B2 0003+38A: A Classical Flat-spectrum Radio Quasar Hosted by a Rotation-dominated Galaxy with a Peculiar Massive Outflow

, , , , , and

Published 2021 June 1 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Qinyuan Zhao et al 2021 ApJ 913 111 DOI 10.3847/1538-4357/abf4de

Download Article PDF
DownloadArticle ePub

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

0004-637X/913/2/111

Abstract

We present a detailed analysis of the single-slit optical spectrum of the flat-spectrum radio quasar (FSRQ) B2 0003+38A, taken by the Echellette Spectrograph and Imager (ESI) on the Keck II telescope. This classical low-redshift FSRQ (z = 0.22911, as measured from the stellar absorption lines) remains underexplored in its emission lines, though its broadband continuum properties from radio to X-ray are well studied. After removing the unresolved quasar nucleus and the starlight from the host galaxy, we obtain a spatially resolved 2D spectrum, which clearly shows three components, indicating a rotating disk, an extended emission-line region (EELR), and an outflow. The bulk of the EELR, with a characteristic mass MEELR ∼ 107 M, and redshifted by vEELR ≈ 120 km s−1 with respect to the quasar systemic velocity, shows a one-sided structure stretching to a projected distance of rEELR ∼ 20 kpc from the nucleus. The rotation curve of the rotating disk is consistent with that of a typical galactic disk, suggesting that the FSRQ is hosted by a disk galaxy. This conclusion is in accordance with the facts that strong absorption in the H i 21 cm line was previously observed, and that Na i λλ5891, 5897 and Ca ii λλ3934, 3969 doublets are detected in the optical ESI spectrum. B2 0003+38A will become the first FSRQ discovered to be hosted by a gas-rich disk galaxy, if this is confirmed by follow-up deep imaging and/or integral field unit mapping with a high spatial resolution. These observations will also help unravel the origin of the EELR.

Export citation and abstract BibTeX RIS

1. Introduction

Blazars are a class of rare radio-loud active galactic nuclei (AGNs) (Sandage 1965), characterized by their flat radio spectra, rapid variability in multiwavelength emission, significant polarization, and bimodal synchrotron/Compton spectral energy distributions (SEDs). These characteristics are likely consequences of shocks driven by a powerful relativistic jet pointing to a direction close to our line of sight and roughly perpendicular to the accretion disk (Urry & Padovani 1995). Blazars may show strong broad emission lines (BELs) in their optical spectra, similar to typical quasars, or be featureless in their optical spectra; these blazars are called flat-spectrum radio quasars (FSRQs) and BL Lacertae (BL Lac) objects, respectively.

Observations show that the majority of blazars, regardless of their type, are hosted by giant elliptical galaxies (for BL Lac objects, e.g., Stickel et al. 1991; Kotilainen et al. 1998a, 1998b; Falomo & Kotilainen 1999; Falomo et al. 2000; Urry et al. 2000; Kotilainen et al. 2005; Hyvönen et al. 2007; León-Tavares et al. 2011; Falomo et al. 2014; for FSRQs, e.g., Olguín-Iglesias et al. 2016). Olguín-Iglesias et al. (2016) presented deep near-infrared (NIR) images of a sample of 19 (0.3 < z < 1.0) FSRQs, finding that the host galaxies of their sample were luminous and apparently follow the μe Reff relation for ellipticals and bulges, consistent with the conclusion based on BL Lac objects (Stickel et al. 1991).

As yet, whether a blazar can be hosted by a disk galaxy remains an open question. Only a handful of cases have been reported where BL Lac objects were found to be hosted by disk-dominated galaxies (e.g., Halpern et al. 1986; Abraham et al. 1991; Wurtz et al. 1996). Nilsson et al. (2003) analyzed 100 BL Lac objects from the ROSAT-Green Bank sample obtained by using the Nordic Optical Telescope, finding that all of their spatially resolved objects are better fitted by an elliptical-galaxy model (β = 0.25) than by a disk-galaxy model (β = 1.0), though with two exceptions that may be hosted by disk galaxies (Abraham et al. 1991; Wurtz et al. 1996 for 1419 + 543 and Urry et al. 1999 for 1540 + 147), whose bulk properties, however, are indistinguishable from normal elliptical galaxies. The hosts of the two BL Lac objects, 1415 + 255 and 1413 + 135, were originally identified to be disk-dominated galaxies (Halpern et al. 1986; Stocke et al. 1992; Lamer et al. 1999, respectively), but recent works, based on imaging with a higher resolution, classify 1415 + 255 as an isolated giant elliptical galaxy (Gladders et al. 1997), highlighting the importance of high-quality data. Furthermore, Scarpa et al. (2000) investigated 69 spatially resolved BL Lac object hosts and found one case of a disk (that of 0446 + 449) and two cases where disk models are preferred (1418 + 546, 0607 + 711). Despite the efforts that have been made, definitive evidence showing that a blazar can be hosted by a disk galaxy (especially for FSRQs) remains to be found. This is mostly due to the lack of high-resolution imaging or specially resolved spectra, and the brightness of blazars is so high in optical and IR bands that the hosts are outshined.

A combination of high-resolution imaging and spatially resolved spectroscopy may shed light on the morphology and dynamics of the host galaxy. Hence, we conduct a case study on a confirmed FSRQ at z = 0.229, known as B2 0003+38A (aka S4 0003+38 or J0005+3820) using the long-slit optical spectroscopy taken by the Echellette Spectrograph and Imager (ESI) at the Keck II Observatory. This paper is structured as follows: After an overview of B2 0003+38A in previous studies, we describe observations and data reduction in Section 2. In Section 3, we present the analysis of the spectral data. In Section 4, we discuss the gas kinematics in the host galaxy and the extended emission-line region (EELR). We conclude with a summary in Section 5. Throughout this paper, we adopt a cosmology with H0 = 0.7 km s−1 Mpc−1, Ωm = 0.3, ΩΛ = 0.7. The wavelengths of all the spectral lines are given in vacuum.

