Interpretation of Departure from the Broad-line Region Scaling in Active Galactic Nuclei

, , , , , , , , , and

Published 2019 January 11 © 2019. The American Astronomical Society. All rights reserved.
, , Citation Bożena Czerny et al 2019 ApJ 870 84 DOI 10.3847/1538-4357/aaf396

Download Article PDF
DownloadArticle ePub

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

0004-637X/870/2/84

Abstract

Most results of the reverberation monitoring of active galaxies showed a universal scaling of the time delay of the Hβ emission region with the monochromatic flux at 5100 Å, with very small dispersion. Such a scaling favored the dust-based formation mechanism of the broad-line region (BLR). Recent reverberation measurements showed that actually a significant fraction of objects exhibit shorter lags than the previously found scaling. Here we demonstrate that these shorter lags can be explained by the old concept of scaling of the BLR size with the ionization parameter. Assuming a universal value of this parameter and a universal value of the cloud density reproduces the distribution of observational points in the time delay–monochromatic flux plane, provided that a range of black hole spins is allowed. However, a confirmation of the new measurements for low/moderate Eddington ratio sources is strongly needed before the dust-based origin of the BLR can be excluded.

Export citation and abstract BibTeX RIS

1. Introduction

The most prominent features in type 1 active galactic nuclei (AGNs) are broad emission lines present in their spectra (for a review, see, e.g., Krolik 1999). The broad-line region (BLR) is unresolved with current instruments, but the reverberation mapping (RM) of nearby active galaxies, pioneered by Liutyi (1977) and done much more intensively since the 1990s (e.g., Kaspi et al. 2000; Peterson et al. 2004; Bentz et al. 2013; Du et al. 2018), has allowed us to measure the size of the BLR from the time delay between the variations of emission lines and the continuum. This, in turn, opened a way to measure the black hole mass by combining the radius from the time delay with the orbital velocity of the BLR clouds estimated from the emission-line width and assuming the Kepler law for that purpose.

Numerous RM campaigns revealed a very strong and tight correlation between the BLR size and the monochromatic luminosity at 5100 Å (Peterson et al. 2004; Bentz et al. 2013). With this scaling, supermassive black hole mass measurements became possible, based just on a single spectrum (e.g., Vestergaard & Peterson 2006). This, in turn, opened a way for cosmological applications, since after proper calibration the time delay measurement allows us to determine the luminosity and to use a generalized standard candle approach to obtain the cosmological parameters (Haas et al. 2011; Watson et al. 2011; Czerny et al. 2013; King et al. 2014).

The problem has started with the detection of some outliers from the radius–luminosity relation. First, outliers have been found among the highly super-Eddington sources (Du et al. 2015, 2016, 2018), and their much shorter time delay than implied by the standard radius–luminosity relation (Bentz et al. 2013) could be interpreted as an effect of the self-shielding in the disk emission (Wang et al. 2014b).

Recently, shorter-than-expected time delays also were measured in some low Eddington ratio sources (Grier et al. 2017; Du et al. 2018). This poses a question about the nature of the standard radius–luminosity relation and the physical reasons for the departures from this law. These shortened lags could be explained by retrograde accretion (Wang et al. 2014a; Du et al. 2018). Actually, the averaged radius of the BLR depends on the ionizing spectral energy distributions (SEDs) and spatial distributions of BLR clouds. The increase of scatterers around the canonical RL relation indicates a lack of understanding of the ionizing sources and the BLR itself.

Most of the models assume that the BLR is quasi-spherical, the radial extension of the cloud formation is not specified (e.g., Pancoast et al. 2014), and the clouds are exposed to the nuclear emission, so the BLR emissivity should respond to the bolometric luminosity of the nucleus or, more precisely, to the available ionizing continuum. This is why the ionization parameter U (see, e.g., Wandel et al. 1999) was used in most of the past BLR modeling. Some models specified the inner BLR radius on some physical grounds (e.g., disk local self-gravity, Wang et al. 2011, 2012; dust presence in the accretion disk atmosphere, Czerny & Hryniewicz 2011; Czerny et al. 2015, 2017).

However, the relation between the 5100 Å monochromatic luminosity, L5100, and the ionizing flux is nonlinear and depends on the SED of the source. Large black hole mass, low Eddington ratio, and low spin lead to significant curvature of the spectrum in the UV band (e.g., Richards et al. 2006; Capellupo et al. 2015). In lower Eddington sources the inner disk may not be well represented by the standard accretion disk, which effectively gives a similar result (see, e.g., Kubota & Done 2018, and the references therein). Therefore, as argued by Wang et al. (2014a), we should expect significant dispersion if we use L5100 as a proxy for the ionizing flux, unless the spectrum shows no curvature owing to a very high black hole spin in all sources. The need to return to the scaling with bolometric luminosity has been suggested by Trippe (2015). Kilerci Eser et al. (2015) showed observationally, using seven well-studied AGNs, that the relation between UV and optical flux is nonlinear, and UV flux offers a much better proxy for the ionizing flux and leads to the lower dispersion in the radius–luminosity relation.

In the present paper we adopt a general approach based on the assumption that the size of the BLR scales with the ionizing flux. However, as in Wang et al. (2014a), we take into account that the ionizing flux is not a linear function of the monochromatic luminosity. We consider the predictions for the ionizing flux from an accretion disk around a spinning black hole, together with the possibility of a counterrotating disk and the possible existence of the inner hot flow instead of a standard optically thick cold accretion disk.

2. Method

We consider the standard scenario of the BLR sensitive to the full ionizing flux available. In this model, the relation between the ionizing flux and the monochromatic luminosity at 5100 Å is not linear, and the size of the BLR is expected to depend on the black hole mass and the Eddington ratio. Additionally, there are two effects: (i) the spin of the black hole plays an important role, and (ii) in the case of low Eddington ratio AGNs the inner radius of the optically thick accretion disk might not be located at the innermost stable circular orbit (ISCO) but farther out, and in the innermost part the flow is replaced by optically thin and hot advection-dominated accretion flow (ADAF). We consider these effects separately. The role of the spin as a possible source of dispersion and systematic departures from the simple radius–luminosity trend has already been studied by Wang et al. (2014a), but here we generalize the method and study the latter option of the inner ADAF.

