This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

TESTING THE UNIFICATION MODEL FOR ACTIVE GALACTIC NUCLEI IN THE INFRARED: ARE THE OBSCURING TORI OF TYPE 1 AND 2 SEYFERTS DIFFERENT?

, , , , , , , , , and

Published 2011 March 29 © 2011. The American Astronomical Society. All rights reserved.
, , Citation C. Ramos Almeida et al 2011 ApJ 731 92 DOI 10.1088/0004-637X/731/2/92

0004-637X/731/2/92

ABSTRACT

We present new mid-infrared imaging data for three Type-1 Seyfert galaxies obtained with T-ReCS on the Gemini-South Telescope at subarcsecond resolution. Our aim is to enlarge the sample studied in a previous work to compare the properties of Type-1 and Type-2 Seyfert tori using clumpy torus models and a Bayesian approach to fit the infrared (IR) nuclear spectral energy distributions. Thus, the sample considered here comprises 7 Type-1, 11 Type-2, and 3 intermediate-type Seyferts. The unresolved IR emission of the Seyfert 1 galaxies can be reproduced by a combination of dust heated by the central engine and direct active galactic nucleus (AGN) emission, while for the Seyfert 2 nuclei only dust emission is considered. These dusty tori have physical sizes smaller than 6 pc radius, as derived from our fits. Unification schemes of AGN account for a variety of observational differences in terms of viewing geometry. However, we find evidence that strong unification may not hold and that the immediate dusty surroundings of Type-1 and Type-2 Seyfert nuclei are intrinsically different. The Type-2 tori studied here are broader, have more clumps, and these clumps have lower optical depths than those of Type-1 tori. The larger the covering factor of the torus, the smaller the probability of having a direct view of the AGN, and vice versa. In our sample, Seyfert 2 tori have larger covering factors (CT = 0.95 ± 0.02) and smaller escape probabilities (Pesc = 0.05% ± 0.080.03%) than those of Seyfert 1 (CT = 0.5 ± 0.1; Pesc = 18% ± 3%). All the previous differences are significant according to the Kullback–Leibler divergence. Thus, on the basis of the results presented here, the classification of a Seyfert galaxy as a Type-1 or Type-2 depends more on the intrinsic properties of the torus rather than on its mere inclination toward us, in contradiction with the simplest unification model.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Observational evidence in the X-rays and the mid-infrared (MIR) indicates that the strong active galactic nucleus (AGN) continuum source must be absorbed by obscuring material over a wide solid angle (see, e.g., Antonucci & Miller 1985; Maiolino et al. 1998; Risaliti et al. 2002). According to observed spectra of different AGN types, the obscuring structure has to block the emission of the subparsec-scale broad-line region (BLR) where the broad lines are produced, but not that of the kiloparsec-scale narrow-line region (NLR).

The unified model for active galaxies (Antonucci 1993; Urry & Padovani 1995) is based on the existence of a dusty toroidal structure surrounding the central region of an AGN. This toroidal geometry explains the biconical shapes observed in Hubble Space Telescope (HST) imaging of several AGNs (Tadhunter & Tsvetanov 1989; Malkan et al. 1998; Tadhunter et al. 1999) and also the polarimetric observations (Antonucci & Miller 1985; Packham et al. 1997). Thus, considering this geometry of the obscuring material, the central engines of Type-1 AGN can be seen directly, resulting in typical spectra with both narrow and broad emission lines, whereas in Type-2 AGN the BLR is obscured.

Pioneering work in modeling the dusty torus (Pier & Krolik 1992, 1993; Granato & Danese 1994; Efstathiou & Rowan-Robinson 1995; Granato et al. 1997; Siebenmorgen et al. 2004) assumed a uniform dust density distribution to simplify the modeling, although from the start, Krolik & Begelman (1988) realized that smooth dust distributions cannot survive within the AGN vicinity. They proposed instead that the material in the torus must be distributed in a clumpy structure, in order to prevent the dust grains from being destroyed by the hot surrounding gas.

The infrared (IR) range (and particularly the MIR) is key to set constraints on the torus models, since the reprocessed radiation from the dust in the torus is re-emitted in this range. However, in comparing the predictions of any torus model with observations, its small-scale emission must be isolated. High angular resolution is then essential to separate torus emission from stellar emission and star-heated dust in the near-infrared (NIR) and MIR, respectively. Indeed, starlight dominates the nuclear NIR emission of Seyfert 2 galaxies when using large aperture data (see, e.g., Alonso-Herrero et al. 1996) and still has a significant contribution for Seyfert 1 galaxies (Kotilainen et al. 1992). Similar contamination problems can be present in the MIR with the star-heated dust and dust in the ionization cones (Alonso-Herrero et al. 2006; Mason et al. 2006).

Another controversial issue about the torus structure is its typical dimensions. Pier & Krolik (1993) and Granato & Danese (1994) reproduced the IR observations of nearby Seyfert galaxies with ∼100 pc scale tori. However, hard X-ray observations showed that about half of nearby Type-2 Seyferts are Compton-thick (i.e., they are obscured by a column density higher than 1024 cm−2; Risaliti et al. 1999). For these highly obscured sources, the torus dimensions are expected to be of a few parsecs, because otherwise the dynamical mass of the obscuring material would be too large to be realistic (Risaliti et al. 1999). In addition, recent ground-based MIR observations of nearby Seyferts reveal that the torus size is likely restricted to a few parsecs. Packham et al. (2005) and Radomski et al. (2008) established upper limits of 2 and 1.6 pc for the outer radii of the Circinus galaxy and Centaurus A tori, respectively. Besides, interferometric observations obtained with the MIR Interferometric Instrument (MIDI) at the Very Large Telescope Interferometer (VLTI) of Circinus, NGC 1068, and Centaurus A suggest a scenario where the torus emission would only extend out to R = 1 pc (Tristram et al. 2007), R = 1.7–2 pc (Jaffe et al. 2004; Raban et al. 2009), and R = 0.3 pc (Meisenheimer et al. 2007), respectively.

In order to solve the discrepancies between observations and previous models, an intensive search for an alternative torus geometry has been carried out in the last decade. The first results of radiative transfer calculations of a clumpy rather than a smooth medium were reported by Nenkova et al. (2002) and Elitzur & Shlosman (2006), and further work was done by Dullemond & van Bemmel (2005). The clumpy dusty torus models (Nenkova et al. 2002, 2008a, 2008b; Hönig et al. 2006; Schartmann et al. 2008) propose that the dust is distributed in clumps, instead of homogeneously filling the torus volume. These models are making significant progress in accounting for the MIR emission of AGNs (Mason et al. 2006, 2009; Mor et al. 2009; Horst et al. 2008, 2009; Nikutta et al. 2009; Ramos Almeida et al. 2009b; Hönig et al. 2010).

In our previous work (Ramos Almeida et al. 2009b; hereafter RA09), we constructed subarcsecond-resolution IR spectral energy distributions (SEDs) for 18 Seyfert galaxies, mostly Type-2 Seyferts. From the comparison between our high angular resolution MIR fluxes and large aperture data, such as those from ISO, IRAS, or Spitzer, we confirmed that the former provide a spectral shape that is substantially different from that of the large aperture data (Rodríguez Espinosa & Pérez García 1997). Since our nuclear measurements allowed us to better characterize the torus emission, we modeled our SEDs with clumpy torus models. In general, we found that Type-2 views are more inclined than those of Type-1s, and more importantly, we derive larger covering factors for the Type-2 tori (i.e., more clumps and wider torus angular distributions). This would imply that the observed differences between Type-1 and Type-2 AGNs would not be due to orientation effects only, but due to intrinsic differences in their tori. However, due to the limited size of the sample analyzed by RA09, and in particular of Type-1 Seyferts compared with Type-2s, our aim is to enlarge the sample studied in the previous work with new Seyfert 1 IR data to compare the properties of Type-1 and Type-2 Seyfert tori.

In this work, we report new subarcsecond MIR imaging data for the three nearby Type-1 Seyfert galaxies NGC 7469, NGC 6221, and NGC 6814, for which we estimate unresolved nuclear MIR fluxes. We enlarge the sample by including the galaxies NGC 1097, NGC 1566, NGC 3227, and NGC 4151, which have similar MIR data, and we compile NIR nuclear fluxes from the literature of similar resolution to construct nuclear SEDs for all the galaxies. We fit these SEDs with clumpy torus models which we interpolate from the Nenkova et al. (2008a, 2008b) database, and compare them with the larger sample studied by RA09. Table 1 summarizes key observational properties of the sources in the sample. Section 2 describes the observations, data reduction, and compilation of NIR and MIR fluxes. Sections 3 and 4 present the main observational results, and in Section 5 we report the modeling results. We discuss differences between Type-1 and Type-2 Seyferts and draw conclusions about the clumpy torus models and AGN obscuration in general in Section 6. Finally, Section 7 summarizes the main conclusions of this work.

Table 1. Basic Galaxy Data

Galaxy Seyfert Type Reference z Distance Scale Reference
        (Mpc) (pc arcsec−1)  
NGC 1097 Sy1a 1 0.0042 19 92 8
NGC 1566 Sy1 2 0.0050 20 97 9
NGC 6221 Sy1 3 0.0050 18 87 10
NGC 6814 HII/Sy1.5 4 0.0052 21 102 11
NGC 7469 Sy1 5 0.0163 65 315 12
NGC 3227 Sy1.5 6 0.0039 17 82 13
NGC 4151 Sy1.5 7 0.0033 13 64 14

Notes. Classification and distance are taken from the literature (references below) and spectroscopic redshift from the NASA/IPAC Extragalactic Database (NED). aOriginally classified as a LINER by Keel (1983) and Phillips et al. (1984). References. (1) Storchi-Bergmann et al. 1997; (2) Kriss et al. 1990; (3) Levenson et al. 2001; (4) Véron-Cetty & Véron 2006; (5) Osterbrock & Martel 1993; (6) Rubin & Ford 1968; (7) Ayani & Maehara 1991; (8) Willick et al. 1997; (9) Sandage & Bedke 1994; (10) Koribalski & Dickey 2004; (11) Liszt & Dickey 1995; (12) Heckman et al. 1986; (13) Garcia 1993; (14) Radomski et al. 2003.

Download table as:  ASCIITypeset image

2. OBSERVATIONS AND DATA REDUCTION

2.1. MIR Imaging Observations

In order to enlarge the sample of 18 Seyfert galaxies presented in RA09, which comprises 12 Seyfert 2 (Sy2), 2 Seyfert 1.9 (Sy1.9), 1 Seyfert 1.8 (Sy1.8), 2 Seyfert 1.5 (Sy1.5), and 1 Seyfert 1 galaxy (Sy1), we obtained new subarcsecond MIR observations of the Type-1 Seyferts NGC 6221, NGC 6814, and NGC 7469 (see Table 1).

The observations were performed with the MIR camera/spectrograph Thermal-Region Camera Spectrograph (T-ReCS; Telesco et al. 1998) on the Gemini-South telescope during the Summer of 2009. T-ReCS uses a Raytheon 320 × 240 pixel Si:As IBC array, providing a plate scale of 0farcs089 pixel−1, corresponding to a field of view of 28farcs5 × 21farcs4. The filters employed for the observations were the narrow Si-2 filter (λc = 8.74 μm, Δλ = 0.78 μm, 50% cut-on/off) and the broad Qa filter (λc = 18.3 μm, Δλ = 1.5 μm, 50% cut-on/off). The resolutions obtained were 0farcs3 at 8.7 μm and 0farcs5 at 18.3 μm, as measured from the observed point-spread function (PSF) stars. A summary of the observations is reported in Table 2.