2. Observation and Data Reduction

2.1. The FSRQ B2 0003+38A

B2 0003+38A is classified as an FSRQ due to its BELs in the optical spectrum and the flat radio spectrum with α = −0.3 (Stickel & Kuehr 1994; Healey et al. 2007; Massaro et al. 2009). Resolved by very long baseline interferometry (VLBI) mapping at 2.29 GHz (Morabito et al. 1982), it shows a prominent core that emits more than 95% of the 2.3 GHz flux, along with a weak jet pointing to the southeast (Fey & Charlot 2000), as shown in Figure 1. Kinematic analyses show quasi-stationary knots at the jet base and relativistic motions downstream (Hervet et al. 2016). Aditya & Kanekar (2018) detected H i 21 cm absorption toward this source, finding the velocity-integrated H i 21 cm optical depths to be 1.943 ± 0.057 km s−1, and the H i column density to be 3.54 ± 0.11 × 1020 cm−2.

Figure 1.

Figure 1. The VLBI 2.292 GHz cutout centered on the B2 0003+38A (Lister et al. 2013). The orientation of the ESI/Keck slit is illustrated by the blue line. The 1farcs25 slit width is larger than the size of the cutout. The slit covers the quasar and jet.

Standard image High-resolution image

2.2. Optical Long-slit Spectrometry

The quasar B2 0003+38A was observed by Keck II ESI long-slit spectrography in its echelle mode on 2015 September 9. The spectrum was taken in the optical band (wavelength coverage λ ∼ 3995–10198 Å in the observer's frame), covering 3250–8297 Å in the rest frame for our target at z ∼ 0.2. A 1farcs25 wide slit was employed, resulting in an instrumental dispersion σ of 22 km s−1. The position angle of the slit is 41° (north to east, Figure 1). The slit width allows for both the blazar and its jet to be covered. Two exposures were taken with an integration time of 10 minutes each. Arc lamps were used during the observing campaign for wavelength calibration, and the standard star BD+28 4211 was observed about one hour before the target with identical settings for the purpose of flux calibration.

2.3. Data Reduction

We use a customized routine based on the IDL package XIDL 5 to obtain a two-dimensional (hereafter 2D) spectrum. Our data reduction consists of four steps: (1) after bias corrections, flat-fielding, and removing cosmic rays, we stack the two exposures; (2) the nucleus of the target quasar is traced along 10 echelle orders, and for each order, we resample the 2D spectrum so that a resultant pixel corresponds to 10 km s−1 in the wavelength direction and 0farcs15 in the spatial direction; (3) sky subtraction and flux calibration are applied on these resampled 2D spectra; (4) we combine these resampled 2D spectra from the 10 orders.

3. Data Analysis and Results

We conduct kinematic analyses using both the one-dimensional (hereafter 1D) and 2D spectrum to investigate the gas motion in the host galaxy. In Section 3.1, we perform a model fitting to the 1D spectra to contextualize the quasar and stellar components, and the information that the fitting delivers is discussed in Section 3.2. With the quasar and stellar contributions removed, we scrutinize the spatially resolved emission lines in the 2D spectra, and the gas kinematics are described in Section 3.3. We observe that three components exist in the gas: a rotating component, an extended emission-line component, and an outflow component. The former two components are further analyzed in Sections 3.4 and 3.5.

3.1. Spectrum of the Nucleus

We extracted the 1D spectrum from the 2D spectrum as a result of data reduction by applying an aperture with a size of 3'' × 1farcs25 centered on the quasar nucleus, as shown in Figure 2 (black lines). AGN features, including BELs, narrow emission lines (NELs) and Fe ii bumps, are seen therein. However, the existence of high-order Balmer absorption lines indicate the continuum contribution from starlight. To disentangle different contributions of the spectrum, we construct a model consisting of four components: a stellar, a power-law, a BEL, and an Fe ii component.

Figure 2.

Figure 2. The observed nuclear spectrum (black) and the best-fitted spectrum (red) are shown in regions where the NELs have little effect. The four components of the best-fitted model are shown as different colored lines: the stellar component in blue, the power-law component in green, the BEL component in cyan, and the Fe ii component in yellow. The spectrum in regions affected by the telluric absorption is plotted as a gray line.

Standard image High-resolution image

  • 1.  
    The stellar component. We assume a single simple stellar population (SSP), utilizing the library of Bruzual & Charlot (2003). When fit to the data, the template is shifted to a redshift z, broadened by convolving to a Gaussian parameterized by a velocity dispersion parameter σ, and multiplied by the extinction curve of the Small Magellanic Cloud (SMC) parameterized by E(BV) (Pei 1992).
  • 2.  
    The power-law component. We assume a formulation of fλ (λ) = C λα , where α is the exponent and C is the normalization.
  • 3.  
    The BEL component. For this component, we mainly account for the Balmer series, He i λ5876 and He ii λ4686. We assume that the BEL profiles can be represented by a linear combination of two Gaussians, whose mean and standard dispersion are fixed but the flux is allowed to vary. Furthermore, we fix the flux ratios of higher-order Balmer BELs to the Hγ BEL to those under the "Case B" situation (Storey & Hummer 1995), assuming a typical circumstance for broad-line regions with Te = 15,000 K and ne = 109 cm−3.
  • 4.  
    The Fe ii component. We employ the Fe ii template constructed by Véron-Cetty et al. (2004). After experiments with various combinations of different Fe ii lines, we conclude that only narrow Fe ii lines are necessary for fitting the observed Fe ii complex. We further assume that these different narrow Fe ii lines can be represented by single Gaussians with a fixed redshift and a fixed width. The flux ratios between different Fe ii lines are taken from the template given in Véron-Cetty et al. (2004). In addition, we assume the SMC extinction curve for the dust attenuation and reddening of the Fe ii component parameterized by E(BV)Fe II .