2.1. Accretion Disk Model and Ionizing Flux

Since Wang et al. (2014a) showed that, most likely, a high value of spin is required, in the present paper we use the Novikov–Thorne model of the accretion disk (Novikov & Thorne 1973; Wang et al. 2014a). We combine the local emissivity with the ray-tracing procedure applied earlier in Czerny et al. (2011), which allows us to include all relativistic corrections, including the light bending, gravitational redshift, and Doppler boosting. The disk model is thus parameterized by the black hole mass, M, accretion rate, ${\dot{M}}_{\bullet }$, and dimensionless spin parameter, a. We consider both prograde and retrograde spin values.

The calculation of the observed monochromatic luminosity in this model depends on the viewing angle, i, between the symmetry axis and the observer. We do not know this angle in individual sources, but the expected range of angles is constrained by the presence of the dusty molecular torus (see Krolik 1999). The torus blocks the view toward the nucleus at high viewing angles; such sources do not show their BLR, and they are classified as type 2 sources. Observational constraints on the torus opening angle are not simple; they typically imply a mean torus opening angle of order of 45, and this value does not depend strongly on the source luminosity (e.g., Tovmassian 2001; Lawrence & Elvis 2010; He et al. 2018). Taking into account this constraint and assuming otherwise random orientation of AGNs with respect to us, we adopt viewing angle i ∼ 30° as a representative value. Thus, we measure the observed fluxes assuming the same viewing angle for all the sources. We neglect here the effect of finite wavelength width for the hydrogen cross section (see Equation (5) of Wang et al. 2014a) and determine the luminosity at 13.6 eV (1 ryd) as

Equation (1)

However, the BLR sees a different part of the spectrum. The BLR covers from 10% to 30% of the sky from the point of view of the inner disk, and it intercepts photons propagating relatively close to the equatorial plane (but not too close since highly inclined photons are intercepted by the disk itself). So, when calculating the number of ionizing photons, we integrate the spectrum above 1 ryd over all viewing angles between 80° and 45°:

Equation (2)

The integration is performed far from the black hole, where the general relativity effects are already negligible. Photons propagating at angles larger than 80 are neglected since they will be absorbed by the disk itself. The ionization level of the BLR clouds can be estimated using either Lion or Q.

2.2. Accretion Disks with Inner ADAF

The transition from the outer cold disk to the inner hot disk is still debated, particularly for the galactic sources. However, it is clear that in very low luminosity sources (the most extreme case is Sgr A*) there is no cold outer disk. A series of papers discussed this issue on the basis of the radiative and conductive interaction between the hot corona above the disk and the underlying cold disk, and it was shown that for very low accretion rates the cold inner disk disappears (see Yuan & Narayan 2014, for a review). The inner flow then proceeds through an optically thin hot flow, such as ADAF (Ichimaru 1977; Narayan & Yi 1994) or its alternatives, e.g., advection-dominated inflow–outflow solutions (ADIOS; Blandford & Begelman 1999). Most of these solutions can be described under the common name RIAF (radiatively inefficient accretion flow), but some are actually quite radiatively efficient if the strong coupling exists between the hot ions and electrons (Bisnovatyi-Kogan & Lovelace 1997; Sironi & Narayan 2015). However, the common property of these solutions is that ion temperature is close to virial temperature, and electrons are also relatively hot, so the emitted radiation concentrates in X-rays instead of far-UV local blackbody emission characteristic for the cold accretion disks.

For the purpose of this paper we use two prescriptions from Czerny et al. (2004). The first one is simply based on the strong ADAF principle, i.e., whenever the ADAF solutions exist, the flow proceeds through an ADAF-type flow, as in the classical papers (Abramowicz et al. 1995; Honma 1996; Kato & Nakamura 1998). The second option is based on evaporation of the cold disk caused by electron conduction between the disk and the two-temperature hot corona (Różańska & Czerny 2000b; Meyer & Meyer-Hofmeister 2002).

In the first case the transition from a cold disk to an ADAF flow occurs at

Equation (3)

(see, e.g., Equation (8) in Czerny et al. 2004), where $\dot{m}$ is the dimensionless accretion rate, α0.1 is the viscosity parameter in units of 0.1, and RSchw is the Schwarzschild radius of the black hole (=2${R}_{{\rm{g}}}=2{GM}/{c}^{2}$). Here $\dot{m}$ is defined for a fixed Newtonian efficiency of the accretion process:

Equation (4)

where ${m}_{{\rm{p}}}$ is the proton mass and σT is the Thomson cross section.

In the description of the second scenario we adopt Equation (11) from Czerny et al. (2004) for the transition radius between the outer disk and an inner hot flow since it contains the effect of the magnetic pressure and compares most favorably with the observed extension of the BLR:

Equation (5)

where β is the ratio of the total (gas + radiation) pressure to the total plus magnetic pressure and varies between 0 (magnetic pressure dominance) and 1 (no magnetic pressure). We neglect the emission from the inner hot flow because this very hot plasma, with electron temperature of the order of tens of keV, does not contribute to the 5100 Å flux. This emission can to some extent affect the BLR by Compton-heating the clouds and the intercloud medium, suppressing the line emission. However, since we are interested only in the ionization flux, and not in full radiative transfer computations with cooling/heating balance, we neglect this emission, effectively locating the inner disk radius at Revap. In this scenario the dependence on the spin practically disappears since the outer disk is only weakly affected by the rotation of the central black hole.

2.3. BLR Radius

We assume the standard view that the localization of the BLR is related to the ionizing flux. When we use Lion defined in Equation (1), we follow the approach of Wang et al. (2014a). We assume the scaling of RBLR with the ionizing flux to have the form of a power law with the same index as determined by Bentz et al. (2013),

Equation (6)

where we specifically took the value of the index from their fit Clean. The value of constant has to be adjusted since now we use Lion instead of L5100 as done in Bentz et al. (2013).

When we use Q as the parameterization of the incident flux, we follow an even more classical approach to modeling of the BLR. It was argued in the past that the BLR properties are well approximated by the fixed value of the ionization parameter, U, and the representative cloud density. The ionization parameter U is defined as

Equation (7)

(see, e.g., Ferland & Netzer 1983), where ne is the representative local density of the cloud and Q is the number of ionizing photons (above 1 ryd) emitted by the accretion disk. We thus calculate the BLR radius from this formula:

Equation (8)

The value of the constant is then related to the universal values of U and ne. The universal characteristic of the cloud density was argued for at the basis of the radiation pressure confinement mechanism (Baskin & Laor 2018).

