Abstract

We present our latest results on the connection between accretion rate and relativistic jet power in active galactic nuclei (AGN), by using a large sample which includes mostly blazars, but contains also some radio galaxies. The jet power can be traced by γ-ray luminosity in the case of blazars, and radio luminosity for both classes. The accretion-disc luminosity is instead traced by the broad emission lines. Among blazars, we find a correlation between broad line emission and the γ-ray or radio luminosities, suggesting a direct tight connection between jet power and accretion rate. We confirm that the observational differences between blazar subclasses reflect differences in the accretion regime, but with blazars only we cannot properly access the low-accretion regime. By introducing radio galaxies, we succeed in observing the fingerprint of the transition between radiatively efficient and inefficient accretion discs in the jetted AGN family. The transition occurs at the standard critical value Ld/LEdd ∼ 10−2 and it appears smooth. Below this value, the ionizing luminosity emitted by the accretion structure drops significantly.

INTRODUCTION

Blazars are active galactic nuclei (AGN) with a relativistic jet directed towards our line of sight. Because of relativistic beaming, the emission from the jet is highly boosted, and dominates the AGN emission at all wavelengths, from the radio to the γ-ray band. They are very useful to study the relation between accretion structure and relativistic jet in AGN.

Blazars are commonly divided into Flat Spectrum Radio Quasars (FSRQs) and BL Lacertae objects (BL Lacs), depending on the rest-frame equivalent width (EW) of their broad emission lines. Specifically, if the EW is smaller than 5 Å, the object is classified as a BL Lac, otherwise as an FSRQ (Urry & Padovani 1995). An analogous classification criterion was introduced by Landt et al. (2004) that found possible to discriminate between objects with intrinsically weak or strong narrow emission lines by studying the [O ii] and [O iii] EW plane. The EW-based classification, besides being simple to apply, is based on the fact that the EW is a good measure of the line emission dominance over the non-thermal continuum emitted from the jet. In this view, the EW could tell if an object had intrinsically strong (FSRQ) or weak (BL Lac) emission lines. However, jet emission is extremely variable, definitely more than the thermal continuum and the emission lines. Hence, the line EW can dramatically vary from one state to another for the same source. A blazar with intrinsically very luminous emission lines can temporarily appear as a BL Lac, with very small EW, if its jet flux happens to be more luminous than usual. On the contrary, during a particularly low state, a BL Lac can show emission lines with EW larger than the 5 Å limit (as it happened to BL Lac itself; Vermeulen et al. 1995; Capetti, Raiteri & Buttiglione 2010). Ghisellini et al. (2011, hereafter G11) already proposed that this classification is not reliable, and moreover does not reflect any intrinsic property or difference within the blazar class, being dependent on the strong continuum variability. These authors introduced a more physical classification, based on the different accretion rates of the two subclasses of blazars, later confirmed in our earlier work (Sbarrato et al. 2012; hereafter TS12): FSRQs have a disc luminosity ≳5 × 10−3 of the Eddington one. In those papers, a strong correlation between accretion and jet emissions was found. The unreliability of the EW classification was also deeply investigated by Giommi et al. (2012) and Giommi, Padovani & Polenta (2013), who suggested that blazars should be divided into high- and low-ionization sources, instead of focusing on observed features that are not physically relevant. In other words, G11, TS12 and Giommi et al. (2012, 2013) based their reclassification on the ionizing luminosity, emitted from the accretion disc.

Radio galaxies are thought to be the parent population of blazars: while blazars are aligned to our line of sight, radio galaxies have their jets oriented at larger viewing angles. This greatly affects the overall emission and the typical spectral energy distribution (SED), that is not dominated by the non-thermal radiation in the case of radio galaxies, contrary to blazars (Urry & Padovani 1995). Radio galaxies are historically divided into two subclasses according to their radio morphology (Fanaroff & Riley 1974): Fanaroff–Riley type I (FR I) show bright jets close to the nucleus, while Fanaroff–Riley type II (FR II) show prominent hotspots far from it. This classification also reflects in a separation in their radio power (below and above L178 MHz = 2.5 × 1033 erg s−1 Hz−1, respectively). As in the case of blazars, the classification is not sharp, but radio galaxies seem to be distributed continuously between the two classes. Similarly to the blazar case, radio galaxies were also classified according to their ionization efficiency. In the nineties many spectroscopy-based classifications were proposed, also connecting the emission line luminosities to the radio power (e.g. Baum & Heckman 1989a,b; Rawlings et al. 1989; Rawlings & Saunders 1991; Morganti et al. 1997; Willott et al. 1999; Labiano 2008). Ghisellini & Celotti (2001) showed that the division between FR I and FR II actually reflected a systematic difference in accretion rate: FR I were shown to have generally an ionizing luminosity ∼10−2–10−3 of the Eddington one, while for FR II this was typically larger. Laing et al. (1994) introduced a subclassification of FR II sources into high-excitation (HEG) and low-excitation galaxies (LEGs), but this was found to be applicable also to some FR I. Therefore, Buttiglione et al. (2009, 2010) decided to perform an extended investigation of the spectroscopic properties of radio galaxies, using a homogeneous sample. They found that all HEG are FR II, while LEGs can be both FR I and FR II. While an ionization-based classification seems to be more physically relevant also in the case of radio galaxies, a univocal, physically based classification method is not yet commonly accepted, as in the case of blazars.

In this work, we try to investigate the relation between accretion and jet emissions in jetted AGN, to understand if a change in the accretion mode happens inside the blazar family, and the whole jetted AGN class. We enlarge the blazar sample with respect to the one in TS12, and we extend our study to another tracer of the jet power: radio luminosity. In this way, we can also include radio galaxies to study the jet-accretion relation in the whole family of jetted AGN. Section 2 presents the samples, Section 3 describes how we trace the accretion luminosity through the broad line region (BLR) emission. In Section 4, we investigate the relationship between BLR and γ-ray power, while the one between BLR and radio power is dealt with in Section 5. The discussion is presented in Section 6, and our results are summarized in Section 7.

THE SAMPLE

We are interested in studying the relationship between jet and accretion in AGN, through γ-ray or radio luminosity and BLR luminosity, respectively. We first tried to study it among the blazar subclass, by collecting a complete sample of γ-ray detected blazars, with measurements of the luminosity of broad lines obtained through optical spectroscopy. We then extended our sample to radio galaxies intrinsically without broad emission lines (but with known redshift), to investigate the jet-accretion relationship in the inefficient accretion regime. Even if not complete, the radio-galaxy sample we describe in Section 2.2 is the most useful for our purposes, as the ionization status of every member is thoroughly studied.

The blazar samples

The Large Area Telescope (LAT) onboard the Fermi Gamma-Ray Space Telescope (Fermi) has detected a large amount of blazars in the γ-rays. In two different papers (Shaw et al. 20122013, respectively, S12 and S13 hereafter), Shaw and collaborators obtained optical spectroscopic data for all the FSRQs (S12) and most of the BL Lacs (S13) included in the Second Catalog of AGN detected by the Fermi LAT (2LAC; Ackermann et al. 2011). We collected from these two papers all the blazars that show broad emission lines in their optical spectra.

In S12, the authors analysed the optical spectra of 229 FSRQs: they obtained new spectra for 165 FSRQs included in the First Catalog of AGN detected by Fermi (1LAC; Abdo et al. 2010), and re-analysed Sloan Digital Sky Survey spectra of 64 other FSRQs (not all with γ-ray data). Along with spectroscopic data, the authors derived virial black hole masses for all their objects. In order to have a complete description of the sources, we selected only the FSRQs with enough data to fit the entire SED. We are left with 191 objects. We cross-correlated this sample with the 2LAC. When the objects were not included in the 2LAC, we collected data from the 1LAC. The sample was completed including radio data from the Combined Radio All-Sky Targeted Eight GHz Survey (CRATES; Healey et al. 2007). When sources were not included in CRATES (only five objects), we obtained radio data at frequencies close to 8 GHz from the ASI Science Data Center (ASDC). We are then left with a sample of 180 FSRQs with optical spectroscopy, along with γ-ray and radio data.