We then fit the spectrum of the nuclear region, where the contamination from NELs is minimal, resulting in a minimized reduced chi-square of 1.99. This fit can be further improved if the regions affected by telluric absorptions are dismissed (reduced χ2 = 1.66, Figure 2). The individual components of the best-fit spectra are depicted by colored lines in the same figure, and the best-fit parameters are tabulated in Table 1.

Table 1. Parameters for Modeling the Nuclear Spectrum

ParameterValueUnit
Stellar Component
Age450Myr
MV −22.77mag
z 0.22911 ± 0.00001
σ 150 ± 20km s−1
E(BV) 0.62 ± 0.02mag
Power-law Component
fλ5100 9.84 ± 0.0610−17 erg s−1 cm−2 Å−1
α 1.32 ± 0.02 
BEL Component
Δv1 263 ± 3km s−1
σ1 974 ± 4km s−1
Δv2 250 ± 10km s−1
σ2 2960 ± 20km s−1
frac1 0.598 ± 0.003 
fHα 4740 ± 2010−17 erg s−1 cm−2
fHβ 587 ± 310−17 erg s−1 cm−2
fHγ 103 ± 410−17 erg s−1 cm−2
fHeI5876 307 ± 310−17 erg s−1cm−2
fHeII4686 55 ± 310−17 erg s−1 cm−2
Fe ii Component
Δv 430 ± 20km s−1
σ 720 ± 20km s−1
fλ4590 1.7 ± 0.210−17 erg s−1 cm−2 Å−1
E(BV)Fe II 0.19 ± 0.04mag

Download table as:  ASCIITypeset image

3.1.1. A Narrow Absorption Line System

The nuclear spectrum of B2 0003+38A shows deep and narrow Na i λλ5890,5896 absorption lines (Figure 3), corresponding to a redshift of ∼0.22883, which is about 70 km s−1 blueshifted relative to the stellar redshift (0.22911). Unlikely to originate from the stellar populations, these are likely absorption lines from absorbers that happen to lie in our line of sight toward the quasar nucleus. Narrow Ca ii λ3934 absorption is detected at the same redshift as that of Na i absorption, implying the same absorption system, though we cannot affirm the existence of the Ca ii λ3969 absorption line due to the influence of the [Ne iii] λ3869 and Hepsilon emission lines. Assuming that the starlight is not affected by the absorber mentioned above, we consider the case of partial coverage:

Equation (1)

where fqso is the flux density of the emission from the quasar nucleus (including the power-law continuum, the BEL, and the Fe ii components), fstellar is the flux density of stellar emission, Cf is the covering factor, and τ(λ) is the optical depth of absorption. The τ(λ) profile that we adopt for each absorption line is Gaussian with a fixed velocity dispersion. We perform this nuclear spectrum fitting in the vicinity of the narrow absorption lines, and the resultant best-fit model and the corresponding parameters are given in Figure 3 and Table 2, respectively. We find the redshift of these absorption lines to be z = 0.228827 ± 0.000002, highly consistent with that of the H i 21 cm absorption (z ∼ 0.2288; Aditya & Kanekar 2018). In addition, the H i 21 cm absorption line is spectrally unresolved, in line with the narrowness of the Na i and Ca ii absorption lines (σ ∼ 23 km s−1). These consistencies strongly imply a relation between the absorption of Na i/Ca ii and that of H i. Aditya & Kanekar (2018) measured the H i column density to be (3.54 ± 0.11) × (Ts /100 K) × 1020 cm−2, where Ts is the spin temperature. The H i column density N(H) and the Na i column density N(Na I) are related by the following equation (e.g., Rupke et al. 2005):

Equation (2)

where y is the ionization fraction, a is the Na abundance, and b describes the depletion onto dust. If we adopt y = 0.9 following Rupke et al. (2005), a solar abundance (so that a = −5.69), and the canonical Galactic depletion value of b = − 0.95, then we see that the measured N(Na I) corresponds to N(H) ∼ 1.1 × 1021 cm−2, a value close to the result of radio spectral analysis.

Figure 3.

Figure 3. The observed spectrum near the Ca ii (left) and Na i (right) narrow absorption lines are shown in black lines. The best-fitted spectra are shown in colored lines based on their components. The best-fitted spectra of the stellar component and stellar plus quasar components, and deep narrow absorption lines are shown in blue, red, and green, respectively.

Standard image High-resolution image

Table 2. Parameters for Modeling the Narrow Absorption Lines

ParameterValueUnit
z 0.228827 ± 0.000002 
σ 23.4 ± 0.6km s−1
Cf 0.78 
NNa I 26 ± 21012 cm−2
NCa II 3.4 ± 0.71012 cm−2

Download table as:  ASCIITypeset image

3.2. The Nature of the Quasar

BELs in the optical spectrum are unambiguously emitted from the quasar nucleus. We measure the Balmer decrement of the BELs to be 8.07 ± 0.05, significantly higher than the theoretical "Case" B value of 2.7 (Gaskell 2017), indicating heavy dust reddening toward the quasar nucleus. Assuming an SMC extinction curve results in an estimation that EBV ∼ 1.2. The power-law component contributes 60%–70% of the total continuum flux in the wavelength range of 4000–7000 Å. The quasar continuum is remarkably red with a power-law index α of 1.32 ± 0.02, probably a result of the synchrotron emission from the radio jet, and/or the reddened thermal emission from the quasar nucleus.