2.4. Observational Data

We compare the model with the size of the BLR measured as a delay with respect to Hβ line in reverberation campaigns. We use a compilation of the results available in the literature (see Table 1). The measurements come from various groups; most of them were performed for nearby sources. The sample of Bentz et al. (2013) has been carefully corrected for the contamination of the 5100 Å flux by the host galaxy. The sample from Grier et al. (2017) comes from the SDSS–RM (Sloan Digital Sky Survey Reverberation Measurement) project and covers larger redshifts up to z = 1.026. The measurements by Lu et al. (2016) provide an independent determination of the delay in NGC 5548, and Wang et al. (2016) give the delay measured for a gamma-ray-loud NLS1. The sample from Du et al. (2014, 2015, 2016, 2018) represents the SEAMBH (Super-Eddington Accretion in Massive Black Holes) project, so on average these objects have higher Eddington ratios than sources from other samples. In diagrams we mark them with a different color, as they might bias the results. We also give in Table 1 the black hole mass values, taken from the references above, and the Eddington ratio, which we calculate from that mass value and from the monochromatic luminosity assuming a fixed bolometric correction of 9.26 after Shen et al. (2011). Absolute values of the luminosity are given assuming the cosmological parameters: H0 = 67 km s−1 Mpc−1, ΩΛ = 0.68, Ωm = 0.32 (Ade et al. 2014).

Table 1.  Time Delays

Name log L5100 ${R}_{{\rm{H}}\beta }$ $\mathrm{log}\,{M}_{\mathrm{BH}}$ $L/{L}_{\mathrm{Edd}}$ Ref
  ($\mathrm{erg}\,{{\rm{s}}}^{-1}$) (ldt) M
SDSS J140518 44.33 ${41.6}_{-8.3}^{+14.8}$ 7.74 0.288 1
SDSS J140759 43.58 ${16.3}_{-6.6}^{+13.1}$ 7.67 0.059  
SDSS J140812 43.15 ${10.5}_{-2.2}^{+1.0}$ 7.26 0.058  
SDSS J140904 44.15 ${11.6}_{-4.6}^{+8.6}$ 8.45 0.036  
SDSS J141004 44.22 ${53.5}_{-4.0}^{+4.2}$ 8.32 0.058  
SDSS J141018 43.58 ${16.2}_{-4.5}^{+2.9}$ 7.65 0.063  
SDSS J141031 44.02 ${35.8}_{-10.3}^{+1.1}$ 7.91 0.094  
SDSS J141041 43.82 ${21.9}_{-2.4}^{+4.2}$ 7.85 0.070  
SDSS J141112 44.12 ${20.4}_{-2.0}^{+2.5}$ 7.41 0.375  
SDSS J141115 44.31 ${49.1}_{-2.0}^{+11.1}$ 7.94 0.174  
SDSS J141123 44.13 ${13.0}_{-0.8}^{+1.4}$ 7.38 0.411  
SDSS J141135 44.04 ${17.6}_{-7.4}^{+8.6}$ 7.56 0.224  
SDSS J141147 44.02 ${6.4}_{-1.4}^{+1.5}$ 6.95 0.865  
SDSS J141214 44.40 ${21.4}_{-6.4}^{+4.2}$ 7.26 1.019  
SDSS J141314 44.52 ${43.9}_{-4.3}^{+4.9}$ 9.21 0.015  
SDSS J141318 43.94 ${20.0}_{-3.0}^{+1.1}$ 7.51 0.200  
SDSS J141324 43.94 ${25.5}_{-5.8}^{+10.9}$ 8.92 0.008  
SDSS J141417 43.40 ${15.6}_{-5.1}^{+3.2}$ 8.03 0.017  
SDSS J141532 44.14 ${26.5}_{-8.8}^{+9.9}$ 7.23 0.591  
SDSS J141606 44.80 ${32.0}_{-15.5}^{+11.6}$ 9.07 0.040  
SDSS J141625 43.96 ${15.1}_{-4.6}^{+3.2}$ 7.58 0.178  
SDSS J141645.15 43.21 ${5.0}_{-1.4}^{+1.5}$ 7.93 0.014  
SDSS J141645.58 43.68 ${8.5}_{-1.4}^{+2.5}$ 6.90 0.438  
SDSS J141706 44.19 ${10.4}_{-3.0}^{+6.3}$ 6.70 2.258  
SDSS J141712 43.21 ${12.5}_{-2.6}^{+1.8}$ 8.99 0.001  
SDSS J141724 43.99 ${10.1}_{-2.7}^{+12.5}$ 7.57 0.195  
SDSS J141729 43.29 ${5.5}_{-2.1}^{+5.7}$ 8.28 0.008  
SDSS J141856 45.38 ${15.8}_{-1.9}^{+6.0}$ 8.90 0.224  
SDSS J141859 44.91 ${20.4}_{-7.0}^{+5.6}$ 8.05 0.534  
SDSS J141923 43.12 ${11.8}_{-1.5}^{+0.7}$ 7.18 0.065  
SDSS J141941 44.52 ${30.4}_{-8.3}^{+3.9}$ 7.60 0.612  
SDSS J141952 44.25 ${32.9}_{-5.1}^{+5.6}$ 9.23 0.008  
SDSS J141955 43.40 ${10.7}_{-4.4}^{+5.6}$ 7.69 0.037  
SDSS J142010 44.09 ${12.8}_{-4.5}^{+5.7}$ 8.64 0.021  
SDSS J142023 44.22 ${8.5}_{-3.9}^{+3.2}$ 8.58 0.032  
SDSS J142038 43.46 ${25.2}_{-5.7}^{+4.7}$ 7.67 0.045  
SDSS J142039 44.14 ${20.7}_{-3.0}^{+0.9}$ 7.57 0.275  
SDSS J142043 43.40 ${5.9}_{-0.6}^{+0.4}$ 7.20 0.115  
SDSS J142049 44.45 ${46.0}_{-9.5}^{+9.5}$ 9.00 0.020  
SDSS J142052 45.06 ${11.9}_{-1.0}^{+1.3}$ 8.73 0.157  
SDSS J142103 43.64 ${75.2}_{-3.3}^{+3.2}$ 7.89 0.041  
SDSS J142112 44.31 ${14.2}_{-3.0}^{+3.7}$ 8.22 0.092  
SDSS J142135 43.47 ${3.9}_{-0.9}^{+0.9}$ 6.60 0.548  
SDSS J142417 44.09 ${36.3}_{-5.5}^{+4.5}$ 7.70 0.180  
 