In S13, instead, the BL Lac objects are directly selected from the 2LAC sample. 2LAC includes 475 BL Lacs, among which only 209 have redshift information. We consider here only objects with broad emission lines, in order to be able to derive an estimate of LBLR, avoiding upper limits. We are left with 26 objects. All of them have 2LAC γ-ray and CRATES radio data. S13 did not derive the black hole masses for these objects. We can anyway assign an average MBH value to these BL Lacs, from an adapted version of the MBHMR relation (from Bettoni et al. 2003), that assumes a mass ratio between bulge and central black hole of 103. In fact, BL Lacs are typically hosted in massive and luminous elliptical galaxies, with a small dispersion in absolute magnitude (〈MR〉 = −22.8 ± 0.5; Sbarufatti, Treves & Falomo 2005). Therefore, we can derive an average value of their black hole mass:
(1)
Since we need the central black hole masses for our studies, we will apply this value to every object without a mass estimate. These BL Lacs and their relevant data are listed in Table 1.
Table 1.

Sources from the S13 BL Lac sample, divided according to their reclassification (see Section 4). Column 1: name; column 2: Fermi/LAT counterpart; column 3: right ascension; column 4: declination; column 5: redshift; column 6: line measured, from which LBLR has been derived; column 7: logarithm of the BLR luminosity (in erg s−1); column 8: logarithm of the γ-ray luminosity from Fermi data (in erg s−1); column 9: logarithm of the radio luminosity, calculated at 8 GHz rest frame (in erg s−1).

NameFermi nameRADec.zLinelog LBLRlog Lγlog Lradio
[1][2][3][4][5][6][7][8][9]
BL Lac
GB6 J0013+19102FGLJ0013.8+190700 13 56.3+19 10 41.50.477Mg ii42.69145.44143.043
PKS 0829+0462FGLJ0831.9+042908 31 48.7+04 29 38.20.174Ha42.61445.52042.923
RBS 09582FGLJ1117.2+201311 17 06.1+20 14 07.60.138Ha41.72244.72341.622
PMN J1125-35562FGLJ1125.6−355911 25 31.3−35 57 05.00.284Ha43.33845.15742.627
SBS 1200+6082FGLJ1203.2+603012 03 03.4+60 31 19.10.065Ha42.00543.76941.190
W Comae2FGLJ1221.4+281412 21 31.6+28 13 58.10.103Ha42.13745.05542.199
PG 1218+3042FGLJ1221.3+301012 21 21.9+30 10 36.20.184Ha42.06345.17441.734
OQ 5302FGLJ1420.2+542214 19 46.5+54 23 15.00.153Ha42.20144.76642.603
RGB J1534+3722FGLJ1535.4+372015 34 47.2+37 15 53.80.144Ha41.72244.34940.991
BL/FS
PKS 0332−4032FGLJ0334.2−400803 34 13.4−40 08 26.91.357Mg ii45.04647.71144.967
TXS 0431−2032FGLJ0434.1−201404 34 07.9−20 15 17.20.928Mg ii43.15346.51143.758
PKS 0437−4542FGLJ0438.8−452104 39 00.7−45 22 23.92.017iv45.20147.66145.363
PKS 0627−1992FGLJ0629.3−200106 29 23.7−19 59 19.71.724iv44.06047.74045.036
4C +14.602FGLJ1540.4+143815 40 49.5+14 47 46.50.606Mg ii43.57545.96644.229
PMN J1754−64232FGLJ1755.5−642317 54 41.8−64 23 44.71.255Mg ii44.16346.91244.113
4C +56.272FGLJ1824.0+565018 24 07.0+56 51 01.10.664Mg ii43.91246.89144.333
S3 2150+172FGLJ2152.4+173521 52 24.7+17 34 37.90.874Mg ii44.16846.33344.310
PMN J2206−00312FGLJ2206.6−002922 06 43.2−00 31 02.31.053Mg ii43.79846.55743.858
B2 2234+28A2FGLJ2236.4+282822 36 22.3+28 28 58.10.79Mg ii44.64547.07944.412
PKS 2244−0022FGLJ2247.2−000222 47 30.1+00 00 07.00.949Mg ii44.10646.54343.950
PKS 2312−5052FGLJ2315.7−501423 15 44.2−50 18 39.70.811Mg ii43.62846.35743.764
PKS 2351−3092FGLJ2353.5−303423 53 47.3−30 37 48.30.737Mg ii43.62846.06643.893
FS
NVSS J020344+3042382FGLJ0204.0+304502 03 44.1+30 42 38.10.761Mg ii44.75746.44943.572
PKS 0516−6212FGLJ0516.8−620705 16 44.5−62 07 04.81.3Mg ii44.44447.48844.523
MG2 J201534+37102FGLJ2015.6+370920 15 28.6+37 10 59.80.859Hb44.34247.97144.794
TXS 2206+6502FGLJ2206.6+650022 08 03.3+65 19 38.71.121Mg ii44.33647.32444.360
NameFermi nameRADec.zLinelog LBLRlog Lγlog Lradio
[1][2][3][4][5][6][7][8][9]
BL Lac
GB6 J0013+19102FGLJ0013.8+190700 13 56.3+19 10 41.50.477Mg ii42.69145.44143.043
PKS 0829+0462FGLJ0831.9+042908 31 48.7+04 29 38.20.174Ha42.61445.52042.923
RBS 09582FGLJ1117.2+201311 17 06.1+20 14 07.60.138Ha41.72244.72341.622
PMN J1125-35562FGLJ1125.6−355911 25 31.3−35 57 05.00.284Ha43.33845.15742.627
SBS 1200+6082FGLJ1203.2+603012 03 03.4+60 31 19.10.065Ha42.00543.76941.190
W Comae2FGLJ1221.4+281412 21 31.6+28 13 58.10.103Ha42.13745.05542.199
PG 1218+3042FGLJ1221.3+301012 21 21.9+30 10 36.20.184Ha42.06345.17441.734
OQ 5302FGLJ1420.2+542214 19 46.5+54 23 15.00.153Ha42.20144.76642.603
RGB J1534+3722FGLJ1535.4+372015 34 47.2+37 15 53.80.144Ha41.72244.34940.991
BL/FS
PKS 0332−4032FGLJ0334.2−400803 34 13.4−40 08 26.91.357Mg ii45.04647.71144.967
TXS 0431−2032FGLJ0434.1−201404 34 07.9−20 15 17.20.928Mg ii43.15346.51143.758
PKS 0437−4542FGLJ0438.8−452104 39 00.7−45 22 23.92.017iv45.20147.66145.363
PKS 0627−1992FGLJ0629.3−200106 29 23.7−19 59 19.71.724iv44.06047.74045.036
4C +14.602FGLJ1540.4+143815 40 49.5+14 47 46.50.606Mg ii43.57545.96644.229
PMN J1754−64232FGLJ1755.5−642317 54 41.8−64 23 44.71.255Mg ii44.16346.91244.113
4C +56.272FGLJ1824.0+565018 24 07.0+56 51 01.10.664Mg ii43.91246.89144.333
S3 2150+172FGLJ2152.4+173521 52 24.7+17 34 37.90.874Mg ii44.16846.33344.310
PMN J2206−00312FGLJ2206.6−002922 06 43.2−00 31 02.31.053Mg ii43.79846.55743.858
B2 2234+28A2FGLJ2236.4+282822 36 22.3+28 28 58.10.79Mg ii44.64547.07944.412
PKS 2244−0022FGLJ2247.2−000222 47 30.1+00 00 07.00.949Mg ii44.10646.54343.950
PKS 2312−5052FGLJ2315.7−501423 15 44.2−50 18 39.70.811Mg ii43.62846.35743.764
PKS 2351−3092FGLJ2353.5−303423 53 47.3−30 37 48.30.737Mg ii43.62846.06643.893
FS
NVSS J020344+3042382FGLJ0204.0+304502 03 44.1+30 42 38.10.761Mg ii44.75746.44943.572
PKS 0516−6212FGLJ0516.8−620705 16 44.5−62 07 04.81.3Mg ii44.44447.48844.523
MG2 J201534+37102FGLJ2015.6+370920 15 28.6+37 10 59.80.859Hb44.34247.97144.794
TXS 2206+6502FGLJ2206.6+650022 08 03.3+65 19 38.71.121Mg ii44.33647.32444.360
Table 1.