In view of the heavy dust extinction in optical bands, we use the infrared continuum to estimate the bolometric luminosity of the quasar. As the first step, we obtain the 5 μm monochromatic luminosity using the infrared SED constructed from the AllWISE photometry of the quasar, finding that ν Lν (5 μm) = 8 × 1044 erg s−1. We then follow Richards et al. (2006) to apply a bolometric correction factor of 8, reaching a bolometric luminosity of 6 × 1045 erg s−1, though this value may have been overestimated due to the possible contribution of synchrotron emission from the jet at 5 μm.

3.3. 2D Spectra of Narrow Emission Lines and Spatial Decomposition

To investigate the gas kinematics of the host galaxy through the 2D profiles of NELs, we remove the emission from the quasar nucleus and the stars, each of which is represented by double Gaussians in the spatial regions where the NELs are negligible. The two top panels of Figure 4 show the rest-frame 2D spectra in the neighborhood of the [O i], [S ii] λ λ 6717, 6731 doublet, the Hα, N ii doublet, and the Hβ, [O iii] λ λ 5007, 4969 doublet (note that we employ the stellar redshift, z = 0.22911). The NEL structure extends to a size of ∼2farcs2, greater than the FWHM of the point-spread function (PSF) 0farcs71 (see Appendix A), and thus is spatially resolved.

Figure 4.

Figure 4. Top two rows: the observed 2D spectra of the [O i], [S ii], Hβ, [O iii], and Hα after the continuum and [Fe ii] subtractions. Only spaxels where the flux is detected with S/N > 4 are plotted. The color of spaxels are scaled by flux density (squared scale, in units of 10−17 erg s−1 cm−2 arcsec−2 Å−1). The spectral direction is horizontal, and the spatial direction is vertical (southwest is positive). The 2D spectra of [O i]- and [S ii]-emitting NELs show a velocity gradient across the spatial extent, indicating a rotation-dominated disk. The 2D spectra of Hα and Hβ NELs imply the existence of an extended region located ∼4–20 kpc southwest of the quasar nucleus, whose position and velocity is reminiscent of EELRs around quasars. The 2D spectra of the [O iii] NEL is more complicated: besides the two components mentioned above, an additional blueshifted component centered on the quasar nucleus is evident. Such a profile is conventionally considered to be suggestive of the existence of outflowing gas. Our multiple-Gaussian decomposition of these emission lines allows us to isolate the contribution from these components. Third row: the reconstructed 2D spectra of blue peaks colored by the flux density. This shows the rotating disk plus the outflow component. Bottom row: the reconstructed 2D spectra of red peaks colored by the flux density. This reveals the EELR extended to the southwest ∼20 kpc.

Standard image High-resolution image

These 2D NEL spectra reveal three components. In particular, the 2D spectra of [O i]- and [S ii]-emitting NELs show a velocity gradient across the spatial extent, indicating a rotation-dominated disk. The velocity gradient can be seen in the 2D NEL spectra produced using both the two independent exposures (see Appendix B), and thus is reliable. The 2D spectra of Hα and Hβ NELs (Figure 4) imply the existence of an extended region located ∼4–20 kpc southwest of the quasar nucleus, whose position and velocity is reminiscent of EELRs around quasars (e.g., Stockton & MacKenty 1987; Fu & Stockton 2009). The 2D spectra of the [O iii] NEL is more complicated: besides the two components mentioned above, an additional blueshifted component centered on the quasar nucleus is evident. Such a profile is frequently seen in the 2D spectra of quasars' [O iii] emission and is conventionally considered to be suggestive of the existence of outflowing gas. Therefore, the three components of the 2D NEL spectra include a rotating disk, an EELR, and an outflow. In addition, we see double-peak profiles in [O iii] Hα and Hβ lines in the southwest of the nucleus, indicating multiple components.

To delineate the gas kinematics, we perform a two-step spectral fit to decompose these three components using the Python package MPFIT. In the first step, for each spatial element (spaxel), we fit a double Gaussian to the profile of Hβ, [O iii], Hα, and N ii. The resultant best-fit Gaussian models for a single spaxel are demonstrated in Figure 5. Our reconstructed 2D spectra of the red Gaussian show the EELR only, while the blue Gaussian is a superposition of a rotating disk and an outflow (Figure 4, the third and fourth rows). In the second step, for each individual spaxel, we perform a single-Gaussian fit to the [O i] emission line and the [S ii] λ λ 6717, 6731 doublet. The best-fit spectra of [O i] and the [S ii] doublet in seven spaxels are shown in Figure 6 (left panels), from northeast to southwest. Due to the higher complication of their profiles, we fit the Hα and Hβ profiles to a single or double Gaussian, for which the decision is made in a way similar to that in Liu et al. (2014). It turns out that the five spectra closest to the nucleus demand a double Gaussian. These fits allow us to successfully decompose the contribution from the rotating disk and a nearly spherical outflow structure, and the best-fit Hα and Hβ spectra of seven spaxels are shown in Figure 6 (right panels), from northeast to southwest.

Figure 5.