Mrk 335 43.76 ${14.0}_{-3.4}^{+4.6}$ 6.93 0.501 2
Mrk 142 43.59 ${6.4}_{-3.4}^{+7.3}$ 6.47 0.983  
IRAS F12397 44.23 ${9.7}_{-1.8}^{+5.5}$ 6.79 2.023  
Mrk 486 43.69 ${23.7}_{-2.7}^{+7.5}$ 7.24 0.208  
Mrk 382 43.12 ${7.5}_{-2.0}^{+2.9}$ 6.50 0.312  
IRAS 04416 44.47 ${13.3}_{-1.4}^{+13.9}$ 6.78 3.577  
MCG 06 42.67 ${24.0}_{-4.8}^{+8.4}$ 6.92 0.042  
Mrk 493 43.11 ${11.6}_{-2.6}^{+1.2}$ 6.14 0.694  
Mrk 1044 43.10 ${10.5}_{-2.7}^{+3.3}$ 6.45 0.322  
SDSS J074352 45.37 ${43.9}_{-4.2}^{+5.2}$ 7.93 2.028  
SDSS J075051 45.33 ${66.6}_{-9.9}^{+18.7}$ 7.67 3.359  
SDSS J075101 44.18 ${30.4}_{-5.8}^{+7.3}$ 7.18 0.733  
SDSS J075949 44.20 ${43.9}_{-19.0}^{+33.1}$ 7.44 0.415  
SDSS J080101 44.27 ${8.3}_{-2.7}^{+9.7}$ 6.78 2.257  
SDSS J080131 43.97 ${11.5}_{-3.7}^{+7.5}$ 6.51 2.121  
SDSS J081441 43.96 ${25.3}_{-7.5}^{+10.4}$ 7.18 0.446  
SDSS J081456 43.99 ${24.3}_{-16.4}^{+7.7}$ 7.44 0.259  
SDSS J083553 44.44 ${12.4}_{-5.4}^{+5.4}$ 6.87 2.722  
SDSS J084533 44.53 ${18.1}_{-4.7}^{+6.0}$ 6.76 4.264  
SDSS J085946 44.41 ${34.8}_{-26.3}^{+9.2}$ 7.30 0.952  
SDSS J093302 44.31 ${19.0}_{-4.3}^{+3.8}$ 7.08 1.245  
SDSS J093922 44.07 ${11.9}_{-6.3}^{+2.1}$ 6.53 2.564  
SDSS J100402 45.52 ${32.2}_{-4.2}^{+43.5}$ 7.44 8.802  
SDSS J101000 44.76 ${27.7}_{-7.6}^{+23.5}$ 7.46 1.456  
SDSS J102339 44.09 ${24.9}_{-3.9}^{+19.8}$ 7.16 0.618  
 
NGC 5548 43.21 ${7.2}_{-0.3}^{+1.3}$ 7.94 0.014 3
1H 0323+342 43.88 ${14.8}_{-2.7}^{+3.9}$ 7.53 0.164 4
 
PG 0026+129 44.97 ${111.0}_{-28.3}^{+24.1}$ 8.15 0.489 5
PG 0052+251 44.81 ${89.8}_{-24.1}^{+24.5}$ 8.64 0.107  
Fairall 9 43.98 ${17.4}_{-4.3}^{+3.2}$ 8.09 0.058  
Mrk 590 43.50 ${25.6}_{-5.3}^{+6.5}$ 7.55 0.065  
3C 120 44.00 ${26.2}_{-6.6}^{+8.7}$ 7.79 0.122  
Ark 120 43.87 ${39.5}_{-7.8}^{+8.5}$ 8.47 0.018  
Mrk 79 43.68 ${15.6}_{-4.9}^{+5.1}$ 7.84 0.050  
PG 0804+761 44.91 ${146.9}_{-18.9}^{+18.8}$ 8.43 0.223  
Mrk 110 43.66 ${25.6}_{-7.2}^{+8.9}$ 7.10 0.265  
PG 0953+414 45.19 ${150.1}_{-22.6}^{+21.6}$ 8.44 0.408  
NGC 3227 42.24 ${3.8}_{-0.8}^{+0.8}$ 7.09 0.010  
NGC 3516 42.79 ${11.7}_{-1.5}^{+1.0}$ 7.82 0.007  
SBS 1116+583A 42.14 ${2.3}_{-0.5}^{+0.6}$ 6.78 0.017  
Arp 151 42.55 ${4.0}_{-0.7}^{+0.5}$ 6.87 0.035  
NGC 3783 42.56 ${10.2}_{-2.3}^{+3.3}$ 7.45 0.009  
Mrk 1310 42.29 ${3.7}_{-0.6}^{+0.6}$ 6.62 0.035  
NGC 4051 41.90 ${2.1}_{-0.7}^{+0.9}$ 5.72 0.110  
NGC 4151 42.09 ${6.6}_{-0.8}^{+1.1}$ 7.72 0.002  
Mrk 202 42.26 ${3.0}_{-1.1}^{+1.7}$ 6.11 0.104  
NGC 4253 42.57 ${6.2}_{-1.2}^{+1.6}$ 6.49 0.088  
PG 1226+023 45.96 ${306.8}_{-90.9}^{+68.5}$ 8.87 0.918  
PG 1229+204 43.70 ${37.8}_{-15.3}^{+27.6}$ 8.03 0.034  
NGC 4593 42.62 ${4.0}_{-0.7}^{+0.8}$ 7.26 0.017  
NGC 4748 42.56 ${5.5}_{-2.2}^{+1.6}$ 6.61 0.064  
PG 1307+085 44.85 ${105.6}_{-46.6}^{+36.0}$ 8.72 0.098  
Mrk 279 43.71 ${16.7}_{-3.9}^{+3.9}$ 7.97 0.040  
PG 1411+442 44.56 ${124.3}_{-61.7}^{+61.0}$ 8.28 0.141  
NGC 5548 43.29 ${17.6}_{-4.7}^{+6.4}$ 7.94 0.014  
PG 1426+015 44.63 ${95.0}_{-37.1}^{+29.9}$ 8.97 0.033  
Mrk 817 43.74 ${19.9}_{-6.7}^{+9.9}$ 7.99 0.042  
Mrk 290 43.17 ${8.7}_{-1.0}^{+1.2}$ 7.55 0.031  
PG 1613+658 44.77 ${40.1}_{-15.2}^{+15.0}$ 8.81 0.068  
PG 1617+175 44.39 ${71.5}_{-33.7}^{+29.6}$ 8.79 0.029  
PG 1700+518 45.59 ${251.8}_{-38.8}^{+45.9}$ 8.40 1.137  
3C 390.3 44.43 ${44.5}_{-17.0}^{+27.6}$ 9.18 0.013  
NGC 6814 42.12 ${6.6}_{-0.9}^{+0.9}$ 7.16 0.007  
Mrk 509 44.19 ${79.6}_{-5.4}^{+6.1}$ 8.15 0.081  
PG 2130+099 44.20 ${9.6}_{-1.2}^{+1.2}$ 7.05 1.043  
NGC 7469 43.51 ${10.8}_{-1.3}^{+3.4}$ 7.60 0.059  
PG 1211+143 44.73 ${93.8}_{-42.1}^{+25.6}$ 7.87 0.530  
PG 0844+349 44.22 ${32.3}_{-13.4}^{+13.7}$ 7.66 0.265  
NGC 5273 41.54 ${2.2}_{-1.6}^{+1.2}$ 7.14 0.002  
KA 1858+4850 43.43 ${13.5}_{-2.3}^{+2.0}$ 6.94 0.225  
Mrk 1511 43.16 ${5.7}_{-0.8}^{+0.9}$ 7.29 0.055  
MCG6-30-15 41.64 ${5.7}_{-1.7}^{+1.8}$ 6.63 0.008  
UGC 06728 41.86 ${1.4}_{-0.8}^{+0.7}$ 5.87 0.073  
MCG+08-11-011 43.33 ${15.7}_{-0.5}^{+0.5}$ 7.72 0.030  
NGC 2617 42.67 ${4.3}_{-1.4}^{+1.1}$ 7.74 0.006  
3C 382 43.84 ${40.5}_{-3.7}^{+8.0}$ 8.67 0.011  
Mrk 374 43.77 ${14.8}_{-3.3}^{+5.8}$ 7.86 0.061  