Table 2. Summary of MIR Observations

Galaxy Filters Observation On-source PSF FWHM
    Epoch Time (s)   
      N Band Q Band N Band Q Band
NGC 1097 Si-5, Qa 2005 Sep 456 912 0farcs41 0farcs52
NGC 1566 Si-2, Qa 2005 Sep 152 304 0farcs30 0farcs53
NGC 6221 Si-2, Qa 2009 Aug 145 203 0farcs32 0farcs55
NGC 6814 Si-2, Qa 2009 Aug 145 203 0farcs28 0farcs53
NGC 7469 Si-2, Qa 2009 Sep 145 203 0farcs31 0farcs55
NGC 3227 N' 2006 Apr 300 ... 0farcs39 ...
NGC 4151 N, IHW18 2001 May 360 480 0farcs53 0farcs58

Notes. Images were obtained in the 8.74 μm (Si-2, Δλ = 0.78 μm at 50% cut-on/off), 11.66 μm (Si-5, Δλ = 1.13 μm), and 18.30 μm (Qa, Δλ = 1.5 μm) T-ReCS filters. NGC 3227 was observed in the 11.29 μm (N', Δλ = 2.4 μm) Michelle/Gemini-North filter, and NGC 4151 in the 10.75 μm (N, Δλ = 5.2 μm) and 18.17 μm (IHW18, Δλ = 1.7 μm) OSCIR/Gemini-North filters.

Download table as:  ASCIITypeset image

The standard chopping–nodding technique was used to remove the time-variable sky background, the telescope thermal emission, and the so-called 1/f detector noise. The chopping throw was 15'', and the telescope was nodded 15'' in the direction of the chop every 45 s. The difference for each chopped pair for each given nodding set was calculated, and the nod sets were then differenced and combined until a single image was created. The data were reduced using in-house-developed IDL routines.

Observations of flux standard stars were made for the flux calibration of each galaxy through the same filters. The uncertainties in the flux calibration are typically ∼5%–10% at 8.7 μm and ∼15%–20% at 18.3 μm. PSF star observations were also made immediately prior to or after each galaxy observation to accurately sample the image quality. These images were employed to determine the unresolved (i.e., nuclear) component of each galaxy. The PSF star, scaled to the peak of the galaxy emission, represents the maximum contribution of the unresolved source (100%), where we integrate flux within an aperture of 2''. The residual of the total emission minus the scaled PSF represents the host galaxy contribution, which is analyzed in Section 3. We require a flat profile in the residual for a realistic galaxy profile over the central pixels. Therefore, we reduce the scale of the PSF from matching the peak of the galaxy emission (100%), when necessary, to obtain the unresolved fluxes reported in Table 3. They include corresponding aperture corrections to take into account possible flux losses when integrating the scaled PSF flux in the 2'' aperture.

Table 3. Unresolved MIR Fluxes

Galaxy Level of PSF Subtraction Flux Density (mJy)
  N Band Qa Band N Band Qa Band
NGC 1097 80% 20% 23 55
NGC 1566a 100% 100% 29 117
NGC 6221 80% 30% 48 314
NGC 6814 100% 40% 53 159
NGC 7469 85% 100% 174 1354
NGC 3227 100% ... 401 ...
NGC 4151 90% 100% 1320 3200

Notes. The percentages of PSF subtraction level are reported in the employed filters (listed in Table 2). Errors in flux densities are dominated in general by uncertainties in the flux calibration and PSF subtraction (∼15% at N band and ∼25% at Qa band). Fluxes include aperture corrections. aMIR fluxes for NGC 1566 are from aperture photometry using 0farcs45 aperture radii (RA09).

Download table as:  ASCIITypeset image

Figure 1 shows an example of PSF subtraction at various levels (in contours) for NGC 7469 in the Si-2 T-ReCS filter, following Radomski et al. (2002, 2003). The residual profiles from the different scales demonstrate the best-fitting result. The uncertainty in the unresolved fluxes determination from PSF subtraction is ∼10%–15%. Thus, we estimated the errors in the flux densities reported in Table 3 by adding quadratically the flux calibration and PSF subtraction uncertainties, resulting in ∼15% at 8.7 μm and ∼25% at 18.3 μm.

Figure 1.

Figure 1. 8.74 μm contour plots of NGC 7469 at the 3σ level, the PSF star, and scaled subtraction of the PSF for this galaxy at various levels (100%, 90%, and 80%). The residuals of the subtraction in the lower right panel show the host galaxy profile.

Standard image High-resolution image

2.2. Compilation of NIR and MIR High Spatial Resolution Data

In addition to the new T-ReCS observations described in Section 2.1, we include here two more Sy1 galaxies with already published T-ReCS data: NGC 1097 and NGC 1566.9 We reduced the images presented in Mason et al. (2007) for NGC 1097 and obtained unresolved MIR fluxes by performing the same technique explained in Section 2.1 (see Table 3). For NGC 1566, we compiled the MIR nuclear fluxes reported RA09. Both galaxies were observed in 2005 September with T-ReCS: NGC 1566 in the Si-2 and Qa filters, and NGC 1097 in the Si-5 (λc = 11.66 μm, Δλ = 1.13 μm, 50% cut-on/off) and Qa filters (see Table 2). The resolutions of the images are 0farcs3 at 8.7 μm, ∼0farcs4 at 11.66 μm, and 0farcs5 at 18.3 μm.

NIR subarcsecond-resolution nuclear fluxes compiled from the literature are reported in Table 4. For NGC 1097, NGC 1566, and NGC 7469, Prieto et al. (2010) reported diffraction-limited and near-diffraction-limited adaptive optics NACO/VLT fluxes. The three galaxies are unresolved in the NIR down to the highest resolution achieved (FWHM ∼ 0farcs15 for NGC 1097 in the L' band, ∼0farcs12 for NGC 1566 in the K band, and ∼0farcs08 for NGC 7469 in the H band).

Table 4. High Spatial Resolution NIR Fluxes

Galaxy Flux Density (mJy) Filters Reference(s)
  J Band H Band K Band L Band    
NGC 1097 1.1 ± 0.1 2.7 ± 0.1 3.9 ± 0.1 11 ± 1 NACO J, H, K, L' 1
NGC 1566 1.1 ± 0.1 ... 2.1 ± 0.1 7.8 ± 0.1 NACO J, K, L' 1
NGC 6221 ... 2.1 ± 0.2 ... ... F160W 2
NGC 6814 ... 5.2 ± 0.5 ... ... F160W 2
NGC 7469a 8.0 ± 0.1 15 ± 1 20 ± 1 84 ± 1 NACO J, H, K, L' 1
NGC 3227b ... 7.8 ± 0.8 16.4 ± 1.7 46.7 ± 9.3 F160W, F222M, NSFCam L 3,4
NGC 4151b 60 ± 6 100 ± 10 197 ± 20 325 ± 65 F110W, F160W, F222M, NSFCam L 3,5

Notes. Ground-based instruments and telescopes are NACO on the 8 m VLT and NSFCam on the 3 m NASA IRTF. Measurements in the F110W, F160W, and F222M filters are from NICMOS on HST. aPrieto et al. (2010) also reported a flux measurement of 19 ± 1 mJy obtained from a NICMOS/HST observation in the filter F187N. bWard et al. (1987) also reported M-band flux measurements of 72 ± 27 mJy for NGC 3227 and 449 ± 34 mJy for NGC 4151 obtained with IRCAM3 on the 3.8 m UKIRT. References. (1) Prieto et al. 2010; (2) This work; (3) Kishimoto et al. 2007; (4) Alonso-Herrero et al. 2003; (5) Ward et al. 1987.

Download table as:  ASCIITypeset image

For NGC 7469, Prieto et al. (2010) reported NACO J-, H-, and K-band nuclear fluxes obtained from observations in 2002 November, as well as in the L' and NB-4.05 μm filters, observed in this case in 2005 December. First, we discarded the narrow-band filter, since it is designed to collect the Brα emission, which in the case of this galaxy is important (as inferred from the Brγ line detected with NIR spectroscopy; Genzel et al. 1995; Hicks & Malkan 2008). The L'-band filter also contains Brα, which is very likely compromising the flux measurement, since the L' and NB-4.05 fluxes do not match the SED shape of the remaining J, H, K, Si-2, and Qa data. Indeed, Prieto et al. (2010) also reported a NICMOS/HST flux measurement in the filter F187N, obtained in 2007. This data point lies in between the NACO H and K measurements in wavelength and flux, which gives us extra-confidence in the reliability of the NACO J, H, and K measurements. Considering all the previous, we finally decided to consider the NACO L'-band flux as an upper limit.10

For NGC 6221 and NGC 6814, which do not have any published subarcsecond-resolution NIR data, we retrieved broadband images from the Hubble Legacy Archive11 obtained with NICMOS on HST. The two galaxies were observed in the F160W filter, using the NIC2 camera, as part of the program 7330 (PI: J. Mulchaey). The typical FWHM for an unresolved PSF is ∼0farcs13 using the F160W filter with NIC2. Details of the observations can be found in Scoville et al. (2000) and Regan & Mulchaey (1999). For the analysis, we first separated the nuclear emission from the underlying host galaxy emission. We then applied the two-dimensional image decomposition GALFIT program (Peng et al. 2002) to fit and subtract the unresolved component (PSF). PSF models were created using the TinyTim software package, which includes the optics of HST plus the specifics of the camera and filter system (Krist 1993). We checked that no prominent emission lines are included in the wavelength range covered by the filter F160W. The resulting unresolved NIR fluxes for NGC 6221 and NGC 6814 are reported in Table 4.

We finally include the Sy1.5 galaxies NGC 3227 and NGC 4151 in this study, which were also part of the RA09 sample. The MIR nuclear fluxes are the same reported in the latter work (see description of the observations in Section 2.1 in RA09), whereas the NIR fluxes have been updated (see Table 4).

Using the NIR nuclear fluxes reported in Table 4 in combination with our MIR unresolved measurements (Table 3) we construct AGN-dominated SEDs for the five Sy1 galaxies.

3. THE MIR EXTENDED EMISSION OF NGC 7469 AND NGC 6221

Our new MIR imaging data reveal complex extended emission for NGC 7469 and NGC 6221, which is known to be intense and associated with emitting dust heated by star formation. On the other hand, NGC 6814 lacks any extended emission. In this section, we present the MIR images of NGC 7469 and NGC 6221, and compare them with published data in different wavelength ranges.

3.1. NGC 7469

The most spectacular morphological feature of this Sy1 galaxy is a circumnuclear ring of powerful starbursts of ∼1.6 kpc diameter, which is deeply embedded in a large cloud of molecular gas and dust. This ring contains several super star clusters and regions of star formation and it accounts for two-thirds of the galaxy bolometric luminosity (Genzel et al. 1995). The star-forming ring has been the subject of study of several works in different wavelengths (see Díaz-Santos et al. 2007, hereafter DS07, and references therein), including the MIR (Miles et al. 1994; Soifer et al. 2003; Horst et al. 2009). Based on VISIR/VLT MIR observations at 12.3 μm, Horst et al. (2009) report the detection of distinct knots of star formation around the nucleus located at a distance of ∼1farcs3 (∼400 pc), although with low signal-to-noise ratio. Using NIR HST data, DS07 identified 30 clusters of star formation in the ring at 1.1 μm. By fitting the individual ultraviolet-to-NIR SEDs of the clusters, the authors reported the presence of a dominant intermediate age population (8–20 Myr) and a younger and more extinguished one (∼1–3 Myr). The latter does not coincide with the optical/NIR continuum-emitting regions, but seems to be traced by the MIR/radio emission, less affected by extinction than the optical/NIR.