Figure 5. Examples of Hβ, [O iii], and Hα velocity profiles at the junction of the "inclined" and "tail" structures. The red and blue dashed lines are the best-fitted spectra of the red and blue peaks, respectively, with the sum of them shown by orange solid lines. The red and blue peaks reveal the "tail" and "inclined" structures, respectively.

Standard image High-resolution image
Figure 6.

Figure 6. The observed and best-fitted spectra of the [O i], [S ii], Hβ, and Hα at different spaxels, from southwest to northeast from top to bottom. The project distance from the nucleus is shown on the upper-right corner in each panel. The velocity of the rotating component is shown at the upper-right corner in blue. For those spectra best fitted by a single Gaussian, the median velocity is marked by a blue dotted line, and the zero-point of the velocity, which is measured by a stellar component, is shown by a gray dashed–dotted line. For those spectra best fitted by two Gaussians, the median velocity of the main Gaussian (i.e., the larger peak value) is marked by a blue dotted line. We note the Hβ spectrum at the nuclear (r = 0'') and Hα spectra within 0farcs3 of the nuclear one (r = 0.3, 0 and −0.3) are best fitted by two Gaussians, due to the outflow component. The green dashed lines, in the second column, are the two Gaussians from the double [S ii] lines. The orange dashed lines, in the third and fourth columns, are the two Gaussians of Hβ and Hα, respectively. The green dashed–dotted lines, in the fourth column, are the best-fitted double N ii lines.

Standard image High-resolution image

3.4. The Rotating Disk

We scrutinize the rotating disk by measuring the velocity relative to the redshift of host galaxy and the velocity dispersions of best-fit spectra of the rotation components in the NELs of [O i], [S ii], Hβ, Hα and N ii. In Figure 7, we plot them as a function of the distance from the quasar nucleus (positive is to the southwest and negative to the northeast). In addition, we only show those spaxels with a high signal-to-noise ratio (S/N of [O i] ≳ 3), corresponding to ∼5 kpc from the nucleus. We see that the overall velocity distribution and dispersion of these NELs are in line with typical galactic disks (Rix et al. 1992; Proshina et al. 2020). We find that the velocity profiles measured from different NELs are consistent, and thus use the median of the measurement results from these different NELs, following the method employed by Courteau (1997), Weiner et al. (2006), and Drew et al. (2018). In detail, assuming an axis ratio of b/a = 1, the distribution of velocity is formulated as:

Equation (3)

where Va is the asymptotic velocity, rt is the knee radius, robs is the projection distance from the center of the galaxy, c is the outer-galaxy slope, Vrel is the velocity of the ionized gas relative to the stars along the line of sight, and i is the inclination angle of the galaxy. Here Va and rt are dimensional scaling parameters, whereas c characterizes the shape of the rotation curve. Meanwhile, if a Gaussian fit is used, the profile of the velocity dispersion is given by:

Equation (4)

where Σ(r) is the value of the fitted Gaussian at each radius, m is the central position (if the ionized gas is isotropically distributed around the center, then m is 0), ω is the width of the Gaussian, and σ0 is the isotroptic component of the velocity dispersion. Here we correct the observed velocity dispersion for the intrinsic instrumental dispersion using $\sigma =\sqrt{{\sigma }_{\mathrm{obs}}^{2}-{\sigma }_{\mathrm{inst}}^{2}}$, where σ is reported in Figure 7, and σinst is the combined instrumental- and spectral-seeing dispersion that we measure to be ∼23 km s−1. The resultant best-fit velocity dispersion σ is 126 ± 10 km s−1.

Figure 7.

Figure 7. Long-slit line-of-sight velocity and velocity dispersion profiles obtained with ESI/Keck. The velocity (top) and velocity dispersion (bottom) of J0005+3820 (PA = −139°) measured by [O i], [S ii], Hβ, Hα, and N ii, respectively, are marked using gray pentagons, purple rhombuses, green squares, red stars, and blue triangles. And the black dots mark the median of the velocity and velocity dispersion measured by these emission lines. The errors on the median are calculated as $\sigma ={\sigma }_{\mathrm{NMAD}}/\sqrt{n-1}$, where σNMAD is the normalized median of the absolute deviations and n is to be 5 on behalf of the number of samples. The cyan solid lines show the best fit of the data. The abscissa zero corresponds to the brightest-continuum bin along the slit (to the galactic nucleus).

Standard image High-resolution image

We fit the velocity and velocity dispersion profiles to the median values of the measurements results from different NELs ([O i], [S ii], Hβ, Hα, and N ii), utilizing a Monte Carlo (MC) simulation that we run 50 times to calculate the errors of parameters. Errors on the median are calculated as $\sigma ={\sigma }_{\mathrm{NMAD}}/\sqrt{n-1}$, where σNMAD is the normalized median of the absolute deviations (Hoaglin et al. 1983) and n is the number of samples, here n = 5. The best-fit profiles are shown in Figure 7, along with the corresponding parameters listed in Table 3. The velocity profile reveals a roughly symmetric gas motion pattern. Gas in the quasar's nuclear region has a velocity of ∼−75 km s−1, and is blue-/redshifted on the southwest/northeast side, respectively. The velocity profile flattens beyond a radius of 4 kpc in both of these two directions, with a velocity of −150 to −180 km s−1 (southwest) and ∼40 km s−1 (northeast) within the distance range of 4–6 kpc. The rotating component is blueshifted relative to the systemic (stellar) velocity with Vrel = − 71 km s−1. Since the narrow absorption line system is blueshifted with a similar relative velocity (−85 km s−1; see Section 3.1.1), this kinetic consistency is suggestive of the same origin of the two.