Sources from the S13 BL Lac sample, divided according to their reclassification (see Section 4). Column 1: name; column 2: Fermi/LAT counterpart; column 3: right ascension; column 4: declination; column 5: redshift; column 6: line measured, from which LBLR has been derived; column 7: logarithm of the BLR luminosity (in erg s−1); column 8: logarithm of the γ-ray luminosity from Fermi data (in erg s−1); column 9: logarithm of the radio luminosity, calculated at 8 GHz rest frame (in erg s−1).

NameFermi nameRADec.zLinelog LBLRlog Lγlog Lradio
[1][2][3][4][5][6][7][8][9]
BL Lac
GB6 J0013+19102FGLJ0013.8+190700 13 56.3+19 10 41.50.477Mg ii42.69145.44143.043
PKS 0829+0462FGLJ0831.9+042908 31 48.7+04 29 38.20.174Ha42.61445.52042.923
RBS 09582FGLJ1117.2+201311 17 06.1+20 14 07.60.138Ha41.72244.72341.622
PMN J1125-35562FGLJ1125.6−355911 25 31.3−35 57 05.00.284Ha43.33845.15742.627
SBS 1200+6082FGLJ1203.2+603012 03 03.4+60 31 19.10.065Ha42.00543.76941.190
W Comae2FGLJ1221.4+281412 21 31.6+28 13 58.10.103Ha42.13745.05542.199
PG 1218+3042FGLJ1221.3+301012 21 21.9+30 10 36.20.184Ha42.06345.17441.734
OQ 5302FGLJ1420.2+542214 19 46.5+54 23 15.00.153Ha42.20144.76642.603
RGB J1534+3722FGLJ1535.4+372015 34 47.2+37 15 53.80.144Ha41.72244.34940.991
BL/FS
PKS 0332−4032FGLJ0334.2−400803 34 13.4−40 08 26.91.357Mg ii45.04647.71144.967
TXS 0431−2032FGLJ0434.1−201404 34 07.9−20 15 17.20.928Mg ii43.15346.51143.758
PKS 0437−4542FGLJ0438.8−452104 39 00.7−45 22 23.92.017iv45.20147.66145.363
PKS 0627−1992FGLJ0629.3−200106 29 23.7−19 59 19.71.724iv44.06047.74045.036
4C +14.602FGLJ1540.4+143815 40 49.5+14 47 46.50.606Mg ii43.57545.96644.229
PMN J1754−64232FGLJ1755.5−642317 54 41.8−64 23 44.71.255Mg ii44.16346.91244.113
4C +56.272FGLJ1824.0+565018 24 07.0+56 51 01.10.664Mg ii43.91246.89144.333
S3 2150+172FGLJ2152.4+173521 52 24.7+17 34 37.90.874Mg ii44.16846.33344.310
PMN J2206−00312FGLJ2206.6−002922 06 43.2−00 31 02.31.053Mg ii43.79846.55743.858
B2 2234+28A2FGLJ2236.4+282822 36 22.3+28 28 58.10.79Mg ii44.64547.07944.412
PKS 2244−0022FGLJ2247.2−000222 47 30.1+00 00 07.00.949Mg ii44.10646.54343.950
PKS 2312−5052FGLJ2315.7−501423 15 44.2−50 18 39.70.811Mg ii43.62846.35743.764
PKS 2351−3092FGLJ2353.5−303423 53 47.3−30 37 48.30.737Mg ii43.62846.06643.893
FS
NVSS J020344+3042382FGLJ0204.0+304502 03 44.1+30 42 38.10.761Mg ii44.75746.44943.572
PKS 0516−6212FGLJ0516.8−620705 16 44.5−62 07 04.81.3Mg ii44.44447.48844.523
MG2 J201534+37102FGLJ2015.6+370920 15 28.6+37 10 59.80.859Hb44.34247.97144.794
TXS 2206+6502FGLJ2206.6+650022 08 03.3+65 19 38.71.121Mg ii44.33647.32444.360
NameFermi nameRADec.zLinelog LBLRlog Lγlog Lradio
[1][2][3][4][5][6][7][8][9]
BL Lac
GB6 J0013+19102FGLJ0013.8+190700 13 56.3+19 10 41.50.477Mg ii42.69145.44143.043
PKS 0829+0462FGLJ0831.9+042908 31 48.7+04 29 38.20.174Ha42.61445.52042.923
RBS 09582FGLJ1117.2+201311 17 06.1+20 14 07.60.138Ha41.72244.72341.622
PMN J1125-35562FGLJ1125.6−355911 25 31.3−35 57 05.00.284Ha43.33845.15742.627
SBS 1200+6082FGLJ1203.2+603012 03 03.4+60 31 19.10.065Ha42.00543.76941.190
W Comae2FGLJ1221.4+281412 21 31.6+28 13 58.10.103Ha42.13745.05542.199
PG 1218+3042FGLJ1221.3+301012 21 21.9+30 10 36.20.184Ha42.06345.17441.734
OQ 5302FGLJ1420.2+542214 19 46.5+54 23 15.00.153Ha42.20144.76642.603
RGB J1534+3722FGLJ1535.4+372015 34 47.2+37 15 53.80.144Ha41.72244.34940.991
BL/FS
PKS 0332−4032FGLJ0334.2−400803 34 13.4−40 08 26.91.357Mg ii45.04647.71144.967
TXS 0431−2032FGLJ0434.1−201404 34 07.9−20 15 17.20.928Mg ii43.15346.51143.758
PKS 0437−4542FGLJ0438.8−452104 39 00.7−45 22 23.92.017iv45.20147.66145.363
PKS 0627−1992FGLJ0629.3−200106 29 23.7−19 59 19.71.724iv44.06047.74045.036
4C +14.602FGLJ1540.4+143815 40 49.5+14 47 46.50.606Mg ii43.57545.96644.229
PMN J1754−64232FGLJ1755.5−642317 54 41.8−64 23 44.71.255Mg ii44.16346.91244.113
4C +56.272FGLJ1824.0+565018 24 07.0+56 51 01.10.664Mg ii43.91246.89144.333
S3 2150+172FGLJ2152.4+173521 52 24.7+17 34 37.90.874Mg ii44.16846.33344.310
PMN J2206−00312FGLJ2206.6−002922 06 43.2−00 31 02.31.053Mg ii43.79846.55743.858
B2 2234+28A2FGLJ2236.4+282822 36 22.3+28 28 58.10.79Mg ii44.64547.07944.412
PKS 2244−0022FGLJ2247.2−000222 47 30.1+00 00 07.00.949Mg ii44.10646.54343.950
PKS 2312−5052FGLJ2315.7−501423 15 44.2−50 18 39.70.811Mg ii43.62846.35743.764
PKS 2351−3092FGLJ2353.5−303423 53 47.3−30 37 48.30.737Mg ii43.62846.06643.893
FS
NVSS J020344+3042382FGLJ0204.0+304502 03 44.1+30 42 38.10.761Mg ii44.75746.44943.572
PKS 0516−6212FGLJ0516.8−620705 16 44.5−62 07 04.81.3Mg ii44.44447.48844.523
MG2 J201534+37102FGLJ2015.6+370920 15 28.6+37 10 59.80.859Hb44.34247.97144.794
TXS 2206+6502FGLJ2206.6+650022 08 03.3+65 19 38.71.121Mg ii44.33647.32444.360

We add to these new samples the objects that show broad emission lines studied in TS12. Specifically, we add the 45 FSRQs and 1 BL Lac (following our reclassification) that were included in Shen et al. (2011), along with the 15 BL Lacs with broad emission lines from G11. In this work, we only consider the objects from TS12 that have broad emission lines detected, excluding then all the BL Lacs with only an upper limit on the BLR luminosity. In TS12, the upper limits were introduced to increase the number of BL Lacs, populating the low-luminosity branch of our sample. They followed the LBLRLγ correlation already found only with the detections, therefore they were not very constraining. In our new work, we increased the number of BL Lacs with broad emission lines thanks to S13, and therefore we do not need the lineless BL Lacs.

In total, we have 225 FSRQs and 42 BL Lacs with detected broad emission lines, both γ-ray and radio counterparts, and a reliable estimate of the central black hole masses (MBH).