References. (1) Grier et al. 2017; (2) Du et al. 2014, 2015, 2016, 2018; (3) Lu et al. 2016; (4) Wang et al. 2016; (5) Bentz et al. 2013.

Download table as:  ASCIITypeset images: 1 2 3

3. Results

We compare the measured time delays in monitored AGNs with the model prediction of the position of the BLR radius taking into account two ways of connecting the BLR position with the incident radiation (through Lion defined in Equation (1) and Q defined in Equation (2)). The incident spectrum is calculated for a realistic range of black hole masses (from 106 to 1010 M) and Eddington luminosities (from 0.01 to 1.0). We allow for a broad range of spin, including the case of a counterrotating black hole, and we also allow for the evaporation of the inner disk given by Equation (5).

Since the observational results are always plotted as a delay versus the monochromatic flux at 5100 Å but the ionization flux in general is not a linear function of that flux, we first show the representative examples of the incident spectra for the adopted parameter range (see Figure 1). The relation between ${L}_{\mathrm{ion}}$ and the 5100 Å luminosity is almost linear when the Eddington ratio is large, the black hole mass is small, and the black hole spin is large: in this regime the spectrum up to 912 Å is still well described by the canonical $\nu {L}_{\nu }\propto {\nu }^{4/3}$ law. Outside this regime, the maximum temperature of the accretion disk drops, and the SED peak moves into the UV band, leading to a strong spectral curvature. A similar effect would be caused by the inner disk evaporation, but the presented examples illustrate accretion disks extending down to the ISCO. The disk spectral curvature leads to a slower rise of Lion in comparison to L5100 in a sequence of models with a constant Eddington rate and rising black hole mass. We take into account this effect in interpreting the data, since we calculate the expected time delay from Lion (or Q).

Figure 1.

Figure 1. Examples of the accretion disk spectra for the parameter range covered by our computations. Top panels: spectra for a fixed Eddington ratio of 0.1 and black hole masses of 106 (red), 108 (black), and 1010 M (blue lines). Bottom panels: spectra for a black hole mass of 108 M and Eddington ratios of 0.01 (red), 0.05 (black), 0.1 (blue), and 0.5 (cyan). The three columns represent three spin values. Dotted lines indicate the position of the usual continuum measurement at 5100 Å and the Lyman edge where ionizing flux should be evaluated.

Standard image High-resolution image

We now calculate the expected time delays from our grid of models at the basis of the ionized flux, but we show the resulting delays as a function of the monochromatic flux since the observational results are customarily presented in this way. We compare the predicted delay grid with the measurements of the Hβ delays given in Table 1.

3.1. BLR Size from the Ionizing Luminosity

Since our sample is larger than the sample considered in Wang et al. (2014a), we first use the same method as was used there, with BLR distance measured from ionizing flux Lion estimated at 912 Å (see Equation (1)), and without any evaporation effect. The results are shown in Figure 2.

Figure 2.

Figure 2. Dependence of the size of the BLR, expressed in terms of the time delay, calculated from Lion for a range of spin (separate panels) and Eddington ratio $\dot{m}=0.01$ (red line), $\dot{m}=0.05$ (black line), $\dot{m}=0.1$ (cyan line), and $\dot{m}=0.5$ (blue line), as a function of monochromatic luminosity at 5100 Å. Observational points come from Table 1; super-Eddington sources are marked in red.

Standard image High-resolution image

We see that the change of the sample essentially affects the conclusion reached by Wang et al. (2014a). Their conclusion was that all the monitored objects must have predominantly high spin since at that time most of the measured delays were located along the $\tau \propto {L}_{5100}^{1/2}$ line. Currently, with the presence of many shorter lags in the sample, good coverage of the measured distribution of the time delays is achieved if we allow a whole range of black hole spins from 0 to maximally rotating black holes. A few objects require counterrotating spin. What is interesting, however, is that these objects have rather moderate luminosities ($\sim {10}^{44}$ erg s−1), relatively low Eddington ratios (∼0.05), and black hole masses that are also moderate (∼108–109 M). These values are roughly consistent with the parameters of the sources in the SDSS–RM sample of Grier et al. (2017). Such accretion rates are relatively low for a quasar sample, but this is the consequence of the selection effect: the monitoring was short, about a year, so the delays were measured only for the sources with delays shorter than 100 days in the observed frame, so only for the low-luminosity tail of the sample. The model predicts even shorter time delays than measured for more massive sources, but observationally determined delays do not populate this region. This is because we allowed for even lower Eddington ratios (0.01) and larger masses (1010 M) in our parameter grid than present in the sample.