Table 3. Parameters for Modeling the Velocity Shift and Velocity Dispersion

ParameterValueUnit
Velocity Shift
Va −196 ± 48km s−1
rt 4.1 ± 1.1kpc
c −0.8 ± 4.8
Vrel −71 ± 4km s−1
i 0.84 ± 0.07rad
Velocity Dispersion
A 24 ± 2
m 0.26 ± 0.01kpc
ω 0.63 ± 0.06kpc
σ0 126 ± 3km s−1

Download table as:  ASCIITypeset image

In the central region, we note that a velocity difference of ∼50 km s−1 exists at a > 3σ significant level between [S ii] and other NELs. This is likely due to the outflow, though the possibility of influence from the sky emission line in the blue outskirts of [S ii] cannot be fully ruled out, but if [S ii] is excluded from these analyses, our results change minimally.

3.5. The Extended Emission-line Region

The EELR extends to the southwest of the nucleus (see Figure 4), which is most evident in the [O iii] emission, where two compact knots and a diffuse region are seen. For the line spectra at each position, we measure the integrated flux, median velocity, and line width (W80, as detailed in Liu et al. 2013) of the Hα, Hβ, and [O iii] emission lines by fitting their profiles to Gaussian models, and conclude that the three NELs depict consistent velocity profiles (Figure 8).

Figure 8.

Figure 8. Long-slit line-of-sight velocity, velocity dispersion, and normalized flux density profiles obtained with ESI/Keck. The velocity, velocity dispersion, and normalized flux measured by [O iii], Hβ, and Hα are marked using red, yellow, and blue, respectively.

Standard image High-resolution image

The integrated flux profile of all three NELs unambiguously show a series of three peaks at galactocentric distances of about 7, 13, and 18 kpc. Hence, we divide the EELR into three annuli, with radius ranges of 2–8, 8–16, and 16–25 kpc, respectively, and for each of which we extract a 1D emission-line spectra (Figure 9). The best-fit integrated fluxes, median velocities, and line widths are listed in Table 4, where it can be seen that the emission lines from 2–16 kpc are relatively narrow and are redshifted relative to the stellar redshift, while at 16–25 kpc they are broader and blueshifted.

Figure 9.

Figure 9. Spectra of the Hα, Hβ, and [O iii] at three different regions. The red line shows the best fitting, the blue dotted line marks the median velocity, and the velocity zero is the gray dashed–dotted line.

Standard image High-resolution image

Table 4. Nonparametric Measurements of Three Parts

ParameterValueUnit
Part One
f[O iii] 28.8 ± 0.1010−17 erg s−1 cm−2 arcsec−2
v[O iii] 123 ± 0.12km s−1
W[O iii] 77 ± 0.31km s−1
fHα 8.6 ± 0.3210−17 erg s−1 cm−2 arcsec−2
vHα 119 ± 0.61km s−1
WHα 79 ± 1.58km s−1
fHβ 2.1 ± 0.1210−17 erg s−1 cm−2 arcsec−2
vHβ 130 ± 2.11km s−1
WHβ 82 ± 5.82km s−1
AV 1.186 ± 0.233mag
Part Two
f[O iii] 52.6 ± 0.7210−17 erg s−1 cm−2 arcsec−2
v[O iii] 108 ± 2.33km s−1
W[O iii] 170 ± 3.03km s−1
fHα 12.3 ± 1.6010−17 erg s−1 cm−2 arcsec−2
vHα 114 ± 6.75km s−1
WHα 149 ± 8.44km s−1
fHβ 3.4 ± 0.1710−17 erg s−1 cm−2 arcsec−2
vHβ 117 ± 10.08km s−1
WHβ 150 ± 25.15km s−1
AV 0.860 ± 0.475mag
Part Three
f[O iii] 21.3 ± 0.2310−17 erg s−1 cm−2 arcsec−2
v[O iii] −122 ± 7.75km s−1
W[O iii] 348 ± 8.38km s−1
fHα 3.3 ± 1.7010−17 erg s−1 cm−2 arcsec−2
vHα −171 ± 4.59km s−1
WHα 175 ± 11.87km s−1
fHβ 2.4 ± 0.1610−17 erg s−1 cm−2 arcsec−2
vHβ −116 ± 6.56km s−1
WHβ 238 ± 17.35km s−1

Note. Fluxes of the Hα, Hβ, and [O iii] are directly measured, without dust attenuation correction. The velocity is the median velocity measured by Hα, Hβ, and [O iii], and W is the W80 (Liu et al. 2013). We use the Milky Way attenuation curve starburst galaxies and the average extinction (Calzetti et al. 2000) to reddening relation at the V band of AV = 4.05E(BV).

Download table as:  ASCIITypeset image

The intensity ratios of the emission lines facilitate our analysis on the physical conditions of the EELR. Specifically, we use the [O iii]/Hβ and Hα/Hβ ratios to quantify the degree of ionization and the dust attenuation, respectively. To obtain higher S/N ratios, we bin the spatial pixels within 0.45'' before measuring the median values of these line ratios, for which the uncertainties are, again, calculated using the MC method (Figure 10). We find that [O iii]/Hβ is larger than 10 at all galactocentric radii under our consideration (top panel therein), implying a high-ionization state in general. Meanwhile, considering the theoretical ratio Hα/Hβ ∼ 2.86 for "case B" (ne = 100 cm−3, Te = 10,000 K; Storey & Hummer 1995), we plot the result in Figure 10 (blue dashed line, bottom panel). The Hα/Hβ ratio at galactocentric radii of 2–16 kpc are roughly constant, corresponding to AV ∼ 1. However beyond 16 kpc, Hβ is too weak for accurate measurement of dust attenuation.

