Multicolor Optical Monitoring of the Quasar 3C 273 from 2005 to 2016

, , , , , , and

Published 2017 March 29 © 2017. The American Astronomical Society. All rights reserved.
, , Citation Dingrong Xiong et al 2017 ApJS 229 21 DOI 10.3847/1538-4365/aa64d2

Download Article PDF
DownloadArticle ePub

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

0067-0049/229/2/21

Abstract

We have monitored the quasar 3C 273 in the optical V, R, and I bands from 2005 to 2016. Intraday variability (IDV) is detected on seven nights. The variability amplitudes on most of the nights are less than 10%, and on four nights, more than 20%. When considering the nights with time spans >4 hr, the duty cycle (DC) is 14.17%. Over the 12 years, the overall magnitude and color index variabilities are ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 67$, ${\rm{\Delta }}R=0\buildrel{\rm{m}}\over{.} 72$, ${\rm{\Delta }}V=0\buildrel{\rm{m}}\over{.} 68$, and ${\rm{\Delta }}(V-R)=0\buildrel{\rm{m}}\over{.} 25$, respectively. The largest clear IDV has an amplitude of 42% over just 5.8 minutes, and the weakest detected IDV is 5.4% over 175 minutes. The BWB (bluer when brighter) chromatic trend is dominant for 3C 273 and appears at different flux levels on intraday timescales. The BWB trend exists for short-term timescales and intermediate-term timescales but different timescales have different correlations. There is no BWB trend for our whole time-series data sets. A significant anticorrelation between the BWB trend and length of timescales is found. Combining with V-band data from previous works, we find a possible quasi-periodicity of P = 3918 ± 1112 days. The possible explanations for the observed variability, BWB chromatic trend, and periodicity are discussed.

Export citation and abstract BibTeX RIS

1. Introduction

Active galactic nuclei (AGNs) are very energetic extragalactic sources powered by accretion on supermassive black holes (Esposito et al. 2015). The blazar subclasses of radio-loud AGNs have jets pointing in the direction of the observer and are characterized by large amplitude and rapid variability at all wavelengths, high and variable polarization, superluminal jet speeds, and compact radio emission (Angel & Stockman 1980; Urry & Padovani 1995). Blazars are often subclassified into BL Lacertae (BL Lac) objects and flat spectrum radio quasars (FSRQs). FSRQs have strong emission lines, while BL Lac objects only have very weak or non-existent emission lines. The emission of blazars is dominated by a relativistic jet, which is boosted by the beaming effect (Sandrinelli et al. 2016). The broadband spectral energy distributions (SEDs) of blazars have a double-peaked structure. The low-energy peak at the IR–optical–UV band is explained by the synchrotron emission of relativistic electrons, and the high-energy peak at the GeV–TeV gamma-ray band is due to the inverse Compton (IC) scattering (e.g., Dermer 1995; Dermer & Schlickeiser 2002; Bottcher 2007). The hadronic model is an alternative explanation for the high-energy emissions from blazars (e.g., Dermer et al. 2012).

Blazars show variability on different timescales, from years down to minutes (e.g., Fan 2005; Zhang et al. 2008; Fan et al. 2009, 2014; Poon et al. 2009). Blazar variabilities can be broadly divided into intraday variability (IDV) or micro-variability, short-term variability (STV), and long-term variability (LTV). Variations in flux of a few tenths or hundredths of a magnitude on a timescale of tens of minutes to a few hr are often known as IDV (Wagner & Witzel 1995). STV has timescales of days to weeks, even months, and LTV ranges from months to years (Gupta et al. 2008; Dai et al. 2015). The IDV of blazars is the least understood type of variation, but it can shed light on the location, size, structure, and dynamics of the emitting regions and radiation mechanism (e.g., Wagner & Witzel 1995; Ciprini et al. 2003, 2007; Dai et al. 2015; Kalita et al. 2015). The discovery of periodicity in the variability could have profound consequences on the global understanding of the sources, constituting a basic building block for models (Sandrinelli et al. 2016). In order to search for periodic variations in many timescales, long-term observations are needed.