Figure 2 shows the high-resolution 8.74 and 18.3 μm T-ReCS images of NGC 7469. At both wavelengths we detect extended emission with high signal-to-noise coincident with the star-forming ring. Indeed, our flux maps appear very similar to the high-resolution 11.7 μm contours presented by Miles et al. (1994) and to the high-resolution (0farcs2) VLA 8.4 GHz radio map presented in Colina et al. (2001). We identified the brightest knots in our MIR images in Figure 2 using the same notation as in Miles et al. (1994), where the AGN is labeled A. The B and C knots in our images correspond to the brightest regions in radio, according to the 5 GHz (Wilson et al. 1991) and 8.4 GHz radio maps (Colina et al. 2001). We also identified the knots D, E, and F. In Table 5, we report the star clusters identified by DS07 which better match the positions of the A to F MIR compact regions. The distances between these knots and the AGN range from 1farcs4 to 1farcs8 (median distance of ∼480 pc). All the knots appear more compact in the 18.3 μm image than in the 8.74 μm, where the ring emission is more extended. This is expected since the 8.74 μm filter contains the 8.6 μm polycyclic aromatic hydrocarbon (PAH) feature, whereas the 18.3 μm filter is mostly probing hot dust emission. As found by Spitzer, the ∼8 μm PAH emission of nearby galaxies appears to be more extended than at ∼24 μm (e.g., Helou et al. 2004; Calzetti et al. 2005).

Figure 2.

Figure 2. 8.74 μm (left) and 18.3 μm (right) T-ReCS images of NGC 7469 smoothed by using a moving box of 3 pixel size. The image size is 7.8 arcsec on each side. North is up and east is to the left. Emitting regions identified in both the Si-2 and Qa images, and also coincident with the radio emission, are labeled from A to F, where A corresponds to the AGN.

Standard image High-resolution image

Table 5. Positions and Flux Densities of NGC 7469 MIR Knots

Knot ID 8.74 μm 18.3 μm
  DS07 (X) (Y) Flux (mJy) (X) (Y) Flux (mJy)
A ... 0.00 0.00 229 0.00 0.00 1320
B C20 −1farcs58 −0farcs89 11.9 −1farcs54 −0farcs93 50
C C6 1farcs09 1farcs07 13.4 0farcs95 1farcs06 79
D C10 −0farcs49 1farcs31 11.5 −0farcs58 1farcs38 107
E C7, C15 −0farcs40 −1farcs36 15.8 −0farcs52 −1farcs49 104
F C12 1farcs25 −0farcs89  3.7 1farcs47 −0farcs83 14.8

Notes. Columns 1 and 2 indicate the label assigned to each knot in Figure 2 and in DS07. B, C, and D are also coincident with the R3, R2, and R1 regions in the VLA 8.4 GHz radio maps in Colina et al. (2001). Columns 3, 4 and 6, 7 indicate the position of each knot relative to A at 8.7 and 18.3 μm, respectively. Columns 5 and 8 list the flux densities of the knots in a 0farcs55 aperture radius. Fluxes include aperture corrections of 19% at 8.7 μm and 37% at 18.3 μm. Errors in flux densities are dominated by uncertainties in the flux calibration (∼5%–10% at 8.7 μm and ∼15%–20% at 18.3 μm).

Download table as:  ASCIITypeset image

In Table 5, we report aperture fluxes for the six identified knots in both the 8.74 and 18.3 μm images calculated using IRAF.12 The aperture radius was defined on the basis of the resolution element in the Qa band (0farcs55), and corresponding aperture corrections were applied, since the knots are only barely resolved in both bands. Positions relative to the nucleus (A) in arcseconds are also given.

3.2. NGC 6221

The galaxy NGC 6221 is a prototypical example of the so-called X-ray–loud composite galaxies (see, e.g., Moran et al. 1996). This classification comes from the comparison between its optical spectrum, which is starburst-like, and its X-ray data, where the AGN is revealed (Levenson et al. 2001). According to the X-ray data, and in particular the width of the Fe Kα line, the orientation of the Seyfert nucleus is Type-1 (Levenson et al. 2001). However, as mentioned above, the optical spectrum of the galaxy resembles more that of a starburst galaxy, implying that there must be a big amount of dust (likely associated with the starburst) along the line of sight (LOS) hiding the BLR. A nuclear optical extinction of AV = 3.0 mag was measured from the optical spectrum by Levenson et al. (2001). They also presented HST images of the central region of NGC 6221 in the optical (F606W/WFPC2) and in the NIR (F160W/NICMOS) and identified the bright central NIR source as the AGN. At optical wavelengths, the nucleus is diffuse and weaker than other bright knots, identified as star clusters.

Our 8.7 and 18.3 μm images of NGC 6221 are shown in Figure 3. They reveal spectacular extended emission with two bright knots. The AGN is centered in both images. The other bright knot, at ∼1farcs9 southwest from the nucleus, which is roughly coincident with the southwest starburst region shown in the F606W optical image in Figure 3 of Levenson et al. (2001), reaches practically the same intensity as the AGN at 8.7 μm and appears brighter at 18.3 μm. We measured aperture fluxes in a 0farcs7 radius for the southwest knot, selected to collect the bulk of its MIR emission, and obtain fluxes of 42 mJy at 8.7 μm and 422 mJy at 18.3 μm. Surrounding the AGN there is more diffuse emission extending toward the north, which in this case is more intense in the 8.7 μm image than in the 18.3 μm one. As already mentioned in Section 3.1, the ∼8 μm PAH emission of nearby galaxies appears more extended than the ∼24 μm emission (e.g., Helou et al. 2004; Calzetti et al. 2005).

Figure 3.

Figure 3. 8.74 μm (left) and 18.3 μm (right) T-ReCS images of NGC 6221 smoothed by using a moving box of 3 pixel size. The image size is 7.8 arcsec on each side. North is up and east is to the left. The AGN is centered in both images.

Standard image High-resolution image

4. SED OBSERVATIONAL PROPERTIES

4.1. Average Seyfert 1 Spectral Energy Distribution

Using the MIR and NIR data reported in Tables 3 and 4, we construct subarcsecond-resolution nuclear SEDs in the wavelength range from ∼1 to 18 μm for the five Type-1 Seyferts analyzed here. Figure 4 shows a comparison between their spectral shapes and the average Sy2 SED from RA09. This mean template was constructed using individual Sy2 data of the same angular resolution as that achieved in this work (≲0farcs55). In the same way, we have constructed an average Type-1 Seyfert template using the IR nuclear SEDs of the seven galaxies studied here.13 The spectral shape of Sy1 and Sy1.5 galaxies is practically identical (as can be seen from Figure 4). Based on the previous, and on the fact that both types of nuclei present broad lines in their optical spectra, in this work we consider them as Type-1 Seyferts.

Figure 4.

Figure 4. Observed IR SEDs for the seven Type-1 Seyfert galaxies (in color and with different symbols) used for the construction of the Sy1 average template (dashed pink). The Sy2 template from RA09 (solid black) is also plotted for comparison. The SEDs have been normalized at 8.74 μm, and the average Sy1 and Sy2 SEDs have been shifted in the Y-axis for clarity.

Standard image High-resolution image

The Sy2 template defines the wavelength grid, and we performed a quadratic interpolation of nearby measurements of the individual Sy1 galaxies onto its scale (1.265, 1.60, 2.18, 3.80, 4.80, 8.74, and 18.3 μm). We did not interpolate the sparse observations of NGC 6221 and NGC 6814, which only have NICMOS 1.6 μm data in addition to the MIR measurements. The interpolated fluxes were used solely for the purpose of deriving the average Sy1 template. The error bars correspond to the standard deviation of each averaged point, except for the 8.74 μm point (the wavelength chosen for the normalization). In this case, we assigned a 15% error, which is the nominal percentage considered for the N-band flux measurements (see Section 2).

We measured the 1.265–18.3 μm IR slope ($f_{\nu }\ \alpha \ \nu ^{-\alpha _{{\rm IR}}}$) of the Sy1 template, which is representative of the individual Sy1 SEDs: αIR = 1.7 ± 0.3. We also calculated the NIR (αNIR = 1.6 ± 0.2, from 1.265 to 8.74 μm) and MIR spectral indexes (αMIR = 2.0 ± 0.2, using the 8.7 and 18.3 μm points). A flat NIR slope indicates an important contribution of hot dust emission (up to ∼1000–1200 K; Rieke & Lebofsky 1981; Barvainis 1987) that comes from the immediate vicinity of the AGN. αIR, αNIR, and αMIR values for the individual Type-1 Seyfert galaxies and for the mean Sy1 and Sy2 SEDs are shown in Table 6. The shape of the Sy2 mean SED is very steep (αIR = 3.1 ± 0.9, αNIR = 3.6 ± 0.8, and αMIR = 2.0 ± 0.2) compared with those of the Type-1 Seyferts. In general, Sy2 have steeper 1–10 μm SEDs than Sy1 (Rieke 1978; Edelson et al. 1987; Ward et al. 1987; Fadda et al. 1998; Alonso-Herrero et al. 2001, 2003). On the contrary, the MIR slope results to be the same (αMIR = 2.0 ± 0.2) for both the Sy1 and Sy2 templates.

Table 6. Spectral Shape Information

Galaxy αIR αNIR αMIR H/N N/Q
Average Sy2 3.1 ± 0.9 3.6 ± 0.8 2.0 ± 0.2 0.003 0.23
Average Sy1 1.7 ± 0.3 1.6 ± 0.2 2.0 ± 0.2 0.071 0.22
NGC 1097 1.4 ± 0.2 1.6 ± 0.2 1.2 ± 0.1 0.117 0.42
NGC 1566 1.8 ± 0.3 1.8 ± 0.3 1.9 ± 0.2 0.040a 0.25
NGC 3227 2.0 ± 0.4 2.0 ± 0.4 ... 0.019 ...
NGC 4151 1.4 ± 0.2 1.4 ± 0.1 1.5 ± 0.1 0.076 0.41
NGC 6221 2.0 ± 0.5 1.8 ± 0.4 2.5 ± 0.3 0.044 0.15
NGC 6814 1.4 ± 0.2 1.4 ± 0.2 1.5 ± 0.1 0.098 0.33
NGC 7469 1.8 ± 0.3 1.5 ± 0.2 2.8 ± 0.4 0.086 0.13

Notes. Columns 2, 3, and 4 show the fitted spectral indexes (fναν−α) measured from the average Sy2 and Sy1 templates and for the individual Sy1 SEDs in the whole range (αIR, from ∼1 to 18 μm), in the NIR (αNIR, from ∼1 to ∼9 μm), and in the MIR (αMIR, using the N- and Q-band data points), respectively. Columns 5 and 6 give the values of the H/N and N/Q band ratios. aDue to the lack of H-band data for NGC 1566, we used the interpolated value at 1.6 μm for calculating the H/N ratio.

Download table as:  ASCIITypeset image

Alonso-Herrero et al. (2003) also reported IR spectral indices measured from 1 to 16 μm for Sy1 and Sy1.5 galaxies (αAlonsoIR = 1.5–1.6). On the other hand, the NIR slopes of the Sy1.8 and Sy1.9 in RA09 have intermediate values between those of Sy2 and Sy1 (mean slope αIR = 2.0 ± 0.4), also in agreement with the IR slopes reported in Alonso-Herrero et al. (2003) for Sy1.8 and Sy1.9 (αAlonsoIR = 1.8–2.6). In summary, the slope of the IR nuclear SED is generally correlated with Seyfert type: Type-2 nuclei show steeper SEDs, whereas Type-1 and intermediate Seyferts are flatter. The NIR excess responsible for flattening the SED of the Type-1 nuclei would come from the contribution of hot dust in the directly illuminated faces of the clouds in the torus, as well as from the direct AGN emission (i.e., the tail of the optical power-law continuum).