Figure 10.

Figure 10. Line ratio of the [O iii]/Hβ and Hα/Hβ of the extended ionized gas profiles along the long-slit line of sight. Blue dashed lines mark the typical [O iii]/Hβ ratio and "Case B" value, respectively.

Standard image High-resolution image

The electron density is achievable from [O ii] I(3729)/I(3726) and [S ii] I(6717)/I(6731) ratios. For this purpose, we stack the spectra taken from locations 2–16 kpc away from the center, and fit the [O ii] and [S ii] doublet emission lines by fixing the kinematics of the two lines (Figure 11). Assuming an electron temperature of 10,000 K and at a 68% confidence level, we estimate the electron density to be ne = 137${}_{-30}^{+36}$ cm−3. This result is generally lower than that of typical narrow-line regions (ne ∼ 1000 cm−3; Greene et al. 2011), but close to that of the EELR (Fu & Stockton 2009).

Figure 11.

Figure 11. Spectra of the [O ii] and [S ii] emission lines in the rest frame. The fitted line is in red. The blue dashed line shows the two Gaussians from the double [O ii] lines. The gray shaded regions are influenced by the sky line.

Standard image High-resolution image

4. Discussion

4.1. Host Galaxy

As we mentioned in the Introduction, there are abundant works studying the host galaxies of blazars, particularly to distinguish whether the host galaxies are disk or elliptical galaxies. Most of them focused on imaging BL Lac objects, while none of them successfully found a disk-galaxy-hosted FSRQ. Because blazars are rare and can only be found in the distant universe, it is hard to spatially decompose the nuclei and the host galaxies well, even when using the Hubble Space Telescope's highest spatial resolution in the optical and NIR bands. However, we could use a spatially resolved spectrum to constrain the host galaxy properties. In disk galaxies, young stars and interstellar gas and dust rotate in disks around bulging nuclei, while in elliptical galaxies, old stars randomly swarm and gas and dust are lacking. Thus, the properties and kinematics of stars and gases can be used to distinguish between disk and elliptical galaxies.

In Section 3.4, we find that the kinematics of gas in the host galaxy of B2 0003+38A are dominated by rotation, and the curvatures of velocity and velocity dispersion are similar to those of disk galaxies (Ho & Martin 2020). In addition, there are other hints that support B2 0003+38A being hosted by a disk galaxy. The velocity of the ionized gas (rotating component) relative to the systemic (stellar) velocity with Vrel = −71 km s−1 is similar to the velocity of the narrow absorption line system (−85 km s−1), suggesting the same origin for these two. Furthermore, emission lines from host galaxy can be detected significantly, indicating a gas-rich host galaxy. The spectrum lacks the features of old stars, such as TiO molecular bands, indicating that old stars are not dominant. Meanwhile, the spectrum is red, suggesting that the starlight is dust reddened. Young stars and rich interstellar dust are also characteristic of disk galaxies. This is consistent with the analysis of a nuclear spectrum, which prefers a young stellar population (∼450 Myr) with large dust reddening (E(BV) = 0.62), under the assumption of a single SSP and quasar spectrum. Therefore, we conclude that B2 0003+38A is mostly likely hosted by a disk galaxy.

4.2. The Extended Emission-line Region

We find the EELR to the southwest of the nucleus, and it extends to a projected distance of up to 25 kpc. EELRs with such sizes were found around a substantial fraction of radio-loud quasars.

Before discussing the origin of the EELR further, we estimate some parameters of the EELR. The gases in parts one and two in Figure 9 have redshifted velocities of v0 ∼ 120 km s−1, and their distance to the galaxy nucleus is R0 ∼ 15 kpc (see Figure 9). We estimate a dynamical timescale to be t ∼ 1.2 × 108 yr, i.e., the time taken by the gas from the nucleus to reach such a distance with an average velocity of v0. The mass of gas in the two parts can be estimated using Hβ luminosity LHβ and electron density ne (e.g Liu et al. 2013; Harrison et al. 2014). After correction for dust attenuation, the summed luminosity in the two parts is ∼2.4 × 1040 erg s−1. The total mass of this ionized extended gas can be estimated as:

Equation (5)

We measure an electron density of ne ∼ 137 cm−3 using [O ii] and [S ii] doublets, if the electron temperature is 10,000 K. We find Mgas ∼ 9.3 × 106 M. Combined with the average velocity of 120 km s−1, the total kinetic energy of the gas can be estimated as:

Equation (6)

If assuming the lifetime of the structure is the dynamical timescale, we can also estimate a mass rate $\dot{M}$ is 7.8 × 10−2 M yr−1, and a kinetic energy rate ${\dot{E}}_{\mathrm{ink}}\sim 3.5\times {10}^{38}$ erg s−1.

However, this electron density value may not be the bulk density of the EELR gas. The EELR might be illuminated by the quasar. If so, the ionization parameter U of the EELR can be estimated as:

Equation (7)

where QH is the rate of the hydrogen ionization photons from the quasar, r is the distance from the quasar nucleus to the EELR, and nH is the hydrogen number density. Assuming that the intrinsic SED of the quasar is that given in Mathews & Ferland (1987), we estimate a QH of 1.2 × 1056 s−1 using the bolometric luminosity estimated in Section 3.2. Assuming a distance of 10 kpc, and assuming ne = 1.2nH for highly ionized plasma, the inferred U is about 10−2.5. However, the observed [O iii]/Hβ values of 12–16 indicate a higher U value of ∼10−1. This may be because the [O iii] emitting gas has a lower density than the previously estimated value. Using the mixed-medium model (Robinson et al. 2000), Stockton et al. (2002) divided the EELR of 4C 37.43 into two components, and found one component with a density of several hundred cm−3, and another with a density near 1 cm−3. Most of the [O iii], Hβ, and Hα come from regions with low density; [O ii] and [S ii] come from the regions with a density several hundred times higher. If this is the case for the EELR in B2 0003+38A, the actual value of Mgas is two orders of magnitude higher.