The radio-galaxy sample

We collected a sample of radio galaxies without broad emission lines from the work by Buttiglione et al. (2010). The authors studied the optical spectroscopical and radio features of all the z < 0.3 radio galaxies (Buttiglione et al. 2009) with F178 MHz > 9 Jy, δ > −5° and an optical counterpart from the Third Cambridge Radio Catalogue (Spinrad et al. 1985). The authors classify the sources in HEG, LEGs and broad line objects (BLOs) according to optical features. Specifically, BLOs clearly show broad emission lines, while HEG and LEGs do not show any broad emission feature, but while the latter have an intrinsic lack of broad line emitting structures, the former show high-excitation fingerprints, suggesting obscuration of the BLR more than a true absence. The introduction of radio galaxies in our work aims at studying the true lineless jetted AGN, so we include in our sample the 37 LEGs studied by Buttiglione et al. (2010).

Along with the optical spectral analysis, they performed a radio morphology study. The radio analysis provides information on the core power, which allows us to have a reliable tracer of the inner jet power, without contamination from the extended structures. In other words, the core power traces the same emission that is traced by the radio luminosity in the case of blazars, apart from the different beaming factor (due to the different orientation angles of the jet, see Section 5). Moreover, the authors provide a measure of the H magnitude, and a calibration to obtain from that the central black hole mass of the objects (Marconi & Hunt 2003):
(2)
The authors find different radio morphologies among the 37 LEGs: 12 FR I, 16 FR II and 9 unclassifiable sources. One object out of each radio group has no MH information, so we cannot derive a black hole mass estimate. We exclude from our sample those objects.

We are then left with 11 LEGs FR I, 15 LEGs FR II and 8 LEGs without a radio classification (FR?), all with an estimate of the radio core power, with narrow emission line information that gives upper limits on the broad emission lines (see Section 3), and black hole mass estimates. We also added to the radio-galaxy sample M87, a ‘classic’ LEG FR I, taking the radio flux from NASA/IPAC Extragalactic Database (NED), while the optical spectroscopic information is taken from Buttiglione et al. (2009). M87 does not show any broad emission line, so in the following we will treat it as the LEGs studied by Buttiglione et al. (2010).

THE BLR LUMINOSITY

To calculate the BLR luminosity, we need an estimate of the broad emission lines of each object, specifically of Lyα, Hα, Hβ, Mg ii or C iv. In the case of the S12 FSRQs, we take directly the emission line fluxes listed in S12, as we did in TS12, while in the case of BL Lacs we directly fit the few broad emission lines present. G11 had already derived from literature the broad emission line values for the BL Lacs included in their sample. Note that the emission lines visible in Mkn421 and Mkn501 spectra have a full width at half-maximum (∼1000 km s−1) that lies on the classification threshold between narrow lines and broad lines. Therefore, we rather put an upper limit on their broad line luminosities, to be more conservative. We nevertheless leave them in our sample, since they are usually listed as known BL Lacs with lines. The objects included in our radio-galaxy sample have the same issue, being explicitly selected as sources intrinsically lacking broad emission lines. They only show narrow emission lines, so we can obtain from them upper limits on broad emission line luminosities, taking as example quasars with both broad and narrow emission lines. Narrow-line-dominated (but with broad emission lines) quasars are in fact included in S11. In their case, the ratio between narrow and broad Hα luminosities is
(3)
In other words, if present, the broad emission lines have a luminosity that is at least 10 per cent the corresponding narrow line luminosity, or more. To be conservative, we choose to fix as a robust upper limit on the broad line luminosity, the luminosity associated with the corresponding detected narrow emission line.

From the broad emission line luminosities or from the upper limits collected for our sample, we can then derive the overall luminosity emitted from the BLR. We follow Celotti, Padovani & Ghisellini (1997) and set the Lyα flux contribution to 100, and the relative weights of the Hα, Hβ, Mg ii and C iv lines to 77, 22, 34 and 63, respectively (see Francis et al. 1991). The total broad line flux is fixed at 555.76. The LBLR value of each source has been derived using these proportions. When more than one line is present, we calculate the logarithmic average of the LBLR estimated from each line.

LBLR as a tracer of the accretion

The BLR luminosity is a reliable tracer of the emission from the accretion structure, since the plasma producing the broad emission lines is directly ionized by its radiation. The fraction of disc luminosity that ionizes the plasma in the BLR (photoionizing luminosity, Lion) depends on the geometry of the disc, and hence on its radiation efficiency. In the simplest hypothesis for a radiatively efficient accretion disc (e.g. Shakura & Sunyaev 1973), the disc is geometrically thin and emits as a blackbody at all radii. For simplicity, we approximate the photoionizing luminosity with the entire disc luminosity (⁠|$L_{\rm d}=\eta \dot{M}c^2$|⁠). Therefore, we can assume that Lion follows the same dependence on the accretion rate as the disc luminosity:
(4)
In this case, then, assuming an average covering factor of ∼10 per cent for the BLR, the disc luminosity is Ld ∼ 10LBLR (e.g. Baldwin & Netzer 1978; Smith et al. 1981). Following the Shakura & Sunyaev hypothesis, such a radiatively efficient disc should occur for |$\dot{m}=\dot{M}/\dot{M}_{\rm Edd}>\dot{m_{\rm c}}$|⁠, where |$\dot{M}_{\rm Edd}=L_{\rm Edd}/c^2$| and |$\dot{m}_{\rm c}\sim 0.1$| (Narayan & Yi 1995). Sharma et al. (2007) suggested instead that the disc remains radiatively efficient for |$\dot{m}_{\rm c}\gtrsim 10^{-4}$|⁠.
For values |$\dot{m}<\dot{m}_{\rm c}$| the accretion structure should become radiatively inefficient, switching from a proper accretion disc to a hot accretion flow. In this case, it does not emit as a blackbody, and the overall luminosity decreases and changes its dependence on the accretion rate to |$L_{\rm d}\propto \dot{M}^2$| (Narayan, Garcia & McClintock 1997). As a consequence of the change in the emission mode, the ionizing luminosity is no more a relevant fraction of the overall disc luminosity, but Lion ≪ Ld. The accretion flow cannot emit as a blackbody, because it is no longer optically thick, and the emission is dominated by synchrotron, bremsstrahlung and Comptonization processes (see Yuan & Narayan 2014, for a review). The SED of the accretion flow is therefore very different compared to a Shakura–Sunyaev-like accretion disc (see fig. 1 in Mahadevan 1997). The photoionizing luminosity (mostly the luminosity emitted in the UV wavelength range) is no more a fixed fraction of the disc luminosity. Since this fraction decreases when the accretion rate decreases, the photoionizing luminosity scales with the accretion rate following a relation steeper than |$\propto \dot{M}^2$|⁠. Following the model in Mahadevan (1997), in TS12 we estimated the ionizing luminosity for the different accretion rates considered by the author. The UV wavelength range is the most affected from the change in accretion rate of the whole SED, and in fact we found a very steep relation:
(5)
This strong change would lead to an analogous change in the dependence of LBLR on |$\dot{M}$|⁠.

THE LBLRLγ RELATION

Table 2.

Results of the partial correlation analysis of the LBLRLγ and LBLR/LEddLγ/LEdd relations using a least-squares fit. The whole blazar sample is taken into account, i.e. 267 objects are considered in the analysis. Correlations are of the form y = mx + q. The listed slopes m refer to the bisector (of the two correlations x versus y and y versus x). r is the correlation coefficient obtained from the analysis. We list also the results when accounting for the common dependence on redshift and/or black hole mass. In all the cases, the probability that the correlation is random is P < 4 × 10−8, i.e. all the correlations are statistically relevant.

mqr
x = log Lγ; y = log LBLR
(x, y)0.92 ± 0.191.2 ± 14.80.81
(x, y), z0.92 ± 0.191.2 ± 14.80.59
x = log (Lγ/LEdd); y = log (LBLR/LEdd)
(x, y)0.84 ± 0.20−2.46 ± 3.10.78
(x, y), z0.84 ± 0.20−2.46 ± 3.10.65
(x, y), z, M0.84 ± 0.20−2.46 ± 3.10.64
mqr
x = log Lγ; y = log LBLR
(x, y)0.92 ± 0.191.2 ± 14.80.81
(x, y), z0.92 ± 0.191.2 ± 14.80.59
x = log (Lγ/LEdd); y = log (LBLR/LEdd)
(x, y)0.84 ± 0.20−2.46 ± 3.10.78
(x, y), z0.84 ± 0.20−2.46 ± 3.10.65
(x, y), z, M0.84 ± 0.20−2.46 ± 3.10.64
Table 2.