The quasar 3C 273 (R.A. = 12h29m06fs7, decl. = 02°03'09'', J2000, redshift z = 0.158) is the first quasar discovered by Smith & Hoffleit (1963). It is classified as a bright gamma-loud FSRQ (Abdo et al. 2010b), and has been observed for more than 50 years in the radio band and more than 100 years in the optical band (Vol'vach et al. 2013). The quasar has blazar-like (superluminal jet and high variability) and Seyfert-like (the strong blue bump, the soft X-ray excess, a weak and neutral iron line, and variable emission lines) characteristics (Esposito et al. 2015; Chidiac et al. 2016). The flux and spectral variations of 3C 273 have been extensively studied over the entire electromagnetic spectrum (e.g., Xie et al. 1999, 2004a; Collmar et al. 2000; Ghosh et al. 2000; Mantovani et al. 2000; Sambruna et al. 2001; Greve et al. 2002; Kataoka et al. 2002; Courvoisier et al. 2003; Attridge et al. 2005; Jester et al. 2005; Savolainen et al. 2006; McHardy et al. 2007; Espaillat et al. 2008; Soldi et al. 2008; Dai et al. 2009; Fan et al. 2009, 2014; Pacciani et al. 2009; Abdo et al. 2010b; Ikejiri et al. 2011; Rani et al. 2011; Vol'vach et al. 2013; Beaklini & Abraham 2014; Esposito et al. 2015; Kalita et al. 2015; Madsen et al. 2015; Chidiac et al. 2016). Soldi et al. (2008) studied the multiwavelength variability of 3C 273. Their results showed variability at all frequencies, with amplitudes and timescales strongly depending on the energy, and implied that either two separate components (possibly a Seyfert-like and a blazar-like component) or at least two parameters with distinct timing properties can account for the X-ray emission below and above ∼20 keV. Soldi et al. (2008) also noted that the optical component is consistent with optically thin synchrotron radiation from the base of the jet, and the hard X-rays would be produced through inverse Compton processes (SSC and/or EC) by the same electron population. Abdo et al. (2010) presented the γ-ray outburst light curves and spectral data from 3C 273 in September 2009. During these flares, 3C 273 reached very high brightness levels for quite short time intervals of 1–10 days, and showed that the rise and the decay are asymmetric on timescales of 6 hr, and that the spectral index is significantly harder during the flares. Esposito et al. (2015) studied the high-energy spectrum of 3C 273 and suggested a two-component model to explain the complete high-energy spectrum. Chidiac et al. (2016) found that the variations at higher frequencies are leading the lower frequencies, which could be expected in the standard shock-in-jet model. Kalita et al. (2015) found high flux variability over long timescales in optical, UV, and X-ray bands, and that on intraday timescales, 3C 273 shows small amplitude variability in X-ray bands. Madsen et al. (2015) investigated the spectral variability in the NuSTAR band and found an inverse correlation between flux and photon index. In the optical band, unlike other blazars that are highly polarized, 3C 273 is a low-polarization quasar (Valtaoja et al. 1991; Xie et al. 1999; Chidiac et al. 2016). Generally, the properties of 3C 273 are very different from that of BL Lac objects and highly polarized quasars in the optical bands. It showed moderate variation in amplitude. Variations of 0.2–0.3 mag can be seen on a timescale of a few days, but ${\rm{\Delta }}V\geqslant 1$ mag was not seen even on a timescale of many years (Smith et al. 1987; Xie et al. 1999). Small amplitude changes over a short time were reported (e.g., Moles et al. 1986; Rani et al. 2011). However, for the quasar, Fan et al. (2009) also presented IDVs of 0.55 mag in 13 minutes and 0.81 mag in 116 minutes. Paltani et al. (1998) presented two distinct variable components in the optical–ultraviolet bands. One rapidly variable component could be from the accretion disk (Shields 1978; Soldi et al. 2008), or illumination by an X-ray source (Ross & Fabian 1993) or a hot corona (Haardt et al. 1994). The second component seems to be related to synchrotron emission from the jet (Paltani et al. 1998). Optical variability is often associated with color/spectral behavior in blazars, which can be used to explore the emission mechanism (e.g., Zheng et al. 2008; Gu & Ai 2011; Xiong et al. 2016). The results from Dai et al. (2009) implied that the spectrum becomes bluer (flatter) when the source becomes brighter both on IDV and LTV for 3C 273. Ikejiri et al. (2011) found that 3C 273 exhibits "bluer-when-brighter" trends in their time-series data sets. Fan et al. (2014) found that the spectrum becomes flat when the source becomes bright, and that the emission in the optical bands consists of two components. The complexity of the 3C 273 emission is due to the presence of different emission components. Therefore, long-term monitoring of 3C 273 is very important to better understand the physics of the quasar.