A similar comparison between Type-1 and Type-2 spectral shapes can be done by using the H/N and N/Q flux ratios. The H/N ratio is larger for Type-1 Seyferts (0.07 ± 0.03) than for Sy2 (0.003 ± 0.002), as measured from the individual values of the Sy1 considered here and the Sy2 in RA09. The difference in the H/N ratio between Sy1 and Sy2 galaxies is significantly different at the 100% confidence level, according to the Kolmogorov–Smirnov (K-S) test. This ratio depends on both the torus inclination and covering factor, as we will discuss in Section 6.1. On the other hand, N/Q is very similar for Type-1 and Type-2 Seyferts (mean values of 0.27 ± 0.11 and 0.23 ± 0.14, respectively). Values of H/N and N/Q for the individual Sy1 galaxies and the average templates are reported in Table 6.

5. SED MODELING

5.1. Clumpy Dusty Torus Models and Bayesian Approach

The clumpy dusty torus models of Nenkova et al. (2002) hold that the dust surrounding the central engine of an AGN is distributed in clumps, instead of homogeneously filling the torus volume. These clumps are distributed with a radial extent Y = Ro/Rd, where Ro and Rd are the outer and inner radius of the toroidal distribution, respectively (see Figure 5). The inner radius is defined by the dust sublimation temperature (Tsub ≈ 1500 K), with Rd = 0.4(1500KT−1sub)2.6(L/1045 erg s-1)0.5 pc. Within this geometry, each clump has the same optical depth (τV, defined at the V band). The average number of clouds along a radial equatorial ray is N0. The radial density profile is a power law (∝rq). A width parameter, σ, characterizes the angular distribution of the clouds, which has a smooth edge. The number of clouds along the LOS at an inclination angle i is $N_{{\rm LOS}}(i) = N_0 e^{(-(i-90)^2/\sigma ^2)}$. Finally, the optical extinction produced by the torus along the LOS is computed as $A_{V}^{{\rm LOS}} = 1.086 N_0 \tau _{V} e^{(-(i-90)^{2}/\sigma ^{2})}$ mag. For a detailed description of the clumpy models, see Nenkova et al. (2002, 2008a, 2008b).

Figure 5.

Figure 5. Scheme of the clumpy torus described in Nenkova et al. (2008a, 2008b). The radial extent of the torus is defined by the outer radius (Ro) and the dust sublimation radius (Rd). All the clouds are supposed to have the same τV, and σ characterizes the width of the angular distribution. The number of cloud encounters is a function of the viewing angle, i.

Standard image High-resolution image

The clumpy database now contains 1.2 × 106 models, calculated for a fine grid of model parameters. The inherent degeneracy between these parameters has to be taken into account when fitting the observables. To this end, we recently developed a Bayesian inference tool (BayesClumpy), that extracts as much information as possible from the observations. Details on the interpolation methods and algorithms employed can be found in Asensio Ramos & Ramos Almeida (2009). Thus, for the following analysis of the Seyfert SEDs, we are not using the original set of models described in Nenkova et al. (2008a, 2008b), but an interpolated version of them (see Figures 3 and 4 in Asensio Ramos & Ramos Almeida 2009 for a comparison between the original and interpolated models). In this work, we use the most up-to-date version of the models that are corrected for a mistake in the torus emission calculations, which in principle only affects the AGN scaling factor (see erratum by Nenkova et al. 2010). The present version of BayesClumpy has also been updated to use the Multinest algorithm of Feroz & Hobson (2008) and Feroz et al. (2009) for sampling. This algorithm is very robust and efficient when sampling from complex posterior distributions.

The prior distributions for the model parameters are assumed to be truncated uniform distributions in the intervals reported in Table 7. Therefore, we give the same weight to all the values in each interval. Apart from the six parameters that characterize the models, there is an additional parameter that accounts for the vertical displacement required to match the fluxes of a chosen model to an observed SED, which we allow to vary freely. This vertical shift scales with the AGN bolometric luminosity (see Section 6). In order to compare with the observations, BayesClumpy simulates the effect of the employed filters on the simulated SED by integrating the product of the synthetic SED and the filter transmission curve. Observational errors are assumed to be Gaussian or upper/lower limit detections. A detailed description of the Bayesian inference applied to the clumpy models can be found in Asensio Ramos & Ramos Almeida (2009). In addition, to see an example of the use of clumpy model fitting to IR SEDs using BayesClumpy, see RA09.

Table 7. Clumpy Model Parameters and Considered Intervals

Parameter Abbreviation Interval
Width of the angular distribution of clouds σ [15°, 70°]
Radial extent of the torus Y [5, 30]
Number of clouds along the radial equatorial direction N0 [1, 15]
Power-law index of the radial density profile q [0, 3]
Inclination angle of the torus i [0°, 90°]
Optical depth per single cloud τV [5, 150]

Download table as:  ASCIITypeset image

5.2. Model Results

5.2.1. Seyfert 1 Individual Fits

The results of the fitting process of the IR SEDs with the interpolated version of the clumpy models of Nenkova et al. (2008a, 2008b) are the posterior distributions for the six free parameters that describe the models and the vertical shift. These are indeed the probability distributions of each parameter, represented as histograms. When the observed data introduce sufficient information into the fit, the resulting posteriors will clearly differ from the input uniform priors, either showing trends or being centered at certain values within the intervals considered. For all the Sy1 fits, we considered uniform priors in the intervals shown in Table 7. The only exception is NGC 7469, for which we use a Gaussian prior of 85° ± 2° for the inclination angle of the torus, based on the value of the accretion disk viewing angle deduced from X-ray observations (Nandra et al. 2007), assuming that the disk and the torus are coplanar.

We fit the individual Sy1 SEDs with BayesClumpy, modeling the torus emission and the direct AGN contribution (the latter as a broken power law). We also consider the IR extinction curve of Chiar & Tielens (2006) to take into account any possible foreground extinction from the host galaxy. The AGN scales self-consistently with the torus flux, and the foreground extinction (separate from the clumpy torus) is another free parameter, which we set as a uniform prior ranging from AV = 0 to 10 mag.14

Table 8. Parameters Derived from the Clumpy Model Fitting

Galaxy σ (°) Y N0 q i (°) τV ALOSV (mag)
  Median Mode Median Mode Median Mode Median Mode Median Mode Median Mode Median Mode
NGC 1097 29±108 28 19±68 23 <2 1 2.7±0.20.3 2.9 36±1719 43 42±6910 36 <15 1
NGC 1566 46±1215 52 21 ± 4 20 2 ± 1 2 0.2 ± 0.2 0.1 79±715 86 104±2855 141 165±95100 160
NGC 6221 42±1512 36 21±56 24 43 7 0.8±0.70.5 0.4 59±1833 79 110±2332 128 <690 1
NGC 6814 23±135 20 18 ± 7 23 2 ± 1 2 1.1±0.60.5 1.1 43±2725 39 113±2230 139 <85 1
NGC 7469 41±176 38 22±56 27 3 ± 1 3 1.9 ± 0.4 1.9 84 ± 2 (fix) 84 141±615 148 500 ± 120 430
NGC 3227 36±1812 33 19 ± 6 19 43 2 0.6±0.50.4 0.5 49±2126 66 118±1827 129 <240 1
NGC 4151 24±176 19 16±87 8 21 2 1.8 ± 0.6 1.8 43±1826 57 110±2326 123 <60 1

Notes. Medians and modes of the Sy1 probability distributions. Those presenting single tails have been characterized with the mode and upper/lower limit at 68% confidence. The models include the intrinsic AGN continuum emission.

Download table as:  ASCIITypeset image

In addition to the Gemini MIR unresolved fluxes reported in Table 3, we consider MIR nuclear fluxes from VISIR compiled from the literature when available (see Table A1 in Appendix A). We find good agreement between the VISIR and T-ReCS unresolved fluxes. We did not include measurements in the VISIR PAH filters (8.59, 11.25, and 11.88 μm) for the galaxies NGC 7469 and NGC 1097 because of their intense star formation and their already well-sampled SED.

Although the solutions to the Bayesian inference problem are the probability distributions of each parameter, we can translate the results into corresponding spectra (Figure 6). The solid lines correspond to the model described by the combination of parameters that maximizes their probability distributions (maximum-a-posteriori, MAP). Dashed lines represent the model computed with the median value of the probability distribution of each parameter. Shaded regions indicate the range of models compatible with the 68% confidence interval for each parameter around the median. In Figure 7, we show the posteriors of the six torus parameters (the vertical shift and the foreground extinction have been marginalized) for the galaxy NGC 1097. Those for the rest of the Sy1 galaxies are presented in Appendix A (Figures A1A6).

Figure 6.

Figure 6. High spatial resolution IR SEDs of the Sy1 galaxies NGC 1097, NGC 1566, NGC 6221, NGC 6814, NGC 7469, NGC 3227, and NGC 4151. Solid and dashed lines correspond to the MAP and median models, respectively. Shaded regions indicate the range of models compatible with the 68% confidence interval for each parameter around the median.

Standard image High-resolution image
Figure 7.

Figure 7. Probability distributions resulting from the fit of NGC 1097. Solid and dashed lines represent the mode and the median of each distribution, and dotted lines indicate the 68% confidence level around the median. The histograms have been smoothed for presentation purposes, occasionally leading to small offsets between the solid line corresponding to the mode and the position of the maximum of the histograms.

Standard image High-resolution image

The more information the IR SEDs provide, the better the probability distributions are constrained. From the analysis performed here and in RA09 it appears important to sample the SED in the wavelength range around 3–4 μm and also at ∼18 μm to constrain the model parameters. A detailed study of the influence of different filters/wavelengths in the restriction of the clumpy models parameter space will be the subject of a forthcoming paper (A. Asensio Ramos et al. 2011, in preparation). The posterior information for the seven Sy1 galaxies is summarized in Table 8.

From the individual fits of the Sy1 galaxies in our sample with clumpy torus models, we obtain the following results.

  • 1.  
    The average number of clouds along an equatorial ray is within the interval N0 = [1, 8].
  • 2.  
    Low values of σ are preferred: σ≃ [25°, 45°], and intermediate inclination angles of the torus are found: i≃ [35°, 85°].
  • 3.  
    The radial extent of the torus (Y = Ro/Rd) is weakly constrained within the interval Y≃ [15,20].
  • 4.  
    Values in the range from τV≃ [40,140] are found for the optical depth of each cloud for all the galaxies.
  • 5.  
    The radial density profile appears constrained within the interval q = [0.2,1.9], with the only exception of NGC 1097, for which q = 2.7 ± 0.20.3.
  • 6.  
    The 10 μm silicate feature appears in shallow emission or absent in the fitted models with the exception of NGC 6221 MAP model. The weak silicate feature arises in the clumpy models because both illuminated and dark cloud sides contribute to the observed spectrum (see Section 5.4 in RA09 for a detailed discussion on the silicate feature modeling).
  • 7.  
    The optical extinction produced by the torus along the LOS results in ALOSV < 690 mag for all the galaxies.
  • 8.  
    The foreground extinction from the host galaxy, which obscures the AGN direct emission in our modeling, results in AV < 5 mag (see Figure 6). The values derived are consistent with those published in the literature, e.g., the nuclear optical extinction of AV = 3 mag measured from the optical spectrum of NGC 6221 (Levenson et al. 2001), AV ∼ 1 mag determined from NACO/VLT color maps (Prieto et al. 2005) for NGC 1097, and AV = 4.5–4.9 mag reported by Mundell et al. (1995) for NGC 3227.All the above intervals or limits of the parameters correspond to median values. We chose the medians instead of the MAPs because the former gives a less biased information about the result, since it takes into account degeneracies, while the MAP does not.