3.2. BLR Size from the Number of Ionizing Photons

Next, we use the RBLR predictions based on the number of photons Q intercepted by the BLR (see Equation (2)), again without any evaporation effect, and the results are shown in Figure 3. The model predictions are qualitatively similar, but not identical. Overall, the Q prescription gives a much stronger bending effect for a counterrotating accretion disk than the model based on Lion for the same parameters. The trend reverses for a high-spin corotating accretion disk, when a smaller departure from a power-law trend is seen for the Q-based model. This is related to the relativistic effects. For large black hole spin in a corotating disk the radiation emitted at higher inclination angles is strongly beamed and enhanced in comparison with the continuum measured by an observer, which compensates for the spectral bending seen in Figure 1. This also means that Q is not strictly proportional to Lion. We illustrate this effect in Figure 4. The departure from the strict linearity between the two quantities results both from the derivation of Q as an integral and from the relativistic effects (a difference between photons going to the observer and photons going toward the BLR). The approach based on Q is more accurate, and the examples calculated later on are based on this assumption, but overall the difference between Lion and Q predictions is not large, and the Lion approach is equally useful for statistical analysis of the samples.

Figure 3.

Figure 3. Same as Figure 2, but for the size of the BLR calculated from Q.

Standard image High-resolution image
Figure 4.

Figure 4. Relation between the ionization luminosity, ${L}_{\mathrm{ion}}$, and the number of ionizing photons, Q, for the size of the BLR for $a=-1.0$ (dashed lines) and a = 0.998 (solid lines), for two values of the accretion rate, $\dot{m}=0.01$ (red lines) and $\dot{m}=0.5$ (blue lines). Both ${L}_{\mathrm{ion}}$ and Q are not monotonic functions of the black hole mass, as reflected on the turnover loop, but the two quantities are not strictly proportional to each other.

Standard image High-resolution image

Now we compare Q-based delay predictions to the measured time delays (see Figure 3). This approach does not require counterrotating black holes to explain even the shortest time lags. Therefore, with very simple and basic assumptions (standard Novikov–Thorne disk, corotating with a black hole, with a range of masses, accretion rates, and spins, BLR responding to the ionizing continuum at fixed ionization parameter and density), we are able to reproduce fully satisfactorily the observed distribution of the time delays.

However, the assumption that the cold Keplerian disk extends all the way down to the ISCO is under discussion, particularly for lower Eddington ratio sources. If the inner hot flow develops, this part of the disk is no longer a source of UV photons but instead, filled with a hot plasma at electron temperatures of order of 100 keV, is a source of X-ray emission. Effectively this decreases the number of photons available for ionizing hydrogen. Available models allow us to determine the position of this transition and thus are the subject of tests if the predictions are consistent with the observed time delays.

3.3. BLR Size with Inner Hot Flow in Classical ADAF Scenario

We first test the prediction of the model based on the strong ADAF principle described by Equation (3). For the viscosity parameter, we assume the value α = 0.02, which was suggested by direct studies of the quasar UV variability (Siemiginowska & Czerny 1989; Starling et al. 2004) and is consistent with damped random walk results (Kelly et al. 2009; see also the discussion in Grzȩdzielski et al. 2017). The results are shown in Figure 5. In this case no counterrotating spin values are necessary, and the whole plane is well covered even if all the objects have high spin (left panel). The longest delays for a given value of L5100 require high spin values, while shorter delays may imply either lower spin or lower Eddington ratio. This degeneracy can be removed if we actually have reliable determination of the Eddington ratio for an individual object. However, in comparison with the predictions for the disk without an inner cutoff, the requested accretion rates of sources with short delays are much higher, and all observed sources are then predicted to have Eddington ratios above ∼0.05. Some of the Eddington ratios in the SDSS–RM sample may be lower than this limit if the bolometric luminosity is estimated as roughly 9 times the L5100 luminosity and the black hole mass measurement from Grier et al. (2017) is adopted. On the other hand, such an estimate of the bolometric luminosity is highly uncertain since it does not take into account the differences in the spectral shapes clearly seen in Figure 1.

Figure 5.

Figure 5. Dependence of the size of the BLR calculated from Q when the ADAF principle is assumed, for a range of spins (separate panels) and Eddington ratios $\dot{m}=0.01$ (red line), $\dot{m}=0.05$ (black line), $\dot{m}=0.1$ (cyan line), and $\dot{m}=0.5$ (blue line) as a function of monochromatic luminosity at 5100 Å. Model parameter: $\alpha =0.02$. Observational points come from Table 1; super-Eddington sources are marked in red.

Standard image High-resolution image

3.4. BLR Size with Inner Hot Flow Defined by Inner Disk Evaporation

Finally, we test the solutions based on the evaporation of the inner disk as described by Equation (5). The results obtained for the parameters $\alpha =0.02$ and β = 0.5 (moderate magnetic field strength; see Equation (5)) are far from being satisfactory (see Figure 6). The longest time delays are not reproduced, even if we allow for high spin since the transition radius between the outer cold disk and an inner hot flow is far too large. A significant contribution of the magnetic field to the total pressure in the disk is inconsistent with the observational data. We thus decreased the role of the magnetic field, assuming β = 0.99 (i.e., only 1% of the contribution from the magnetic pressure to the gas plus radiation pressure), but such a parameter adjustment still did not fully solve the problem (see Figure 7). We thus additionally allowed for a significant decrease of the viscosity parameter down to the value α = 0.001, and this allowed to reconcile the model and the data (see Figure 8). For these parameters, evaporation of the outer cold disk is indeed inefficient. The transition radius between the outer cold disk and an inner hot flow takes place at 5.7RSchw for the Eddington ratio 0.01, and it is closer in or absent for larger accretion rates and/or lower spin. When we assumed the expected values of the viscosity parameter α = 0.02, the transition radii start at 62.3RSchw (for the lowest Eddington ratio 0.01). This low viscosity α ∼ 0.001 is thus strongly required if the measured time delays are to be consistent with the disk evaporation phenomenon.