In view of these facts, we monitored the source in the optical band from 2005 to 2016. We analyze the variability and spectral properties of 3C 273. Combining with V-band data from Soldi et al. (2008) and Fan et al. (2014), we search the periodic signals of 3C 273 in 48 year time spans. This paper is organized as follows. The observations and data analysis are described in Section 2. Section 3 presents the results. Discussion and conclusions are reported in Section 4. A summary is given in Section 5.

2. Observations and Data Analysis

Our optical monitoring program of 3C 273 was carried out using the 2.4 m and 1.02 m optical telescopes located at the Yunnan Astronomical Observatories (YAO) of China. The 2.4 m optical telescope is equipped with two photometric detectors: PI VersArray 1300B CCD camera (PICCD) and Yunnan Faint Object Spectrograph and Camera (YFOSC).8 The 2.4 m optical telescope used PICCD from 2008 to 2012 but YFOSC after 2012. The PICCD, with 1340 × 1300 pixels, covers a field of view (FOV) of 4.48 × 4.40 arcmin2 at the Cassegrain focus. The readout noise and gain are 6.05 electrons and 1.1 electrons/ADU, respectively. The YFOSC has a FOV of about 10 × 10 arcmin2 and 2000 × 2000 pixels for photometric observation. Each pixel corresponds to 0.283 arcsec/pixel. The readout noise and gain of the YFOSC CCD are 7.5 electrons and 0.33 electrons/ADU, respectively (Liao et al. 2014). For the 1.02 m optical telescope,9 prior to 2006, the entire CCD chip (1024 × 1024 pixels) covered 6.5 × 6.5 arcmin2. The readout noise and gain of the CCD are 3.9 electrons and 4.0 electrons/ADU, respectively. After 2009, the telescope was equipped with an Andor DW436 CCD (2048 × 2048 pixels) camera at the Cassegrain focus (f = 13.3 m), with pixel size 13.5 × 13.5 μm2. The readout noise and gain are 6.33 electrons and 2.0 electrons/ADU, respectively, with 2 μs (readout time per pixel) or 2.29 electrons and 1.4 electrons/ADU with 16 μs. The FOV of the CCD image is 7.3 × 7.3 arcmin2 and its pixel scale is 0.21 arcsec/pixel (Dai et al. 2015; Xiong et al. 2016). The standard Johnson broadband filters were used for the two telescopes.

Our photometry observations were performed in the V, R, and I bands through two modes. The first mode was that all of the observations were completed for the same optical band and then moved to the next band. The other was a cyclic mode among the V, R, and I bands. The optical observations in the V, R, and I bands were in a corresponding cyclic mode for half of the total data. The exposure times from 0.17 to 8.3 minutes were chosen according to different seeing and brightness of the source. Moreover, in order to search for very fast variability (<10 minutes) and obtain more data, the short exposure time in the first mode was adopted. For the same optical band in our observatories, the time resolutions (minimum times between two adjacent data points for the same optical band) were from 0.2–20.5 minutes, and most of the time resolutions were <15 minutes. Therefore, we can consider these data in the cyclic mode to be quasi-simultaneous measurements (<20 minutes). The rest of the data were not explored in the analysis of the interband time lag and color index because their V, R, and I bands were not observed in a corresponding cyclic mode. The observation log is given in Table 1 where we have listed the observation date, time spans, time resolutions, and number of data points for each date in different bands. From Table 1, there are many nights with time spans <3 hr due to bad weather and time devoted to other targets.

