 9.

Figure 9. Schematic view of the galaxy color–stellar mass diagram, or equivalently CMD. The blue, red, and green regions represent the blue cloud of star-forming galaxies, the red sequence of quiescent galaxies, and the green valley occupied by AGN host galaxies, respectively. The implied location of EELR galaxies is also shown by the yellow triangle. The bulge-mass–SMBH mass correlation indicates that SMBH mass increases toward the right, hence the habitats of RQQs and SSRQs are separated in the green valley as indicated. The lower right corner of this diagram enclosed by the dotted lines seems to be the "forbidden" area for observed galaxies. The seven-rayed star marks the starting point of the galaxy evolution discussed in the text. The blue and red arrows represent wet and dry mergers, respectively, while the open arrows represent the stable color change due to star formation fading. The filled parts of the open arrows show the contribution of the violent star formation quenching by the AGN feedback.

Standard image High-resolution image

The above AGN feedback scenario gives a solution to the major problems of the classical hierarchical galaxy formation models, i.e., the overproduction of massive galaxies and the inverted color–magnitude–morphology relation (see Section 1), by suppressing star formation in massive galaxies. However, many details of the process are still unknown and are debated. What is observationally clear is that EELR is found at the exact region on the CMD where the AGN feedback is expected to work—the massive blue corner that seems to be the forbidden area for observed galaxies. This provides evidence for the presence of the AGN feedback, which may be playing a leading role in the co-evolution of galaxies and central SMBHs.

The author is grateful to B. Husemann for kindly providing the unpublished data on re-analysis of the EELR measurements. An anonymous referee has provided many suggestive comments to improve the manuscript. Great assistance was provided for the Subaru/Suprime-cam observation by S. Masaki, R. Asano, and the staff at the Hawaii observatory, NAOJ. This work was supported by a Grant-in-Aid for Young Scientists (22684005) and by the Global COE Program of Nagoya University "Quest for Fundamental Principles in the Universe" from JSPS and MEXT of Japan.

Footnotes

  • This work is based in part on data collected at Subaru Telescope, which is operated by the National Astronomical Observatory of Japan (NAOJ).

  • Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participating Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel, University of Cambridge, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington.

Please wait… references are loading.
10.1088/0004-637X/750/1/54