The clumpy models successfully reproduce the observed Sy1 SEDs studied here with compatible results among them. This is indicating that the NIR and MIR unresolved fluxes employed here are dominated by a combination of reprocessed emission from dust in the torus and direct AGN emission.

Our modeling results for NGC 1097 somehow contradict those presented in Mason et al. (2007). The latter authors unsuccessfully tried to reproduce the T-ReCS 11.66 and 18.3 μm aperture fluxes that they measured for this galaxy using the clumpy models of Nenkova et al. (2002). The difference with our result is probably due to (1) the fact that we obtained unresolved fluxes using PSF subtraction over the same images presented in Mason et al. (2007), reducing the potential contamination from star formation; (2) they did not consider NIR data, but only the MIR fluxes; and finally (3) they faced the degeneracy problem of the clumpy models without using any sophisticated tool as, e.g., BayesClumpy.

5.2.2. Sy2 and Intermediate-type Seyfert Results

As a consequence of the publication of the erratum Nenkova et al. (2010), where the authors report a mistake in the torus emission calculations, we repeated all the fits presented in RA09 using our updated version of BayesClumpy. In order to do a proper comparison with the results for the Sy1 galaxies presented here, we performed the fits of the Sy2 and intermediate-type Seyferts considering exactly the same priors as for the Sy1. As in RA09, we did not consider the direct AGN contribution for either Sy1.8/1.9 or Sy2.

For the new fits we considered MIR nuclear fluxes from VISIR when available, in addition to the NIR and MIR fluxes reported in RA09. We did not include measurements in the VISIR PAH filters (8.59, 11.25, and 11.88 μm) for those galaxies with very intense star formation such as NGC 7582. Many of the SEDs in RA09 have been also updated with NIR data from recent publications. In Table B1 (Appendix B), we report the NIR-to-MIR SEDs for all the sample.

In general, the results from the fitting with the most up-to-date version of the interpolated clumpy models, which are reported in Table B2 in Appendix B, are compatible with those presented in RA09 at the 1σ level. Indeed, if we compare the results for the five galaxies for which we fitted exactly the same SEDs as in the previous work (Circinus, Mrk 573, NGC 1386, NGC 1808, and NGC 1365) we find that they are practically identical. The only fits that are completely different from those presented in the previous work correspond to Centaurus A and NGC 3281. This is a consequence of adding new MIR data from VISIR and/or a priori information for the inclination angle of the torus (see Appendix B).

In Table 9, we show the ranges of the parameters found for the Sy1, intermediate-type Seyferts, and Sy2. We have excluded the unreliable fits of NGC 1808 and NGC 7582 as we did in RA09. In the case of NGC 1808, due to the intense star formation that is taking place in its nuclear region and to the lower spatial resolution IR SED (all from 3–4 m telescopes), it is likely contaminated with starlight, resulting in its peculiar shape. For NGC 7582, the intense circumnuclear star formation and the edge-on orientation of the galaxy make it difficult to isolate the torus emission from that of the host galaxy. In the fit, the silicate feature is predicted in emission, while from MIR spectroscopy shows it in strong absorption (Siebenmorgen et al. 2004).

Table 9. Ranges of Parameters for Sy1, Intermediate-type Seyferts, and Sy2

Type Galaxies σ (°) Y N0 q i (°) τV
Sy1 & Sy1.5 7 [25, 45] [15, 20] [1, 8] [0.2, 1.9]a [35, 85] [40, 140]
Sy1.8 & Sy1.9 3 [35, 70] [20, 25] [1, 7] [0.9, 3.0] [25, 85] [40, 85]
Sy2 9 [20, 65] [10, 20] [6, 14] [0.0, 3.0] [45, 85] [5, 95]

Notes. General ranges of parameters resulting from the fits with the clumpy models for the seven Type-1 Seyferts, the three Sy1.8 and Sy1.9 nuclei, and for the nine Sy2 with reliable fits (the extremes of the intervals have been rounded to simplify the comparison). All the above intervals correspond to median values. aWith the exception of NGC 1097, for which q = 2.7 ± 0.20.3.

Download table as:  ASCIITypeset image

We considered the Sy1.8 and Sy1.9 types as a separate group in between the Sy1 & Sy1.5 and Sy2 because the IR slopes measured from their SEDs are intermediate between those of Sy1 and Sy2, as measured for our sample and also as reported in the literature (Section 4.1 and references therein). By looking at the ranges of parameters for the Sy1.8 and Sy1.9 types, we find similarities with the Sy1 & Sy1.5 group in Y, N0, and τV, whereas σ and q are more similar to those of Sy2.

6. COMPARISON BETWEEN TYPE-1 AND TYPE-2 SEYFERT NUCLEI

The main aim of this work is to enlarge the number of Type-1 Seyferts in the original sample of RA09 in order to better compare between Type-1 and Type-2 tori under the assumption that the SEDs studied here are torus/AGN dominated. Despite the relatively low number of objects considered (seven Type-1 and nine Type-2 Seyferts15), we find that some of the parameters are significantly different between Sy1 and Sy2.

To take full advantage of the Bayesian approach employed here for the individual fits, the best way to compare the results for Sy1 and Sy2 galaxies is to derive joint posterior distributions for the full Type-1 and Type-2 data sets, respectively. If Di contains the observed data from the ith SED, assuming that the different SEDs are statistically independent, we can use the Bayes theorem to write the posterior for all galaxies together as

Equation (1)

where $\mbox{\boldmath $\theta $} = (\sigma,Y,N_0,q,\tau _V,i)$.

Thus, we normalized all the Sy1 SEDs at 8.74 μm and fitted them together using BayesClumpy, and we did the same for the Sy2. For those galaxies without flux measurements in the Si-2 filter, we performed a quadratic interpolation of the SED and used the interpolated values at 8.74 μm to normalize the real data. We did not use the interpolated values in the fits. We considered the mean redshift for the Sy1 (z = 0.0061 ± 0.0045) and for the Sy2 (0.0078 ± 0.0051) in the fits.16 In Figure 8, we show the Sy1 (left panel) and Sy2 fits (right panel). Note that the MAP and median models predict a flat SED with the silicate feature in weak emission for the Sy1 galaxies, and steeper and with the silicate band in shallow absorption for Sy2.

Figure 8.

Figure 8. Same as in Figure 6, but for the Sy1 (left) and Sy2 SEDs (right) normalized at 8.74 μm.

Standard image High-resolution image

The comparison between the Sy1 and Sy2 posterior distributions is shown in Figure 9. From a visual inspection it is clear that the joint posteriors of the parameters N0, q, τV, and σ are completely different between Sy1 and Sy2. There is not overlap between the 1σ intervals. In Table 10, we report the median and mode values of the histograms in Figure 9.

Figure 9.

Figure 9. Same as in Figure 7, but for the joint Sy1 and Sy2 SEDs. KLD values derived from the comparison between Sy1 and Sy2 for each parameter are labeled.

Standard image High-resolution image

Table 10. Statistics of the Comparison between Sy1 and Sy2 Parameters

Type σ (°) Y N0 q i (°) τV
  Sy1 Sy2 Sy1 Sy2 Sy1 Sy2 Sy1 Sy2 Sy1 Sy2 Sy1 Sy2
Medians 44±87 63±45 21 ± 4 23±45 4 ± 1 11±21 0.8 ± 0.2 2.3 ± 0.1 47±76 54±1011 133±89 30 ± 1
Modes 42 65 21 29 4 11 0.9 2.2 46 55 132 30

Download table as:  ASCIITypeset image

In order to quantify how different the probability distributions are, we calculated the Kullback–Leibler divergence (KLD; Kullback & Leibler 1951) between the Sy1 and Sy2 posteriors. This divergence takes into account the full shape of the posterior and it is always a positive value, and it is equal to zero when two distributions are identical. Therefore, the larger the value of KLD, the more different the posteriors. We find KLD > 1 for σ (KLD = 5.2), N0 (KLD = 25), q (KLD = 1.2), and τV (KLD = 21). These are indeed the four parameters whose 1σ regions do not overlap (see Figure 9) and thus we consider their differences significant between Sy1 and Sy2.17 For both Y and i, we find KLD < 1 and similar median values between Sy1 and Sy2.

Sy1 tori are narrower and have fewer clouds ($\sigma = 44\mbox{$^\circ $}\pm ^{8\mbox{$^\circ $}}_{7\mbox{$^\circ $}}$; N0 = 4 ± 1) than those of Sy2 ($\sigma = 63\mbox{$^\circ $}\pm ^{4\mbox{$^\circ $}}_{5\mbox{$^\circ $}}$; N0 = 11 ± 21). The radial density distribution of the clouds is also different between the two Seyfert types according to this analysis: in Sy2, the majority of the clumps are distributed very close to the nucleus (i.e., steep radial density distribution; q = 2.3 ± 0.1), whereas for Sy1 the clouds distribution is flatter (q = 0.8 ± 0.2). On the other hand, the optical depth of the clouds in Sy1 tori is larger (τV = 133 ± 89) than in Sy2 (τV = 30 ± 1).

By taking a closer look to the right panel of Figure 8, which corresponds to the Sy2 fit, it is clear that some of the data points are underestimated by the model. This happens because of the Circinus SED, which has the shortest errors bars, has more weight in the fit than the rest of the SEDs. In order to check that our Sy2 results are not completely biased by Circinus, we have repeated the fit excluding it, as well as NGC 1386, which is the least restricted SED in terms of data points, getting rid of the extremes. From the new fit, we find even larger differences with the Sy1 values of σ, N0, and τV. On the other hand, the q joint posterior becomes comparable to that of the Sy1, probably as a consequence of getting rid of the two SEDs fitted with the largest q values among the Sy2. Considering all the previous, we prefer to be cautious about the q parameter, and only consider σ, N0, and τV genuinely different between Sy1 and Sy2.

Interestingly, we find high as well as low values of the inclination angle of the torus for Sy1 and Sy2 (see Table 9). This variety in the i values translates into the similar median values found for the joint Sy1 and Sy2 posterior distributions ($47\mbox{$^\circ $}\pm ^{7\mbox{$^\circ $}}_{6\mbox{$^\circ $}}$ for Sy1 and $54\mbox{$^\circ $}\pm ^{10\mbox{$^\circ $}}_{11\mbox{$^\circ $}}$ for Sy2), which are also intermediate within the considered prior (i = [0°,90°]). This is telling us that, in the clumpy torus scenario, the classification of a Seyfert galaxy as a Type-1 or Type-2 depends more on the intrinsic properties of the torus rather than its inclination.

Our results contradict those presented by Hönig et al. (2010), who find similar averaged values of N0 for both the Sy1 and Sy2 galaxies in their sample. However, it is worth mentioning that they fixed low values of the inclination angle of the torus for Sy1 and large for Sy2, which could likely have influenced their results.