Figure 6.

Figure 6. Dependence of the size of the BLR calculated from Q when evaporation of the inner disk is included, for a maximally counterrotating spin (left panel; ISCO radius 9Rg) and maximally corotating spin (right panel; ISCO radius 1.24Rg) and Eddington ratios $\dot{m}=0.01$ (red line), $\dot{m}=0.05$ (black line), $\dot{m}=0.1$ (cyan line), and $\dot{m}=0.5$ (blue line) as a function of monochromatic luminosity at 5100 Å. Evaporation model parameters: $\alpha =0.02$, $\beta =0.5$. Observational points come from Table 1; super-Eddington sources are marked in red.

Standard image High-resolution image
Figure 7.

Figure 7. Same as Figure 6, but for the magnetization parameter β = 0.99.

Standard image High-resolution image
Figure 8.

Figure 8. Dependence of the size of the BLR calculated from Q when evaporation of the inner disk is included, for a range of spin (separate panels) and Eddington ratio $\dot{m}=0.01$ (red line), $\dot{m}=0.05$ (black line), $\dot{m}=0.1$ (cyan line), and $\dot{m}=0.5$ (blue line) as a function of monochromatic luminosity at 5100 Å. Evaporation model parameters: $\alpha =0.001$, $\beta =0.99$. Observational points come from Table 1; super-Eddington sources are marked in red.

Standard image High-resolution image

However, as we mentioned before, such small values of the viscosity parameter are not supported by the quasar variability (see Grzȩdzielski et al. 2017). This implies that the physically based description of the disk evaporation and the transition to inner hot flow still requires a more advanced approach. The issue is difficult and has been the subject of studies for several years in the context of cataclysmic variables (Meyer & Meyer-Hofmeister 1994), binary black holes, and AGNs (Liu et al. 1999; Różańska & Czerny 2000a). The transition process is further complicated by the possibility of the existence of the inner cool disk separated by the gap of pure hot flow from the outer cold disk (Liu et al. 2006; Meyer et al. 2007; see Meyer-Hofmeister & Meyer 2014; Meyer-Hofmeister et al. 2017; Taam et al. 2018 for recent developments). Here we considered only the inner radius of the outer cold disk.

4. Discussion

4.1. The BLR Models

Emission lines of the BLR respond to the emission originating close to the black hole, as clearly seen from the response of the line with respect to the variable incident flux. Traditionally, this response was modeled by parameterizing the incident continuum with a single parameter in the form of the ionization parameter, or ionizing flux. Then, if a range of radii and densities were allowed, like, for example, in the LOC model (Baldwin et al. 1995), line ratios were successfully modeled. However, this general approach did not give a direct insight into the reason for the presence of the numerous clouds above the disk. Wind models are also parametric and not quite constraining the cloud formation mechanisms. The observational discovery of the scaling of the BLR size with a monochromatic flux instead of the ionizing flux opened a way for testing the cloud origin. Such a scaling, with a very small dispersion, is consistent with the dust-based formation mechanism of the BLR (Czerny & Hryniewicz 2011; Czerny et al. 2015, 2017; Baskin & Laor 2018), at least with respect to the Hβ formation region. In this model the BLR position is controlled not by the ionizing flux but by the availability of the material for the irradiation, lifted above the disk by the dust-driven radiation pressure. This scaling also strongly disfavored self-gravitational instability as the BLR origin (Czerny et al. 2016).

New observational data presented in Figure 2 drastically change this view. Now the dispersion around the delay–monochromatic flux relation is very large, and most of the new measurements lie below the previous tight scaling law of Bentz et al. (2013). Part of the outliers come from the super-Eddington ratio sample by Du et al. (2014, 2015, 2016, 2018), and in this case the departure from the previous scaling can be explained as a departure from the thin Keplerian disk approximation for the incident continuum. The average value of the Eddington ratio in these sources is 1.72. However, other numerous points come from the SDSS–RM sample of Grier et al. (2017), and these objects have low to moderate Eddington ratios, with the average value of 0.24, not significantly higher than in the Bentz et al. (2013) sample (0.14), well within the dispersion in the two samples.

The measurements we use here come from the literature, as described in Section 2.4, and were not prepared in a uniform way. All sources in the Bentz et al. (2013) sample were carefully corrected for the host galaxy contamination using Hubble Space Telescope (HST) observations, and the corrections were important not just for the faintest nearby AGN but also for PG quasars (Bentz et al. 2006). A good example from Bentz et al. (2013) is the source SBS 116+583A (z = 0.02787), where the AGN emission contributes only 12% to the total flux at 5100 Å, i.e., not correcting this source for starlight gives a shift of 0.92 in the logarithm of L5100, or, equivalently, a factor of 2.9 in the expected time delay. The sources from Du et al. (2014) were also corrected for the starlight using HST images, while for the sources reported later (Du et al. 2015, 2016, 2018) the authors used an empirical relation from Shen et al. (2011). Grier et al. (2017) used the information on host contamination from Shen et al. (2015), who employed principal component analysis and the measurements of the measured stellar velocity dispersion to decompose the spectra into AGNs and host components. The accuracy of such methods is difficult to estimate. The error in the host subtraction can be responsible for the incorrect determination of L5100.

New results, if reliable, take us all the way back to the original concept of the simple BLR response to the ionizing flux. The range of delays is well consistent with the predictions of the ionized continuum based on the thin Keplerian disk since this ionized continuum does not scale linearly with the monochromatic flux at 5100 Å owing to the curvature of the spectrum in the UV band (see Figure 1). Models based on the dusty origin of the BLR seem not justified, as in this case no departure from the Bentz et al. (2013) scaling is predicted for a broad range of black hole masses and accretion rates. Instead, the observed delays are consistent with a universal ionization parameter and universal density in the Hβ formation region for a very broad range of parameters.