Results of the partial correlation analysis of the LBLRLγ and LBLR/LEddLγ/LEdd relations using a least-squares fit. The whole blazar sample is taken into account, i.e. 267 objects are considered in the analysis. Correlations are of the form y = mx + q. The listed slopes m refer to the bisector (of the two correlations x versus y and y versus x). r is the correlation coefficient obtained from the analysis. We list also the results when accounting for the common dependence on redshift and/or black hole mass. In all the cases, the probability that the correlation is random is P < 4 × 10−8, i.e. all the correlations are statistically relevant.

mqr
x = log Lγ; y = log LBLR
(x, y)0.92 ± 0.191.2 ± 14.80.81
(x, y), z0.92 ± 0.191.2 ± 14.80.59
x = log (Lγ/LEdd); y = log (LBLR/LEdd)
(x, y)0.84 ± 0.20−2.46 ± 3.10.78
(x, y), z0.84 ± 0.20−2.46 ± 3.10.65
(x, y), z, M0.84 ± 0.20−2.46 ± 3.10.64
mqr
x = log Lγ; y = log LBLR
(x, y)0.92 ± 0.191.2 ± 14.80.81
(x, y), z0.92 ± 0.191.2 ± 14.80.59
x = log (Lγ/LEdd); y = log (LBLR/LEdd)
(x, y)0.84 ± 0.20−2.46 ± 3.10.78
(x, y), z0.84 ± 0.20−2.46 ± 3.10.65
(x, y), z, M0.84 ± 0.20−2.46 ± 3.10.64

Fig. 1 represents the first result of our work. The left-hand panel shows the BLR luminosity as a function of the γ-ray luminosity. The right-hand panel shows the same quantities divided by the Eddington luminosity (LEdd). The objects included in TS12 are marked as FSRQs or BL Lacs, according to how we classified them in our previous work. The new FSRQs from S12 are included as FSRQs, while the objects classified as BL Lacs in S13 are marked as BL Lacs, BL/FS or FS in the plots. Note that the correlation we found in TS12 is confirmed by the new data, both when directly comparing the two luminosities and when normalizing them by the Eddington luminosity. This clearly strengthens the hypothesis of a tight relation between the accretion rate and the jet power in blazars. As explained in Section 3.1, the LBLR is a very good tracer of the accretion rate, while the γ-ray luminosity traces well the jet power. We calculate the best fit of the relation between the two luminosities, both normalized by the Eddington luminosity and not. We find that both are consistent with the results found in TS12. We apply a partial correlation analysis, to take also into account the possible common dependence on z and MBH of the values. The LBLRLγ and LBLR/LEddLγ/LEdd relations results are linear and statistically relevant (see Table 2).

BLR luminosity as a function of γ-ray luminosity (left-hand panel) and the same, normalized at the Eddington luminosity (right-hand panel). Different symbols correspond to different classifications, as labelled. The dashed lines are the results of the least-squares fit described in Table 2. The grey stripe in the right-hand panel indicates the luminosity divide between FSRQs and BL Lacs, located at LBLR/LEdd ∼ 5 × 10−4.
Figure 1.

BLR luminosity as a function of γ-ray luminosity (left-hand panel) and the same, normalized at the Eddington luminosity (right-hand panel). Different symbols correspond to different classifications, as labelled. The dashed lines are the results of the least-squares fit described in Table 2. The grey stripe in the right-hand panel indicates the luminosity divide between FSRQs and BL Lacs, located at LBLR/LEdd ∼ 5 × 10−4.

Contrary to our previous work, instead, the apparent ‘divide’ between FSRQs and BL Lacs (located at LBLR/LEdd ∼ 5 × 10−4) seems no longer valid, since some BL Lacs from S13 are located in the high-luminosity branch of the correlation, in the area typically occupied by FSRQs (S13 BL Lacs are marked as purple and orange asterisks, or included among the blue open squares in all the figures). To understand this discrepancy with our previous results, we first inspected visually the overall SEDs of the BL Lacs from S13. We notice that the sources show three different SED behaviours (as shown by the individual SEDs in the appendix).

  • Nine have the synchrotron emission dominant or comparable to the high-energy component, and the thermal emission from the accretion structure completely swamped by the non-thermal jet emission. These features define a BL Lac, according to the classification scheme adopted in G11 and TS12, and first introduced by Padovani & Giommi (1995).

  • Four of them show a clear Compton dominance, and the emission from the accretion disc is clearly visible. We then classify them as FSRQs, and claim for a misclassification in S13. However, we highlight them differently and label them as ‘FS’, to keep track of them in the plots.

  • 13 objects have the high-energy component that slightly dominates the synchrotron emission, as in the case of non-extreme FSRQs. On the other hand, the synchrotron component completely swamps the accretion emission, leading to a BL Lac-like optical appearance. We classify them as ‘BL/FS’, since they show both an FSRQ and a BL Lac fingerprint.

The objects classified as FS and FS/BL are labelled accordingly in all our figures. From Fig. 1, we immediately notice that all these ‘reclassified’ BL Lacs are the S13 BL Lacs that occupy the high-luminosity branch of our correlations. The FS have all the typical FSRQ features, so we expect to find them in the high-luminosity branch of the LBLRLγ correlations (see Fig. 1). Interestingly, all the others objects from S13 that were located in the FSRQ branch are the 13 that we classified as BL/FS. Their location allows us to better understand their peculiar SED features. We can in fact infer that they have an intrinsic powerful jet and a highly luminous accretion disc (i.e. high accretion rate), as common FSRQs, even if their optical spectroscopical features are BL Lac-like. In other words, the BL/FS were classified as BL Lacs because of an unusually powerful synchrotron emission that reduced the EW of their broad emission lines, but are instead FSRQs. Even a very luminous thermal continuum, with the related emission features, can in fact be overcome by a very luminous non-thermal continuum (Giommi et al. 2012, 2013). Since the synchrotron emission is mainly driven by the energy density of the magnetic field, we can expect that these objects have it unusually high. This is confirmed by the SED modelling. We fitted the overall SEDs with a one-zone leptonic model (Ghisellini & Tavecchio 2009), and the results show that the energy density of the magnetic field of all the BL/FS is unusually high (detailed results will be shown in Ghisellini & Tavecchio, in preparation).

Considering the reclassification, the division between FSRQs and BL Lacs at LBLR/LEdd ∼ 5 × 10−4 becomes even more relevant. The BL/FS are actually FSRQs ‘disguised’ as BL Lacs and the canonical classification based on the EW of their broad emission lines fails in classifying them. The division based on LBLR/LEdd represents a more physical classifying system, since it discriminates the objects in terms of their accretion rate. However, the divide is not sharp, and again our blazar sample seems to be distributed continuously in both the LBLRLγ and the LBLR/LEddLγ/LEdd planes.

Along with the divide, we are interested in studying at what accretion rate (and if) a change in the accretion structure happens. As detailed in Section 3.1, a standard Shakura–Sunyaev disc should occur for accretion rates larger than a critical value |$\dot{m_{\rm c}}$|⁠. Below that value, the accretion structure is no longer radiatively efficient, also ionizing less efficiently the plasma in the BLR. This change in accretion should hence be reflected in a change of slope in the LBLR/LEddLγ/LEdd plot. The jet power is in fact directly correlated to the accretion rate at all values of the accretion rate itself (Celotti & Ghisellini 2008; Ghisellini et al. 2010). In TS12, the low-luminosity branch of the LBLR/LEddLγ/LEdd plot was not populated enough to draw a firm conclusion. The new BL Lacs have increased the number of objects that could help in understanding the possible existence of a break in the relation, but the data are still too sparse to draw a firm conclusion. Hence, we try to have a new perspective on the problem, by introducing another tracer for the jet power, that allows us to reach smaller accretion rates and observe directly the behaviour of the jet–disc system in the case of truly inefficient accretion structures, i.e. objects intrinsically without broad emission lines.