In Figure 10, we represent the median values of σ and N0 for the different Seyfert types over the covering factor contours.18 The covering factor is defined as $C_T=1-\int e^{-N_{{\rm LOS}}(i)}d \cos (i)$. Type-1 nuclei tend to be located within lower CT contours (CT ⩽ 0.6) than those of Type-2s, for which CT ⩾ 0.5, with the exception of Centaurus A and Mrk 573.19 We have represented with larger symbols the median values from the joint σ and N0 posteriors reported in Table 10.

Figure 10.

Figure 10. σ vs. N0 for the individual galaxies. Either median values or upper/lower limits are taken from the fits presented here. Dots correspond to Type-1 Seyferts, triangles to Sy1.8 and Sy1.9, and squares to Sy2. Error bars indicate 68% confidence level around the median. Note the segregation between Seyfert types, indicating the intrinsic difference between their tori in terms of covering factor (indicated in contours). The big dot and square correspond to the average σ and N0 values for Sy1 and Sy2 from Table 10.

Standard image High-resolution image

Since the covering factor is a nonlinear function of the torus parameters, we took full advantage of our Bayesian approach and generated joint posterior distributions for CT from those in Figure 9, which are shown in the left panel of Figure 11. The median values of the histograms are CT(Sy2) = 0.95 ± 0.02 and CT(Sy1) = 0.5 ± 0.1. The divergence between the Sy1 and Sy2 CT posteriors is KLD = 28, indicating that the difference is significant (the 1σ regions do not overlap). Thus, Sy1 tori in our sample have lower CTs than those of Sy2, implying that they are intrinsically different. This difference was theoretically predicted by Lawrence & Elvis (2010) using tilted disk models for the torus.

Figure 11.

Figure 11. Joint posterior distributions of the torus covering factor (left panels) and escape probability (right panels) for Sy1 (top) and Sy2 galaxies (bottom). The values of the Kullback–Leibler divergence obtained from the comparison between Sy1 and Sy2 are KLD = 28 for CT and KLD = 29 for Pesc.

Standard image High-resolution image

6.1. Near-infrared Emission and Torus Angular Width

As reported in Section 4.1, Type-1 Seyferts present characteristically higher H/N ratios and flatter SEDs than those of Sy2 nuclei. The enhancement on the NIR emission of Type-1 AGN is produced by the hot dust from the directly illuminated faces of the clumps in the torus which are close to the central engine, and also by direct AGN emission (i.e., the tail of the optical/ultraviolet power-law continuum), which strongly flattens their IR SEDs (Rieke & Lebofsky 1981; Barvainis 1987).

Based on the IR SEDs presented here, the relative NIR contribution to the SED is generally correlated with the Seyfert type. Sy2 galaxies show lower NIR to MIR ratios (H/N = 0.003 ± 0.002) than Sy1 (0.06 ± 0.03).

In the context of the clumpy models, the presence of a cloud along the LOS, which may occur from any viewing angle, results in a Type-2 classification. Cloud encounters are more probable at large inclination angles (i), but there is always a finite probability of having an unobscured view of the AGN. In fact, the likelihood of intercepting a cloud along a LOS depends on the combination of i, N0, and σ. Thus, the preference for lower values of these parameters (especially N0 and σ) in Type-1 Seyferts increases the likelihood of unimpeded views of some directly illuminated cloud faces (i.e., those on the "back" side of the torus) and direct AGN emission, resulting in an increase of the NIR flux. The latter can be represented in terms of the escape probability Pesc (see Equation (4) in Nenkova et al. 2008a). For a total number of clouds NLOS along a path, Pesc ≃ exp(−NLOS) when the clouds are optically thick (τλ>1).

In Figure 12, we show the dependence of the H/N ratio on the escape probability. All the Sy2 are in the bottom left corner of the diagram, whereas the Sy1 have higher values of the H/N ratio and Pesc = [1%, 92%]. Sy1.8 and Sy1.9 galaxies present intermediate values between those of Sy1 and Sy2. The derived joint posterior distributions of the escape probabilities for Sy1 and Sy2 (KLD = 29 between them) are shown in Figure 11, and the median values of the histograms are Pesc(Sy1) = 18% ± 3% and Pesc(Sy2) = 0.05% ± 0.080.03%.

Figure 12.

Figure 12. H/N vs. Pesc. Lower values of N0, σ, and i result in higher Pesc, resulting in higher NIR-to-MIR ratios. The inset shows an amplification of the region occupied by the Sy2. Symbols are the same as in Figure 10. The big dot and square correspond to the median values of Pesc for Sy1 and Sy2 from Figure 11, and to the H/N values of the Sy1 and Sy2 average templates (Table 6).

Standard image High-resolution image

Thus, while for tori with high values of i, N0, and σ the probability of having a direct view of the AGN is very little, which increases for objects with narrower and less inclined tori, and containing less clumps. In this work, we show for the first time that, in the clumpy torus scenario, the classification as a Type-1 or Type-2 Seyfert depends more on the intrinsic properties of the torus, rather than on the inclination angle itself. The Sy1 galaxies in our sample have larger Pesc than the Type-2 Seyferts, and as a consequence of that, we detect the broad lines in their spectra.

6.2. AGN Luminosities

The clumpy model fits yield the intrinsic bolometric luminosity of AGN (LAGNbol) by means of the vertical shift applied to match the observational data points. Combining this value with the torus luminosity (Ltorbol), obtained by integrating the corresponding model torus emission (without the AGN contribution), we derive the reprocessing efficiency (RE) of the torus (Ltorbol/LAGNbol). The previous values are calculated on the Bayesian framework by combining the posterior distributions of the model parameters. Median values and 1σ intervals for Sy1 galaxies are reported in Table 11, and those for Sy2 and intermediate-type Seyferts are shown in Table B3 (Appendix B).

Figure 13.

Figure 13. Reprocessing efficiency vs. torus covering factor for the individual galaxies. Higher values of CT translate into more efficient reprocessors. In general, Type-2 tori are more efficient than Type-1 tori, with the exceptions of Centaurus A and Mrk 573. Symbols are the same as in Figure 10.

Standard image High-resolution image

Table 11. Bolometric Luminosity Predictions

Galaxy LAGNbol Ltorbol/LAGNbol Ro (pc) LAGNXbol LAGNXbol / LAGNbol Reference
NGC 1097 2.4±0.80.4 × 1042 0.4 ± 0.1 0.4±0.10.2 1.0 × 1042 0.4 1
NGC 1566 5.9±2.21.8 × 1042 0.4 ± 0.1 0.6 ± 0.2 6.3 × 1042 1.1 2
NGC 6221 6.2±4.33.0 × 1042 0.7±0.50.3 0.7±0.30.2 1.3 × 1043 2.1 3
NGC 6814 8.1±6.02.4 × 1042 0.4 ± 0.2 0.7 ± 0.3 2.7 × 1043 3.3 4
NGC 7469 3.7 ± 0.8 × 1044 0.5 ± 0.1 12 3.4 × 1044 0.9 5
NGC 3227 1.3±1.30.5 × 1043 0.7±0.50.3 0.9±0.40.3 3.8 × 1043 2.9 6
NGC 4151 4.6±1.50.9 × 1043 0.5 ± 0.1 1.4±0.80.6 1.7 × 1044 3.7 7

Notes. Bolometric luminosities corresponding to the AGN luminosity (LAGNbol in erg s−1). Columns 3 and 4 correspond to the RE (Ltorbol/LAGNbol) and the outer radius of the torus calculated using LAGNbol and Y. Absorption-corrected 2–10 keV X-ray luminosities are taken from the literature (references below). LAGNXbol (erg s−1) is derived from 20×LAGNX. References. (1) Terashima et al. 2002; (2) Levenson et al. 2009; (3) Levenson et al. 2001; (4) Gandhi et al. 2009; (5) Nandra et al. 2007; (6) Lamer et al. 2003; (7) Beckmann et al. 2005.

Download table as:  ASCIITypeset image

Sy2 tori in our sample are more efficient reprocessors than Sy1, absorbing and re-emitting the majority of the intrinsic AGN luminosity in the IR: RE(Sy2) = [0.4, 1.0], with a median value of 0.8 and RE(Sy1) = [0.2,0.7], with median of 0.5. It makes no sense to compare the derived joint Sy1 and Sy2 posterior distributions of RE, because we normalized the SEDs to perform the fits, and consequently the derived LAGNbol are meaningless.

We considered a possible dependency of the RE (or alternatively the covering factor, CT) on LAGNbol, since the amount of incoming radiation from the AGN could possibly have some influence on the reprocessed energy or even in the torus properties (e.g., receding torus scenario; Lawrence 1991). However, we find no relationship between the two quantities in the luminosity range considered. This means that the RE depends primarily on the total number of clouds available to absorb the incident radiation, i.e., on the torus covering factor. However, the possible dependence of the torus properties on the AGN luminosity considering a broader luminosity range is further investigated in Alonso-Herrero et al. (2011).

Figure 13 shows that there is a correlation between RE and the torus covering factor for the galaxies in our sample. The larger CT the more efficient reprocessor. Type-2 tori have in general larger RE than those of Type-1, with the exceptions of Centaurus A and Mrk 573. Despite the relatively small sample, Figure 13 shows a segregation between Sy1 and Sy2 galaxies in terms of RE and covering factor.

If all Seyfert nuclei are identical, as the unified model predicts, only the viewing angle should determine the classification, not the properties of the torus itself. While our results are limited by the small sample analyzed, they suggest instead that the Type-1/Type-2 classification depends on the torus intrinsic properties rather than in the mere torus inclination.

The bolometric luminosity of the intrinsic AGN derived from the fits (LAGNbol; column 2 in Table 11) can be directly compared with those from the absorption-corrected 2–10 keV luminosities compiled from the literature (LAGNXbol; Column 6 in Table 11), which is an effective proxy for the AGN bolometric luminosity (Elvis et al. 1994). To obtain LAGNXbol from the intrinsic 2–10 keV values, we applied a bolometric correction factor of 20 (Elvis et al. 1994).

We find similar values of LAGNXbol and LAGNbol for all the Sy1 (see Table 11). In the case of NGC 1097, we expect to have some contamination from the nuclear starburst (see Section 5.2.1 and Mason et al. 2007). Thus, if the LAGNXbol value represents the AGN contribution only, LAGNbol may be overestimated because of the starburst contribution, thus resulting in a low ratio between the two measurements. The good agreement with the observational X-ray measurements reinforces the results from our SED modeling. The same comparison with X-ray data for Sy2 and intermediate-type Seyferts is shown in Table B3 in Appendix B.

6.3. Torus Size

In general, the IR SED fitting does not constrain the size of the torus (Y) as well as other model parameters (see Section 5.3 in RA09). The NIR and MIR observations are sensitive to the warm dust (located within ∼10 pc of the nucleus), which depends on the combination of model parameters N0, q, and Y (Thompson et al. 2009). FIR observations would be more sensitive to the torus extent independently. However, the fits performed here for Type-1 and Type-2 Seyferts are consistent with a small torus size, confined to scales of less than 6 pc (see below).

Uniform density models require the dusty torus to extend over large dimensions, to provide cool dust that produces the IR emission (e.g., Granato & Danese 1994). In contrast, in a clumpy distribution, different dust temperatures can coexist at the same distance, including cool dust at small radii (Nenkova et al. 2002), so large tori, which are inconsistent with imaging and interferometry results, are not necessary. Indeed, for the Seyfert galaxies considered here, we found Y ranging from 10 to 25 (see Table 8), showing that small tori can account for the observed IR nuclear emission.