The range of spin plays an important role, and all spins from nonrotating black holes to maximally spinning ones (a = 0.998) are requested to create the appropriate representation of the reverberation studied sources. Counterrotating disks are not strongly required, but a small fraction of such sources is not excluded. Evaporation of the inner disk and a transition to ADAF are not really required for Eddington ratios above 1% studied here, and models that predict inner ADAF flow at such high Eddington ratios are disfavored, so the models with an inner cold disk separated from the other cold disk sound more attractive. However, the predictions for such models are more complex and were not tested in detail in the current paper. In this paper we also did not test again the self-gravity scenario, but since self-gravity in general predicts shorter delays than the Bentz et al. (2013) relation, it remains to check whether this option offers equally good coverage of the parameter space as a simple universal ionization parameter with the universal density model.

Recently, a new model of the BLR origin was suggested by Wang et al. (2017). In this model, clumps in the dusty/molecular torus are tidally captured and disrupted by the central black hole. A population of inflowing clouds forms at the first stage, a tiny fraction of clumps is channeled into outflows (less than 10%), and then most of the disrupted clumps form a BLR disk with Keplerian rotation (virialized component). Unlike the dust-based clouds (Czerny & Hryniewicz 2011), the supply of the BLR clouds originates from the dusty torus. The virialized component in this model can produce the canonical RL relation for sub-Eddington AGNs provided that the ionization parameter, density, and temperature are universal. Additionally, the infalling component is supported by PG 2130+099 (see Section 4.3). This model avoids difficulties of the disk self-gravity model or dust-based failed winds.

4.2. Implications for BH Evolution

What we found in this paper is the role of black hole spin in the observed RL relation jointly with accretion rates. The evidence in support of retrograde accretion onto black holes has very important implications for cosmological evolution of black holes. If the black holes are fueled in a stochastic manner, with no preferred orientation, they are slowly spinning owing to cancellation of random angular momentum of accreted gas (King et al. 2008). Wang et al. (2009) built up an equation of the radiative efficiency (η) across cosmic time from observed data of galaxy and quasar surveys. The authors derived the evolutionary curve and obtained that the radiative efficiency changes from η ≈ 0.3 at redshift z ≈ 2, where quasar density peaks, down to 0.03 at low redshift. This supports the role of episodic accretion at later stages of the galaxy evolution. The downsizing behavior of spins was further discussed by Li et al. (2012). Subsequently, Volonteri et al. (2013) found a similar behavior of spin evolution from numerical simulations. The sensitive dependence of the RL relation on spin offers a new tool of estimating black hole rotation, in particular for those AGNs with the retrograde accretion disks, which may have too weak gravitational effects on iron Kα line profiles to measure their spins from X-ray observations.

4.3. Future RM Campaigns

To draw firm conclusions, however, the measured delays have to be accurate. The SDSS–RM campaign concentrated on average on higher-redshift sources, the campaign was relatively short, and the observational cadence was not very dense. Determination of the time delays in AGNs is not very straightforward since the variability has a red-noise character and the reprocessing region is extended. Frequently, two peaks show up in the cross-correlation function (e.g., Du et al. 2016), and if the campaign is too short, only one solution (the shorter one) may be found, although it may not be actually the correct one. Therefore, an extension of this campaign is clearly necessary to ensure that the measured delays are not affected by the way in which they are performed.

The cadence selected for the RM campaign is very important. A low-candence campaign will smear short-timescale variations, so that the measured lags tend to be longer. PG 2130+099 is an example. Kaspi et al. (2000) measured an Hβ line lag of 188 days with a low cadence of about ∼20 days and a couple of seasonal gaps, with the total campaign duration as long as about 8 yr. With a cadence of a few days, however, Grier et al. (2008) measured a lag of ∼23 days with large uncertainties (but the total length in their case was only about 100 days). Moreover, Grier et al. (2012) got a lag of ∼10 days with a cadence of 1 day. Hu et al. (2018) (ApJ, submitted) measured a lag of ∼24 days with a cadence of 3 days, confirming the results of Grier et al. (2008), but found that the lag of ∼188 days follows the dust reverberation scaling relation (Koshida et al. 2014), suggesting that the reverberations with a lag of 188 days are from the inner edge of the torus. This supports the ideas of Wang et al. (2017). Additionally, PG 2130+099 is a super-Eddington source, and the other two lags of ∼10 and ∼24 days can be explained by the self-shadowing effects of slim accretion disks (Wang et al. 2014b). Overall, the BLR is extended, and the choice of the cadence can focus the monitoring on a particular part of the BLR, making a comparison of results for different sources very difficult. Also, the nonlinear response of the BLR to the line emission, particularly if the characteristic variability timescale is short in comparison to the average time delay, can easily lead to apparent shortening of the time delay, as pointed out by Goad & Korista (2014). Future RM campaigns should be planned very carefully with respect to the cadence.

5. Conclusions

In this paper, we have tested roles of the energy distribution of accretion disks governed by black hole spins and accretion rates to the RL relation. Our main conclusions can be summarized in several points:

  • 1.  
    New measurements of the time delays in AGNs, inconsistent with the simple scaling relationship of Bentz et al. (2013) with the monochromatic flux, favor a model where the BLR responds to the ionizing continuum, and both the local density and the ionization parameter are universal, independent of the black hole mass and Eddington ratio. Inconsistency between the results in Grier et al. (2017) and Bentz et al. (2013) does not seem to be related to the Eddington ratio, only slightly higher in the Grier et al. sample.
  • 2.  
    Since new BLR scaling is sensitive to the SED shape, the radius–luminosity relation is a potential tool to examine the SED from an accretion disk.
  • 3.  
    If the transition to the inner hot flow is based on the strong ADAF principle, the measured delays are consistent with the model for a realistic value of the viscosity parameter α because it does not overestimate the cold disk evaporation.
  • 4.  
    If the transition to the inner hot flow is based on the disk evaporation through the electron conduction between the disk and the corona, low values of the magnetic pressure and very low values of the viscosity parameter α are required, so this description is less satisfactory than the simple strong ADAF principle and most likely implies that the inner cold disk formation takes place.
  • 5.  
    New measurements can be seen as a counterargument against the dust-based model of the BLR formation since the dust-based model implies a scaling of the BLR size with the monochromatic flux.
  • 6.  
    A confirmation of the new reverberation measurements for the outliers from the Bentz et al. (2013) relationship at low/moderate Eddington ratios is strongly required.

The project was partially supported by National Science Centre, Poland, grant no. 2017/26/A/ST9/00756 (Maestro 9). V.K. acknowledges Czech Science Foundation no. 17-16287S.

Please wait… references are loading.
10.3847/1538-4357/aaf396