THE LBLRLradio RELATION

We aim to introduce in our study objects that do not have broad emission lines, but with reliable estimates of z and MBH, and a direct proxy for the jet power. We also want to be able to derive an upper limit on their BLR luminosity, which we will use as a proxy for Lion. As we saw from S13 BL/FS objects, the non-thermal continuum emitted from the jet, highly boosted because of relativistic effects, can dilute dramatically even strong broad emission lines. In the case of less luminous lines, this problem is obviously even bigger. In fact, the majority of γ-ray detected BL Lacs lack a reliable redshift estimate, since their optical spectra are completely dominated by the non-thermal emission, and they do not show any emission features. This means that we cannot discriminate whether an object is genuinely lineless or its faint emission lines are simply not visible. Hence, to select only truly lineless object, we choose to introduce in our study a sample of radio galaxies, i.e. jetted AGN in which the optical emission is not completely dominated by the non-thermal, boosted jet emission (sample description in Section 2.2). We choose a group of LEGs, to be sure that their broad emission lines are not present, likely because of a radiatively inefficient accretion disc. Radio galaxies are usually not γ-ray detected, so we cannot use the γ-ray luminosity as a tracer of the jet power. We then consider radio luminosity at 8 GHz rest frame as an alternative jet tracer, with the following caveat: the radio luminosity is emitted from the jet, and is therefore beamed in the emission direction. We will take into account the different beaming factors that characterize blazars and radio galaxies in the discussion.

Fig. 2 shows the comparison between BLR and radio luminosities in all the sources of our samples. All the radio galaxies have upper limits on their LBLR values, since they are explicitly selected to be lineless (see Section 3 for the upper limits derivation). Note that the radio luminosities calculated for blazars and radio galaxies (and plotted in Fig. 2) are differently beamed, because of different viewing angles. Therefore, the linear correlation over the whole luminosity range is only apparent. To properly study the LradioLBLR relation, we have to homogenize the beaming factors. This is true also if we consider the two luminosities normalized to the Eddington luminosity, as shown in Fig. 3, which we analyse in detail in Section 6.

BLR luminosity as a function of radio luminosity. Different symbols correspond to different classification of the objects. The dashed line is the result of a least-squares fit calculated among the detection, i.e. the blazars.
Figure 2.

BLR luminosity as a function of radio luminosity. Different symbols correspond to different classification of the objects. The dashed line is the result of a least-squares fit calculated among the detection, i.e. the blazars.

Luminosity of the BLR (in Eddington units) for the sources from our samples as a function of the radio luminosity (in Eddington units). Different symbols correspond to different samples or a different classification of the sources, as labelled. The dashed lines indicate the bisector, rescaled to pass through the FSRQs. The dark grey horizontal stripe indicates the luminosity divide between FSRQs and BL Lacs at LBLR/LEdd ∼ 5 × 10−4. The light grey stripe indicates the expected distribution of the luminosities if they were produced by a Shakura–Sunyaev accretion disc for LBLR/LEdd ∼ 5 × 10−4 and an ADAF with a Mahadevan-like spectrum ($L_{\rm d}\propto \dot{m}^{3.5}$, see Section 3.1 and Section 6). The leftmost ellipse include the core of the radio galaxies of our sample. The central and rightmost ellipses show where the radio galaxies would be located if they were beamed according to Lorentz factors Γ = 3 or 10, respectively (as indicated by the arrows, and described in Section 6).
Figure 3.

Luminosity of the BLR (in Eddington units) for the sources from our samples as a function of the radio luminosity (in Eddington units). Different symbols correspond to different samples or a different classification of the sources, as labelled. The dashed lines indicate the bisector, rescaled to pass through the FSRQs. The dark grey horizontal stripe indicates the luminosity divide between FSRQs and BL Lacs at LBLR/LEdd ∼ 5 × 10−4. The light grey stripe indicates the expected distribution of the luminosities if they were produced by a Shakura–Sunyaev accretion disc for LBLR/LEdd ∼ 5 × 10−4 and an ADAF with a Mahadevan-like spectrum (⁠|$L_{\rm d}\propto \dot{m}^{3.5}$|⁠, see Section 3.1 and Section 6). The leftmost ellipse include the core of the radio galaxies of our sample. The central and rightmost ellipses show where the radio galaxies would be located if they were beamed according to Lorentz factors Γ = 3 or 10, respectively (as indicated by the arrows, and described in Section 6).

Note that the objects reclassified as BL/FS are located at the highest radio luminosity edge of the correlation in both Figs 2 and 3. This clearly highlights their FSRQ nature, associated with an uncommonly luminous synchrotron emission, very well traced by the radio luminosity itself. They can easily be considered as the tail at high magnetic field energy density of the class of FSRQs.

Note that in both plots the radio galaxies are located at LBLR (normalized by LEdd and not) lower than the BL Lacs, with a small overlap. By including these objects, we finally manage to achieve very low disc luminosities, and hence very low accretion rates, which is crucial for our understanding. We discuss the implications of this in Section 6. The upper limits hint also at a change in relationship between the tracers of Lion and jet power but their presence in the low-luminosity branch of the sample prevents us from a proper parametric characterization of this relation. To understand if the relationship between LBLR/LEdd and Lradio/LEdd is better explained by a single or a broken power law, we do the following: we treat the upper limits as detections and, since these cluster mostly at the low power end, the slope of the power law we get from the fit has to be considered as a lower limit (i.e. the true value will be steeper). With this approach, we can compare the two hypotheses of a single and a broken power law by using an F-test. By minimizing the χ2 values in the hypotheses of a single power-law (log LBLR/LEdd = (0.87 ± 0.12)log Lradio/LEdd + 0.28) and a broken power law with break at Lradio/LEdd ∼ −4, we infer that the data are better described by a broken power law with a 99.97 per cent level of confidence.1 The break value is fixed to correspond to the dividing value between FSRQs and BL Lacs in LBLR/LEdd discussed in Section 4, following the single power law. The broken power law that better fits our data is described by the relation
(6)
This shows that a single power law is not a good representation of our data. We discuss in the next section the meaning of this result.

DISCUSSION

Fig. 3 is the main result of our work. As we have already pointed out, the radio luminosity has a different physical meaning in the case of blazars and radio galaxies, because of different beaming levels. To properly compare them, we have to beam the radio luminosity of the radio galaxies, assuming an average bulk Lorentz factor Γ and a viewing angle θ. This will shift them at higher radio luminosities, rejoining them with their aligned analogous AGN. Note that the different orientation of blazars and their parent population does not affect the BLR luminosity, since it is emitted isotropically. Therefore, the parent population of a group of blazars would be located at the same LBLR/LEdd, with a Lradio/LEdd smaller than the corresponding aligned blazars. From their position in Figs 2 and 3, the LEG FR I radio galaxies are not the parent population of the BL Lacs included in our study (and see Chiaberge et al. 2000), In fact, it is important to remember that the BL Lacs in our sample have broad emission lines, while the FR I we collected are intrinsically without broad emission lines. This spectral difference explains why our FR I and our BL Lacs are intrinsically different. There is likely a population of truly lineless BL Lacs, of which these LEGs are actually the parent population, that we are not able to include in our study. The only upper limits on LBLR that we derived for a group of BL Lacs in TS12 (from Plotkin et al. 2011) were anyway located in the same LBLR range as the broad line BL Lacs. None of the known BL Lacs with a measured redshift represent the re-oriented analogues of the LEG FR I.

However, we are considering only the tip of the iceberg of the BL Lac population: 2LAC includes 475 BL Lacs, and most of them do not have a reliable redshift estimate, since their optical spectra are completely featureless. Without a redshift estimate, we cannot derive their intrinsic luminosity in any band, nor calculate an upper limit on their broad line luminosity. Therefore, they cannot be compared to the other blazars in our work. Among them, there are most likely the truly lineless BL Lacs that would be necessary to study the very low accretion regimes, and of course they would be the aligned analogues to the LEGs that we include in our study. This makes the radio galaxies without broad emission lines even more relevant for our work, since they are the only valid tracer of the low-accretion regime. But to use them to explore that regime, we have to uniform their beaming to the blazar one.