The outer size of the torus scales with the AGN bolometric luminosity: Ro = YRd, so assuming a dust sublimation temperature of 1500 K, Ro = 0.4Y(LAGNbol/1045)0.5 pc. We derived Ro posterior distributions from those of LAGNbol and Y and find that all tori in our sample have outer radii smaller than 6 pc (Tables 11 and B3), in agreement with MIR direct imaging of nearby Seyferts (Packham et al. 2005; Radomski et al. 2008) and also interferometric observations (Jaffe et al. 2004; Tristram et al. 2007; Meisenheimer et al. 2007; Raban et al. 2009).

For example, the estimated outer radius for NGC 1097 (Ro = 0.4 ± 0.10.2 pc) is in agreement with the value derived from NIR NACO/VLT observations (Prieto et al. 2005), which placed the radius of the central compact source in r < 5 pc. Mason et al. (2007) also derived an upper limit of 19 pc for the size of the unresolved component from the MIR images employed in this work for NGC 1097. For the case of NGC 7469, for which we derive Ro = 5 ± 12 pc, Tristram et al. (2009) reported an estimation of the size of the dust distribution of 10 pc, based on MIDI/VLT interferometric observations, although compromised by the high level of noise and the lack of fringes.

7. CONCLUSIONS

We present new subarcsecond-resolution MIR imaging data at 8.7 and 18.3 μm for the three Type-1 Seyfert galaxies NGC 6221, NGC 6814, and NGC 7469. NGC 7469 and NGC 6221 appear extended, with part of this extended emission associated with emitting-dust heated by star formation. On the contrary, NGC 6814 lacks of any extended emission. Nuclear MIR and NIR fluxes for the three galaxies as well as for NGC 1566, NGC 1097, NGC 3227, and NGC 4151 are reported. We construct nuclear SEDs that the AGN dominates and fit them with clumpy torus models and a Bayesian approach to derive torus parameters. The main results of this work for the individual Sy1 galaxies and from the comparison with the Sy2 and intermediate-type Seyferts in RA09 are summarized as follows.

  • 1.  
    We derived an average Type-1 Seyfert template from the individual Sy1 SEDs, which is flatter (mean IR slope αIR = 1.7 ± 0.3) than the Type-2 mean SED presented in RA09 (αIR = 3.1 ± 0.9).
  • 2.  
    The NIR-to-MIR flux ratios measured from the individual SEDs are larger for Sy1 (0.07 ± 0.03) than for Sy2 (0.003 ± 0.002). Indeed, the distributions of NIR-to-MIR values are significantly different between the two types at the 100% confidence level.
  • 3.  
    The interpolated version of the clumpy models of Nenkova et al. (2008a, 2008b) employed here successfully reproduces the nuclear IR SEDs of Type-1 Seyferts with compatible results among them. Consequently, the observed nuclear IR emission of these galaxies can be accounted for by dust heated by the central engine and direct AGN emission.
  • 4.  
    We derive joint posterior distributions for Sy1 and Sy2 and find that the differences in N0, τV, and σ between Type-1 and Type-2 tori are significant according to the Kullback–Leibler divergence and the lack of overlap between their 1σ confidence intervals.
  • 5.  
    We find that Sy1 tori are narrower and have fewer clouds than those of Sy2. In addition, the optical depth of the clouds in Sy1 tori is larger than in Sy2.
  • 6.  
    There is not a clear trend in the values of the inclination angle of the torus for Sy1 and Sy2 (slightly larger values are found for Sy2).
  • 7.  
    The larger the covering factor of the torus, the smaller the likelihood of intercepting a cloud along a LOS. In our sample, Seyfert 2 tori have larger covering factors and smaller escape probabilities than those of Seyfert 1.
  • 8.  
    Despite the limited number of galaxies considered, we find that Type-2 tori are in general more efficient reprocessors than those of Type-1. Indeed, there is a correlation between the RE and the torus covering factor.
  • 9.  
    For the Seyfert galaxies studied here, we find that tori with outer radii smaller than 6 pc can account for the observed NIR/MIR nuclear emission, in agreement with MIR interferometric observations.

Summarizing, we find tantalizing evidence, albeit for a small sample of Seyfert galaxies and under the clumpy torus hypothesis, that the classification as a Type-1 and Type-2 depends more on the intrinsic properties of the torus than on its mere inclination toward us.

C.R.A. acknowledges the Spanish Ministry of Science and Innovation (MICINN) through Consolider-Ingenio 2010 Program grant CSD2006-00070: First Science with the GTC (http://www.iac.es/consolider-ingenio-gtc/). C.R.A. acknowledges financial support from STFC PDRA (ST/G001758/1). A.A.H acknowledges support from the Spanish Plan Nacional del Espacio under grant ESP2007-65475-C02-01 and Plan Nacional de Astronomía y Astrofísica under grant AYA2009-05705-E. A.A.R. acknowledges the Spanish Ministry of Science and Innovation (MICINN) through project AYA2007-63881. A.M.P.G. acknowledges the Spanish Ministry of Science and Innovation (MICINN) through project AYA2008-06311-C02-01. C.P. acknowledges support from the NSF under grant number 0904421.

This work is based on observations obtained at the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agreement with the NSF on behalf of the Gemini partnership: the National Science Foundation (United States), the Science and Technology Facilities Council (United Kingdom), the National Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), Ministério da Ciência e Tecnologia (Brazil), and Ministerio de Ciencia, Tecnología e Innovación Productiva (Argentina). The Gemini programs under which the data were obtained are GS-2005B-DD, GS-2005B-Q-29, and GS-2009B-Q-43.

This work is based on observations made with the NASA/ESA Hubble Space Telescope, and obtained from the Hubble Legacy Archive, which is a collaboration between the Space Telescope Science Institute (STScI/NASA), the Space Telescope European Coordinating Facility (ST-ECF/ESA), and the Canadian Astronomy Data Centre (CADC/NRC/CSA).

This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.

We finally acknowledge useful comments from the anonymous referee.

APPENDIX A: FITTING RESULTS FOR Sy1 GALAXIES

Here we include the posterior distributions of the Sy1 galaxies NGC 1566, NGC 6221, NGC 6814, NGC 7469, NGC 3227, and NGC 4151 (Figures A1A6).

Figure A1.

Figure A1. Same as in Figure 7, but for the galaxy NGC 1566.

Standard image High-resolution image
Figure A2.

Figure A2. Same as in Figure 7, but for the galaxy NGC 6221.

Standard image High-resolution image
Figure A3.

Figure A3. Same as in Figure 7, but for the galaxy NGC 6814.

Standard image High-resolution image
Figure A4.

Figure A4. Same as in Figure 7, but for the galaxy NGC 7469.

Standard image High-resolution image
Figure A5.

Figure A5. Same as in Figure 7, but for the galaxy NGC 3227.

Standard image High-resolution image
Figure A6.

Figure A6. Same as in Figure 7, but for the galaxy NGC 4151.

Standard image High-resolution image

Table A1. VISIR MIR Fluxes from the Literature for the Sy1 Galaxies

Galaxy Flux Density (mJy) Filter(s) Reference(s)
NGC 1097 28 ± 7, 49 ± 12 12.27, 18.72 1, 2
NGC 1566 63 ± 9, 128 ± 32 11.88, 18.72 2
NGC 6814 99 ± 6, 96 ± 6 11.25, 13.04 3
NGC 7469 448 ± 16, 595 ± 18, 630 ± 17, 1270 ± 317 10.49, 12.27, 13.04, 18.72 1, 2, 4
NGC 3227 180 ± 11, 320 ± 22 8.99, 11.88 4

References. (1) Horst et al. 2008; (2) Reunanen et al. 2010; (3) Gandhi et al. 2009; (4) Hönig et al. 2010.

Download table as:  ASCIITypeset image

APPENDIX B: FITTING RESULTS FOR Sy2 AND INTERMEDIATE-TYPE SEYFERT GALAXIES

Here, we include the results of the fits of the intermediate-type Seyferts and the Sy2 galaxies in RA09 (Figures B1 and B2) using the data reported in Table B1 and the most up-to-date version of the Nenkova et al. (2008a, 2008b) models (see erratum Nenkova et al. 2010). The resulting model parameters for the individual galaxies are shown in Tables B2 and B3. In Figure B3, we show the posterior distributions from the fit of Centaurus A (see the electronic edition of the journal for the rest of the Sy2 and intermediate-type Seyfert posteriors).

Figure B1.

Figure B1. Same as in Figure 6, but for the Sy2 galaxies Centaurus A, Circinus, IC 5063, Mrk 573, NGC 1386, NGC 1808, NGC 3081, and NGC 3281.

Standard image High-resolution image
Figure B2.

Figure B2. Same as in Figure 6, but for the Sy2 galaxies NGC 4388, NGC 7172, and NGC 7582; the Sy1.8 NGC 1365, and the Sy1.9 galaxies NGC 2992 and NGC 5506.

Standard image High-resolution image
Figure B3.

Figure B3. 

Same as in Figure 7, but for (a) the galaxy Centaurus A, (b) the Circinus galaxy, (c) the galaxy IC 5063, (d) the galaxy Mrk 573, (e) the galaxy NGC 1386, (f) the galaxy NGC 1808, (g) the galaxy NGC 3081, (h) the galaxy NGC 3281, (i) the galaxy NGC 4388, (j) the galaxy NGC 7172, (k) the galaxy NGC 7582, (l) the galaxy NGC 1365, (m) the galaxy NGC 2992, and (n) the galaxy NGC 5506. Panels (b)–(n) are available in the online version of the journal. (An extended version of this figure is available in the online journal.)

Standard image High-resolution image

    Table B1. NIR and MIR Fluxes from the Literature for Sy2 and Intermediate-type Seyferts

    Galaxy Seyfert Flux Density (mJy) Filter(s) Reference(s)
    Centaurus A 2 1.3 ± 0.1, 4.5 ± 0.3, 34 ± 2, 200 ± 40 NACO J, H, K, L' 1
        643 ± 27, 947 ± 29, 1100 ± 100, 1451 ± 73, 2300 ± 575 VISIR 10.49,11.25,11.88,12.27,18.72 2, 3
    Circinus 2 1.6 ± 0.2, 4.8 ± 0.7, 19 ± 2, 31 ± 3, 380 ± 38, 1900 ± 190 F160W, NACO J, K, 2.42, L, M 4
    IC 5063 2 0.3 ± 0.1, 4.77 ± 0.95 F160W, F222M 5, 6
        609 ± 22, 727 ± 25, 925 ± 25, 1036 ± 56 VISIR 10.49,11.25,11.88,12.27 7
    Mrk 573 2 0.15 ± 0.06, 0.54 ± 0.04, 3.2 ± 0.6, 18.8 ± 3.8, 41.3 ± 8.3 F110W,F160W, NSFCam K', L, M 8
    NGC 1386 2 0.2 ± 0.1 F160W 5
    NGC 1808 2 15.5 ± 4.5, 30.5 ± 8.5, 27.5 ± 10.5 ISAAC J, Ks, L' 9
    NGC 3081 2 0.22 ± 0.13 F160W 5
        138 ± 11 VISIR 13.04 10
    NGC 3281 2 1.3 ± 0.2, 7.7 ± 0.8, 103 ± 9, 207 ± 25 IRAC-1 H, K, IRCAM3 L', M 11
        481 ± 24, 1016 ± 52 VISIR 11.25,13.04 10
    NGC 4388 2 0.06 ± 0.02, 0.71 ± 0.28, 40 ± 8 F110W,F160W, NSFCam L 8
        147 ± 11, 375 ± 28 VISIR 11.25,13.04 10
    NGC 7172 2 <0.4, 3.4 ± 0.7, 30 ± 6, 61 ± 12 IRCAM3 H, K, L', M 12
        165 ± 27 VISIR 12.27 2
    NGC 7582 2 11 ± 1, 18 ± 2, 96 ± 10, 110 ± 11, 142 ± 21 F160W, NACO 2.06,L,4.05, ISAAC M 3, 13
        236 ± 27, 550 ± 130 VISIR 8.99, 18.72 3, 7
    NGC 1365 1.8 8.3 ± 0.8 F160W 14
    NGC 2992 1.9 <1, 2.8 ± 0.6, 22.7 ± 4.5, 35.7 ± 7.1 IRCAM3 H, K, L', M 12
        312 ± 31 VISIR 11.25 15
    NGC 5506 1.9 13 ± 1, 53 ± 1, 80 ± 1, 290 ± 10, 351 ± 29 NACO J, H, K, L', NSFCam M 3, 16
        900 ± 100, 1400 ± 350 VISIR 11.88, 18.72 3