There are various possibilities for the origin of an EELR. In the case of an isolated galaxy, the gas can be an inflow, i.e., a cold accretion flow from the intergalactic medium, an outflow triggered by an AGN or a starburst, or the recycling of gas (of the outflow). In the case of a galaxy with interactions, the gas can also belong to tidal features. First, inflows are generally isotropic, which excludes the possible of the EELR being inflow driven. Second, EELRs are common in radio quasars, and these EELRs are generally related with outflows driven by radio jets (e.g., Fu & Stockton 2009), so the EELR in B2 0003+38A may also be the same. Third, verifying the possibility of the recycling of gas requires a panoramic image of the circumgalactic gas. The long-slit spectrum is not sufficient, and future integral field unit (IFU) observation data are needed. We did not see a companion galaxy around B2 0003+38A, and neither did we see any asymmetry from the brightness profile of the starlight. As for the possibility of galaxy interactions, we do not find any direct evidence of them, including any signature of asymmetry from the brightness profile of the starlight, or any companion galaxy around it. However, we note that the current data quality is not enough to rule out the possibility. The EELR may also originate in tidal features. If so, starlight accompanied with the EELR should be seen. This cannot be tested using the existing data. Higher-quality future observations, such as IFU, are required to fully investigate the origin of EELRs.

5. Conclusions

In this paper, we present long-slit observations taken from ESI/Keck to study the gas in the host galaxy of FSRQ B2 0003+38A at redshift z = 0.22911. Based on multiple Gaussian-fitting processes, we separate the 2D NEL spectra into three components, indicating a rotation disk, an EELR, and an outflow, respectively. To model the rotating disk, we measure and analyze the curves of velocity and velocity dispersion. We also analyze the EELR, which extends to a projected distance up to 25 kpc from the nucleus. We summarize our results below.

  • 1.  
    For the first time, we discover a rotating gaseous disk from optical spectroscopy in an FSRQ host galaxy. The curvatures of velocities and the velocity dispersions derived from different emission lines agree with an identical rotating-disk model. The rotating gas disk has a mean velocity of v = −75 km s−1 relative to the stellar redshift. The velocity has little difference with that of the absorber seen in Na i, Ca ii, and H i lines (v = −85 km s−1), suggesting that the rotating gas disk and the absorber are related.
  • 2.  
    According to the kinematics and morphology of the EELR, we divide them into three parts, including two knots and a diffused region. We calculate that the two knots have an averaged electron density ne of ${137}_{-30}^{+36}$ cm−3 and dust attenuation AV of 1.19 ± 0.23/0.86 ± 0.48, respectively. After correcting for the dust attenuation, we estimate a corresponding mass to be 9.3 × 106 M. The velocity of ionized gas in these two knots is redshifted by 120 km s−1. The dynamical timescale of the knots can be estimated at ∼1.2 × 108 yr as the travel time of clouds to reach the observed distances from the center. There are various possibilities for the origin of an EELR, including inflows, outflows, gas recycling, and galaxy interactions. If the EELR is from an outflow/inflow, the mass rate is 7.8 × 10−2 M yr−1, and the kinetic energy carried by the ionized gas is estimated as 1.46 × 1054 erg.

Q.Z., L.S., and G.L. acknowledge the grant from the National Key R&D Program of China (2016YFA0400702), the National Natural Science Foundation of China (Nos. 11673020 and 11421303), and the Fundamental Research Funds for the Central Universities. We acknowledge the support from the Chinese Space Station Telescope (CSST) Project. We acknowledge support by the Fundamental Research Funds for the Central Universities.

Appendix A: Measuring the Point-spread Function

We measure the PSF due to the seeing and instrumental effects using the spatial brightness profiles of BELs. We can do it in this way because the BEL region, with a typical size less than 1 pc, is a point source at z ∼ 0.2. We extract the spatial brightness profiles of the Hα and Hβ BELs (Figure 12). Both the two profiles can be well fit using a Gaussian function with a FWHM of 0farcs71.

Figure 12.

Figure 12. The spatial brightness profiles of Hα and Hβ BELs, and the Gaussian functions that fit them.

Standard image High-resolution image

Appendix B: Verifying the Velocity Gradient of NELs across the Spatial Extent

One may doubt that velocity gradient of NELs across the spatial extent seen from the 2D spectra might be artificial: it might be caused by a tiny inclined movement of the target along the slit during the observation. As there were two exposures of this target, we test this presumption by independently analyzing the data from the two exposures. We reduce the data and extract the 2D NEL spectra again following the methods described in Sections 2.3 and 3.3, while this time we do not stack the two exposures. The results are shown in Figure 13. The velocity gradient of NELs are seen in both the two 2D NEL spectra, indicating that the rotation-dominated disk structure in B2 0003+38A is reliable.

Figure 13.

Figure 13. Same as Figure 4. First row: the observed 2D spectra of the [O i], Hα, N ii and [S ii] from the data of the first exposure. Second row: those from the second exposure. Note that the cosmic rays are not removed.

Standard image High-resolution image

Footnotes

Please wait… references are loading.
10.3847/1538-4357/abf4de