How much do we have to beam the radio luminosity of the LEGs in our sample to compare them with blazars? The beaming factor of a source with generic Γ, β and θ is
(7)
We take an average viewing angle θLEG ∼ 40°, assuming that the LEGs are misaligned. We want to beam their radio luminosity as they were oriented as a blazar, i.e. with θBL ∼ 3°. Therefore, we have to boost the radio luminosity by a factor
(8)
If we assume their jets are beamed with a Lorentz factor similar to a common blazar, i.e. Γ ≃ 10, this beaming factor would be ∼5 × 104, shifting the LEGs as drawn in Fig. 3 (rightmost ellipse). None of the existing blazars could populate that region of the plot, even considering the BL Lacs without redshift. In fact, if we assign an arbitrary z = 1 to all BL Lacs without a redshift estimate in 2LAC, we obtain an average radio luminosity at 8 GHz rest frame, log Lradio ∼ 43.2, which would correspond to log (Lradio/LEdd) ∼ −3.3. This would not be enough to populate the region occupied by the rightmost ellipse, i.e. the radio galaxies beamed of a factor ∼5 × 104 due to Γ ≃ 10. To populate that region, one should postulate the existence of AGN with an extremely powerful and relativistic jet, associated with an accretion structure extremely radiatively inefficient, or even without any accretion structure present. This beaming level seems then quite unlikely.

We can consider another beaming option. From VLBI (very long baseline interferometry) studies, there is evidence that in strong TeV BL Lacs the pc-scale jets move slowly (Edwards & Piner 2002; Piner & Edwards 2004). At the same time, the extreme variability of their intense TeV luminosity implies that the jet should be highly relativistic, at least in the region where the TeV emission originates. To justify such a discrepancy, the two less demanding hypothesis that have been advanced are: (i) a deceleration of the emitting region between the TeV and the radio locii (Georganopoulos & Kazanas 2004); and (ii) a spine-layer structure of the jet (Ghisellini, Tavecchio & Chiaberge 2005). Furthermore, detailed observations performed with the VLBI show a morphology that suggests the presence of a slower external layer, surrounding a faster core in the jet in the lineless BL Lac Mkn 501 (Giroletti et al. 2004). Similar results have also been obtained in the case of some radio galaxies (Owen, Hardee & Cornwell 1989; Swain, Bridle & Baum 1998; Giovannini et al. 1999). Moreover, a velocity structure helps in explaining other features typical of radio galaxies, such as the configuration of their magnetic field (Komissarov 1990; Laing 1993). According to this hypothesis, the radio emission should then be characterized by a rather small Lorentz factor, Γ ∼ 3, being emitted by the external layer. In this case, the LEG radio luminosity can be boosted by a smaller factor (δBLLEG)3 ∼ 100.

The central ellipse in Fig. 3 represents this hypothesis. Such a beaming factor shifts the radio power of the LEGs by such an amount that in the LBLR/LEddLradio/LEdd plane they now follow a much steeper relationship than the almost linear best fit derived in Section 5 with the broken power law. As previously done, we perform an F-test to compare the single and a broken power-law hypotheses, obtained as best fits of our data, all assumed as detections. The two best fits are log LBLR/LEdd = (0.89 ± 0.19)log Lradio/LEdd + 0.36 and again a broken power law with break at Lradio/LEdd ∼ −4:
(9)
The F-test shows that, again, a broken power law provides a better description of the data at the >99.99 per cent level. We stress again that the slope derived with this method at values below the break provides only a lower limit to the true slope. In other words, at luminosities smaller than the break, jetted AGN follow a relation steeper than
(10)
This clearly means that the jetted AGN highlight the break that we expect from standard theory of accretion discs, i.e. there is a change in the accretion process at LBLR/LEdd ≈ 10−3.5 (the light grey stripe in Fig. 3). In fact, if we interpret the LEGs as the jetted AGN with lowest accretion rate, as suggested by their LBLR, they would follow a relation that is steeper than the one expected in the case of an efficient accretion structure (i.e. linear). This relation is even steeper than the simple dependence of Ld on the accretion rate (⁠|$L_{\rm d}\propto \dot{M}^2$|⁠), since the slope we derive is only a lower limit on the true slope. This result is consistent with highly radiatively inefficient models, closer to the most inefficient one, i.e. a pure advection-dominated accretion flow (ADAF), from which we expect the relation
(11)
below the value of |$\dot{m_{\rm c}}$|⁠, which separates the two accretion regimes.

We find that such a transition occurs at |$\dot{m}_{\rm c}\sim 0.1$|⁠, i.e. LBLR/LEdd ∼ 5 × 10−4–10−3 if a radiative efficiency η ∼ 0.1 is assumed. This threshold is consistent also with the accretion rate transition between FR I and FR II found by Ghisellini & Celotti (2001). The hypothesis of a transition at |$\dot{m}_{\rm c}\sim 10^{-4}$| would not be consistent with the beamed LEG data. In any case, we do not expect a sharp transition, but more likely a smooth one, since we do not observe a clear bimodality in the LBLR/LEdd distribution.

CONCLUSIONS

In this work, we have explored the connection between jet and accretion structure in jetted AGN, using 267 broad emission line blazars and 38 broad-line-less radio galaxies, all with known redshift, a measure of the jet power and an estimate of the black hole mass. In the case of blazars, we have used both γ-ray and radio luminosities to trace their jet power, while the radio galaxies only have the radio core power as a jet tracer. Since they do not show broad emission lines, we have derived robust upper limits on their BLR luminosity from the luminosity of their narrow lines. They are crucial to explore the low-accretion regime of jetted AGN. The results we obtained can be summarized as follows.

  • With a sample composed by both blazars and radio galaxies, we finally can identify the transition between efficient and inefficient accretion structures. With only blazars, we are not able to include the very low accreting objects, since they would be lineless and dominated by the jet non-thermal emission, and therefore again without a redshift estimate. LEG radio galaxies are therefore the only means to study the radiatively inefficient accretion regime.

  • The most reasonable beaming option for the radio galaxies we included is due to jets structured with a central extremely relativistic spine, surrounded by a slower layer. A high Lorentz factor Γ = 10, necessary to justify some observational properties, would characterize only the central part of the jet. A slower layer likely surrounds this extreme spine, and would be the responsible for the radio emission from the jet. This external layer is characterized by a smaller Lorentz factor (Γ ∼ 3), implying a smaller beaming factor to homogenize radio galaxies to blazars.

  • The transition between efficient and inefficient accretion regimes occurs at the standard critical value |$\dot{m}_{\rm c}\sim 0.1$|⁠, i.e. at LBLR/LEdd ∼ 5 × 10−4–10−3 assuming an accretion efficiency η ∼ 0.1. At accretion values lower than that, the ionizing luminosity decreases with a slope steeper than |$\propto \dot{m}^{2}$|⁠, clearly traced by the radio galaxies. This is consistent with a transition from an efficient to an inefficient regime at low accretion rates. A relevant decrease in the ionizing luminosity is in fact expected in all the highly inefficient accretion regimes (e.g. the ADAF model).

We thank the referee for useful comments that improved the paper. Part of this work is based on archival data, software or online services provided by ASDC. This research also made use of NED which is operated by the Jet Propulsion Laboratory, Caltech, under contract with NASA.

1

The uncertainties on log LBLR/LEdd, necessary to calculate the χ2, are derived from the uncertainties on the broad line luminosities and on the black hole mass measurements, and are typically ∼0.3 dex.

Associated to INAF – Osservatorio Astronomico di Roma, via Frascati 33, I-00040 Monteporzio Catone, Italy.