    Notes. Ground-based instruments and telescopes are NACO and ISAAC on the 8 m VLT, NSFCam on the 3 m NASA IRTF, IRCAM3 on the 3.8 m UKIRT, and IRAC-1 on the 2.2 m ESO telescope. Measurements in the F110W, F160W, and F222M filters are from NICMOS on HST. References. (1) Meisenheimer et al. 2007; (2) Horst et al. 2008; (3) Prieto et al. 2010; (4) Prieto et al. 2004; (5) Quillen et al. 2001; (6) Kulkarni et al. 1998; (7) Hönig et al. 2010; (8) Alonso-Herrero et al. 2003; (9) Galliano & Alloin (2008); (10) Gandhi et al. 2009; (11) Simpson 1998; (12) Alonso-Herrero et al. 2001; (13) Prieto et al. 2002; (14) Carollo et al. 2002; (15) Haas et al. 2007; (16) Ward et al. 1987.

    Download table as:  ASCIITypeset image

    Table B2. Parameters from the Fits for Sy2 and Intermediate-type Seyferts

    Galaxy Seyfert σ Y N0 q i τV ALOSV
      Type Median Mode Median Mode Median Mode Median Mode Median Mode Median Mode Median Mode
    Centaurus A 2 20 ± 2 20 13 ± 2 13 13±12 15 0.4 ± 0.2 0.5 70 ± 2 69 45±1311 42 235±11090 200
    Circinus 2 63±47 68 20±610 27 11 ± 2 11 2.6±0.20.4 2.7 85 ± 2 (fix) 85 31 ± 2 31 365±7055 340
    IC 5063 2 51 ± 10 56 12±43 10 13±12 14 0.8±0.90.6 0.3 81±58 85 90 ± 19 96 1135±315240 990
    Mrk 573 2 30±1810 23 17 ± 8 25 42 5 0.9±0.90.6 0.6 85 ± 2 (fix) 85 72±4133 53 155±10080 105
    NGC 1386 2 49±1118 52 17±87 17 11±23 13 1.5±0.80.9 2.3 85 ± 2 (fix) 85 95±5166 51 805±645440 425
    NGC 1808 2 26±127 20 18±78 18 8 ± 4 12 1.3±0.90.8 0.9 41±1624 53 111±2331 122 <100 1
    NGC 3081 2 55±817 61 18±78 22 10 ± 3 12 1.4 ± 0.9 2.3 54±2131 79 29±2312 21 <230 70
    NGC 3281 2 34 ± 3 34 10 ± 2 9 14±13 15 <0.2 0.0 63±42a 60 15 ± 2 14 120±2015 120
    NGC 4388 2 60±610 67 20±56 22 11±23 14 0.5±0.60.3 0.2 45±2522 28 31±168 25 <270 110
    NGC 7172 2 58±711 66 14±96 8 10±32 9 1.9±0.71.0 2.5 63±1626 77 21 6 50 ± 15 50
    NGC 7582 2 34±2013 23 18±78 18 42 2 2.3±0.40.6 2.6 43±1922 45 16±75 13 <15 1
    NGC 1365 1.8 35±1410 32 18 ± 7 22 7 ± 4 4 1.1±0.80.7 0.7 27±1916 19 86±3638 96 <110 1
    NGC 2992 1.9 40±1614 28 18 ± 7 24 43 5 0.9±1.00.7 0.2 75±822 78 39±1210 34 210±145105 235
    NGC 5506 1.9 >68 69 23±46 26 <2 1 >2.9 3.0 85 ± 2 (fix) 85 42 ± 3 43 60 ± 8 60

    Notes. For the galaxies Circinus, Mrk 573, NGC 1386, and NGC 5506, the i parameter has been introduced as a Gaussian prior into the computations, centered at 85° with a width of 2°, based on other observations. Probability distributions presenting a single tail have been characterized with the mode and upper/lower limits at 68% confidence. Sy1.8 and Sy1.9 have been fitted with the geometry corresponding to torus emission-only. aThe inclination angle of the torus was introduced as a uniform prior i = [60°,90°] following Storchi-Bergmann et al. (1992).

    Download table as:  ASCIITypeset image

    Table B3. Bolometric Luminosity Predictions

    Galaxy LAGNbol Ltorbol/LAGNbol Ro (pc) LAGNXbol LAGNXbol/LAGNbol Reference
    Centaurus A 4.7±0.80.7 × 1042 0.4 ± 0.1 0.4 ± 0.1 1.3 × 1043 3 1
    Circinus 1.0±0.20.1 × 1043 0.8 ± 0.1 0.8±0.30.4 1.2 × 1043 1.2 2
    IC 5063 2.7±1.20.7 × 1044 0.8±0.10.2 2.5±1.30.8 1.7 × 1044 0.6 3
    Mrk 573 1.5±1.10.6 × 1044 0.4 ± 0.2 2.8±1.41.2 4.4 × 1044 3 4
    NGC 1386 2.8±2.90.8 × 1042 0.7±0.10.3 0.4 ± 0.2 1.3 × 1043 5 5
    NGC 1808 3.1±1.40.8 × 1042 0.8±0.40.2 0.4 ± 0.2 2.2 × 1041 0.1 6
    NGC 3081 7.6±4.62.0 × 1042 0.9 ± 0.3 0.7 ± 0.3 1.0 × 1044 13 4
    NGC 3281 1.1±0.20.1 × 1044 0.7 ± 0.1 1.3 ± 0.3 3.0 × 1044 3 7
    NGC 4388 2.3±1.10.6 × 1043 1.0 ± 0.3 1.3 ± 0.4 1.5 × 1044 6 8
    NGC 7172 8.4 ± 1.6 × 1042 0.9±0.20.1 0.5±0.30.2 1.1 × 1044 13 9
    NGC 7582 1.4±0.40.2 × 1043 0.6 ± 0.1 0.8 ± 0.4 9.8 × 1043 7 10
    NGC 1365 8.1±2.91.4 × 1042 1.0±0.30.2 0.7 ± 0.3 3.0 × 1043 4 11
    NGC 2992 3.0±2.51.2 × 1043 0.6±0.30.2 1.3±0.70.6 3.0 × 1043 1.0 12
    NGC 5506 9.3±0.40.2 × 1043 0.5 ± 0.1 2.9±0.40.7 2.2 × 1044 2 13

    Note. Same as in Table 11, but for the Sy2 and intermediate-type Seyfert galaxies in RA09. References. (1) Markowitz et al. 2007; (2) Soldi et al. 2005; (3) Turner et al. 1997; (4) RA09; (5) Levenson et al. 2006; (6) Jiménez-Bailón et al. 2005; (7) Vignali & Comastri 2002; (8) Elvis et al. 2004; (9) Awaki et al. 2006; (10) Turner et al. 2000; (11) Risaliti et al. 2005; (12) Yaqoob et al. 2007; (13) Lamer et al. 2000; (14) Lamer et al. 2003; (15) Beckmann et al. 2005.

    Download table as:  ASCIITypeset image

    As mentioned in Section 5.2.2, the results presented in this appendix are compatible in general with those presented in RA09 at the 1σ level. For the majority of the galaxies, we have included IR data from recent publications. In particular, the use of VISIR MIR data have been key in predicting the silicate feature in absorption in the case of Centaurus A20 and NGC 3281, and in emission in NGC 3227.

    In the case of Centaurus A, the fit presented in RA09 predicted the silicate feature in emission, in clear contradiction with MIR spectroscopic data (Siebenmorgen et al. 2004; Meisenheimer et al. 2007). By including the VISIR data for this galaxy, the fitted models predict a broad silicate feature in agreement with the observations. In addition, we have also introduced the inclination angle of the torus as a Gaussian prior centered in 70° as inferred from the inclination of the inner radio jet of Centaurus A (Tingay et al. 1998). This value is also coincident with the 66° inclination angle of the obscuring structure derived from MIR interferometric observations (Burtscher et al. 2010).

    For the Sy2 galaxy NGC 3281, we also have considered a uniform prior in the range [60°, 90°] for the inclination angle of the torus, based on the inclination of the ionization cone axis derived in Storchi-Bergmann et al. (1992) from optical imaging and spectroscopic observations, which implies a nearly edge-on orientation of the obscuring structure.

    Footnotes

    • NGC 1097 and NGC 1566 were not included in the analysis of the SEDs presented in RA09 because of the lack of high angular resolution NIR fluxes for them at the time of publication.

    • 10 

      The nucleus of NGC 7469 had undergone different periods of activity, including a maximum in the optical happening in the period 1996–2000, followed by a relaxation epoch in the following years (Prieto et al. 2010). For this reason, here we only consider data obtained from 2002.

    • 11 
    • 12 

      IRAF is distributed by the National Optical Astronomy Observatory, which is operated by the Association of Universities for the Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation (http://iraf.noao.edu/).

    • 13 

      We have not considered the NIR ground-based data reported in Table 4 of RA09 for NGC 3227 and NGC 4151 in the construction of the mean template to be consistent with the angular resolutions of the other SEDs.

    • 14 

      Note that AV is the foreground extinction from the host galaxy, which is different from the ALOSV value reported in Table 8, corresponding to the extinction produced by the torus. $A_V^{{\rm LOS}} = 1.086 N_0 \tau _{V} e^{(-(i-90)^{2}/\sigma ^{2})}$ mag.

    • 15 

      Here, we consider the nine Sy2 in RA09 with reliable fits (NGC 1808 and NGC 7582 are excluded).

    • 16 

      Indeed, since all the galaxies are local Seyferts, the results are the same if we consider that the SEDs are rest frame.

    • 17 

      See further discussion on the Sy2 q parameter results below.

    • 18 

      A similar plot showing values for Sy2 galaxies from RA09 and PG quasars from Mor et al. (2009) was shown in the talk by M. Elitzur at the Physics of Galactic Nuclei conference held 2009 June 15–19 at Ringberg Castle. Proceedings published online at http://www.mpe.mpg.de/events/pgn09/online_proceedings.html.

    • 19 

      As discussed in RA09, the fit of Centaurus A is complicated by the presence of a dust lane of AV ∼ 7–8 mag that is likely affecting the NIR nuclear fluxes, as well as the possible synchrotron contamination of the MIR fluxes. Mrk 573 has been recently reclassified as an obscured narrow-line Seyfert 1 (NLSy1) based on NIR spectroscopy (Ramos Almeida et al. 2008, 2009a). However, here we considered it as a Sy2 because of the similarity in SED shape with the rest of the sample.

    • 20 

      In order to account for the extinction caused by the dust lane of AV ∼ 7–8 mag in Centaurus A, we have considered the Chiar & Tielens (2006) law in the fit fixing AV = 8 mag.

    Please wait… references are loading.
    10.1088/0004-637X/731/2/92