REFERENCES

Abdo
A. A.
et al. 
ApJ
2010
, vol. 
715
 pg. 
429
 
Ackermann
M.
et al. 
ApJ
2011
, vol. 
743
 pg. 
171
 
Baldwin
J. A.
Netzer
H.
ApJ
1978
, vol. 
226
 pg. 
1
 
Baum
S. A.
Heckman
T.
ApJ
1989a
, vol. 
336
 pg. 
681
 
Baum
S. A.
Heckman
T.
ApJ
1989b
, vol. 
336
 pg. 
702
 
Bettoni
D.
Falomo
R.
Fasano
G.
Govoni
F.
A&A
2003
, vol. 
399
 pg. 
869
 
Buttiglione
S.
Capetti
A.
Celotti
A.
Axon
D. J.
Chiaberge
M.
Macchetto
F. D.
Sparks
W. B.
A&A
2009
, vol. 
495
 pg. 
1033
 
Buttiglione
S.
Capetti
A.
Celotti
A.
Axon
D. J.
Chiaberge
M.
Macchetto
F. D.
Sparks
W. B.
A&A
2010
, vol. 
509
 pg. 
6
 
Capetti
A.
Raiteri
C. M.
Buttiglione
S.
A&A
2010
, vol. 
349
 pg. 
77
 
Celotti
A.
Ghisellini
G.
MNRAS
2008
, vol. 
385
 pg. 
283
 
Celotti
A.
Padovani
P.
Ghisellini
G.
MNRAS
1997
, vol. 
286
 pg. 
415
 
Chiaberge
M.
Celotti
A.
Capetti
A.
Ghisellini
G.
A&A
2000
, vol. 
358
 pg. 
104
 
Edwards
P. G.
Piner
B. G.
ApJ
2002
, vol. 
579
 pg. 
67
 
Fanaroff
B. L.
Riley
J. M.
MNRAS
1974
, vol. 
167
 pg. 
31
 
Francis
P. J.
Hewett
P. C.
Foltz
C. B.
Chaffee
F. H.
Weymann
R. J.
Morris
S. L.
ApJ
1991
, vol. 
373
 pg. 
465
 
Georganopoulos
M.
Kazanas
D.
ApJ
2004
, vol. 
604
 pg. 
81
 
Ghisellini
G.
Celotti
A.
A&A
2001
, vol. 
379
 pg. 
L1
 
Ghisellini
G.
Tavecchio
F.
MNRAS
2009
, vol. 
397
 pg. 
985
 
Ghisellini
G.
Tavecchio
F.
Chiaberge
M.
A&A
2005
, vol. 
432
 pg. 
401
 
Ghisellini
G.
Tavecchio
F.
Foschini
L.
Ghirlanda
G.
Maraschi
L.
Celotti
A.
MNRAS
2010
, vol. 
402
 pg. 
497
 
Ghisellini
G.
Tavecchio
F.
Foschini
L
Ghirlanda
G.
MNRAS
2011
, vol. 
414
 pg. 
2674
  
(G11)
Giommi
P.
Padovani
P.
Polenta
G.
Turriziani
S.
D'Elia
V.
Piranomonte
S.
MNRAS
2012
, vol. 
420
 pg. 
2899
 
Giommi
P.
Padovani
P.
Polenta
G.
MNRAS
2013
, vol. 
431
 pg. 
1914
 
Giovannini
G.
Taylor
G. B.
Arbizzani
E.
Bondi
M.
Cotton
W. D.
Feretti
L.
Lara
L.
Venturi
T.
ApJ
1999
, vol. 
522
 pg. 
101
 
Giroletti
M.
et al. 
ApJ
2004
, vol. 
600
 pg. 
127
 
Healey
S. E.
Romani
R. W.
Taylor
G. B.
Sadler
E. M.
Ricci
R.
Murphy
T.
Ulvestad
J. S.
Winn
J. N.
ApJS
2007
, vol. 
171
 pg. 
61
 
Komissarov
S. S.
SvA
1990
, vol. 
16
 pg. 
L284
 
Labiano
A.
A&A
2008
, vol. 
488
 pg. 
L59
 
Laing
R.
Burgarella
D.
Livio
M.
O'Dea
C.
Astrophysical Jets
1993
Cambridge
Cambridge Univ. Press
pg. 
95
 
Laing
R. A.
Jenkins
C. R.
Wall
J. V.
Unger
S. W.
Bicknell
G. V.
Dopita
M. A.
Quinn
P. J.
ASP Conf. Ser. Vol. 54, The First Stromlo Symp. The Physics of Active Galaxies
1994
San Francisco
Astron. Soc. Pac.
pg. 
201
 
Landt
H.
Padovani
P.
Giommi
P.
Perlman
E. S.
MNRAS
2004
, vol. 
351
 pg. 
83
 
Mahadevan
R.
ApJ
1997
, vol. 
447
 pg. 
585
 
Marconi
A.
Hunt
L. K.
ApJ
2003
, vol. 
589
 pg. 
L21
 
Morganti
R.
Tadhunter
C. N.
Dickson
R.
Shaw
M.
A&A
1997
, vol. 
326
 pg. 
130
 
Narayan
R.
Yi
I.
ApJ
1995
, vol. 
452
 pg. 
710
 
Narayan
R.
Garcia
M. R.
McClintock
J. E.
ApJ
1997
, vol. 
478
 pg. 
L79
 
Owen
F. N.
Hardee
P. E.
Cornwell
T. J.
ApJ
1989
, vol. 
340
 pg. 
698
 
Padovani
P.
Giommi
P.
ApJ
1995
, vol. 
444
 pg. 
567
 
Piner
B. G.
Edwards
P. G.
ApJ
2004
, vol. 
600
 pg. 
115
 
Plotkin
R. M.
Markoff
S.
Trager
S. C.
Anderson
S. F.
MNRAS
2011
, vol. 
413
 pg. 
805
 
Rawlings
S.
Saunders
R.
Nature
1991
, vol. 
349
 pg. 
138
 
Rawlings
S.
Saunders
R.
Eales
S. A.
Mackay
C. D.
MNRAS
1989
, vol. 
240
 pg. 
701
 
Sbarrato
T.
Ghisellini
G.
Maraschi
L.
Colpi
M.
MNRAS
2012
, vol. 
421
 pg. 
1764
  
(TS12
Sbarufatti
B.
Treves
A.
Falomo
R.
ApJ
2005
, vol. 
635
 pg. 
173
 
Shakura
N. I.
Sunyaev
R. A.
A&A
1973
, vol. 
24
 pg. 
337
 
Sharma
P.
Quataert
E.
Hammet
G. H.
Stone
J. M.
ApJ
2007
, vol. 
667
 pg. 
714
 
Shaw
M. S.
et al. 
ApJ
2012
, vol. 
748
 pg. 
49
  
(S12
Shaw
M. S.
et al. 
ApJ
2013
, vol. 
764
 pg. 
135
  
(S13
Shen
Y.
et al. 
ApJS
2011
, vol. 
194
 pg. 
45
 
Smith
M. G.
et al. 
MNRAS
1981
, vol. 
195
 pg. 
437
 
Spinrad
H.
Marr
J.
Aguilar
L.
Djorgovski
S.
PASP
1985
, vol. 
97
 pg. 
932
 
Swain
M. R.
Bridle
A. H.
Baum
S. A.
ApJ
1998
, vol. 
507
 pg. 
29
 
Urry
C. M.
Padovani
P.
PASP
1995
, vol. 
107
 pg. 
803
 
Vermeulen
R. C.
Ogle
P. M.
Tran
H. D.
Browne
I. W. A.
Cohen
M. H.
Readhead
A. C. S.
Taylor
G. B.
Goodrich
R. W.
ApJ
1995
, vol. 
452
 pg. 
L5
 
Willott
C. J.
Rawlings
S.
Blundell
K. M.
Lacy
M.
MNRAS
1999
, vol. 
309
 pg. 
1017
 
Yuan
F.
Narayan
R.
ARA&A
2014
, vol. 
525
 pg. 
529
 

APPENDIX A

We show in Figs A1A2 and A3 the SEDs of the 25 objects from S13, divided according to our reclassification, discussed in Section 4. The SEDs are fitted with a one-zone leptonic model, fully described in Ghisellini & Tavecchio (2009).

Sources reclassified as FS. The blue solid lines represent the overall models. The green solid line is the synchrotron components, the black dashed line is the thermal emission from accretion disc, torus and corona, while the grey long-dashed line is the synchrotron self-Compton emission.
Figure A1.

Sources reclassified as FS. The blue solid lines represent the overall models. The green solid line is the synchrotron components, the black dashed line is the thermal emission from accretion disc, torus and corona, while the grey long-dashed line is the synchrotron self-Compton emission.

Sources reclassified as BL/FS, with optical spectra showing weak broad emission lines, but SEDs more typical of FSRQs. Lines as in Fig. A1.
Figure A2.

Sources reclassified as BL/FS, with optical spectra showing weak broad emission lines, but SEDs more typical of FSRQs. Lines as in Fig. A1.

(a) Sources reclassified as BL/FS. (b) Sources that can be classified as BL Lacs, both from the optical spectra and from the SED aspect. Lines as in Fig. A1.
Figure A3.

(a) Sources reclassified as BL/FS. (b) Sources that can be classified as BL Lacs, both from the optical spectra and from the SED aspect. Lines as in Fig. A1.