Table 1.  Observation Log of Photometric Observations

Date(UT) Number(I, R, V) Time Spans(hr) Time Resolutions(minutes)
2005 Apr 05 4, 4, 4 2.7 20
2005 Apr 06 15, 15, 15 2.3 9
2006 Jan 09* 8, 8, 8 0.6 2
2006 Jan 10* 8, 8, 8 0.6 2
2006 Jan 11* 8, 8, 8 0.65 1.5
2006 Apr 02* 4, 4, 3 0.1 2
2006 Apr 03* 4, 4, 4 0.15 2
2006 May 05* 11, 6, 11 1.42, 0.26, 1.42 2.5, 3, 2.4
2006 May 06* 6, 6, 6 0.2 2.2
2007 Mar 27* 39, 0, 40 1, 1.75 1.1, 1.6
2007 Mar 28* 20, 0, 11 2.5, 2.3 2, 4
2007 Mar 29* 15, 0, 15 2.2 2.1
2007 Mar 30* 15, 0, 15 1.5 1.5
2007 Apr 15* 40, 0, 10 2.89, 0.58 2.1, 3.8
2007 Apr 22* 16, 0, 5 0.68, 0.17 2.3, 2.3
2007 Apr 23* 10, 0, 7 1.85, 0.5 2.1, 3.8
2007 Apr 24* 8, 0, 6 3, 0.33 2.2, 3.4
2007 May 08* 20, 0, 10 1.6, 1.39 2.3, 3.9
2007 May 09* 11, 0, 5 1.3, 0.3 2.3, 3.9
2007 May 10* 40, 0, 5 1.5, 1.1 2.1, 3.9
2008 Jan 10* 15, 15, 0 0.5 2
2008 Jan 11* 30, 10, 0 0.78, 0.31 1.5, 2
2008 Jan 12* 20, 20, 0 0.6, 0.63 1.4, 1.9
2008 Apr 22* 10, 10, 10 0.24 0.75
2008 Apr 23* 15, 15, 15 0.5 0.42
2009 Jan 17* 20, 15, 0 2.39, 1.64 2.64, 3.48
2009 Jan 18* 10, 10, 0 0.8 2.6
2009 Jan 19* 7, 7, 7 0.12, 0.12, 0.12 0.72, 0.73, 0.84
2009 Mar 21* 5, 10, 5 0.04, 0.33, 0.04 0.6, 0.6, 0.6
2009 Mar 22* 5, 10, 5 0.04, 0.31, 0.04 0.6, 0.6, 0.6
2009 Mar 23 5, 5, 5 0.21 3.08
2009 Mar 26 5, 5, 5 0.15 2.3
2009 Apr 13 12, 12, 12 0.47 2.4
2009 Apr 14 24, 24, 24 1.12 2.7
2009 Apr 15 32, 32, 32 1.26 2.4
2009 Apr 16* 10, 10, 0 0.27 1.8
2009 May 11* 75, 0, 20 2.41, 1.17 1.15, 2.15
2009 Dec 14 11, 11, 11 0.22 1.3
2010 Jan 04* 20, 20, 20 0.1 0.3
2010 Jan 09* 15, 15, 15 1 2.19
2010 Jan 10* 5, 5, 5 0.2 2
2010 Jan 14* 5, 5, 5 0.2 2
2010 Jan 16* 0, 109, 0 2.14 0.2
2010 Feb 15* 10, 18, 10 0.07, 0.45, 0.08 0.5, 0.5, 0.6
2010 Feb 20* 5, 5, 5 0.2 2
2010 Feb 21* 5, 5, 5 0.1 1
2010 Feb 22* 5, 5, 5 0.1 1
2010 Feb 26* 40, 40, 40 0.62 0.2
2010 Feb 27* 20, 30, 30 0.28, 1.72, 1.75 0.2, 0.2, 0.4
2010 Feb 28* 10, 10, 10 0.03 0.2
2010 Mar 11* 20, 20, 20 2.31 0.3
2010 Mar 14 5, 6, 5 0.18 3
2010 Mar 19* 70, 20, 15 2.36, 0.76, 0.98 0.8, 1.2, 2.2
2010 Apr 04 30, 30, 30 5.25 10
2010 Apr 05 30, 30, 30 5.04 10
2010 Apr 07* 9, 10, 10 0.08 0.5
2010 Apr 23* 30, 30, 30 0.3 0.2
2010 Apr 25* 30, 30, 30 0.33 0.2
2010 May 02* 5, 5, 5 0.08 1
2010 May 03* 5, 5, 5 0.03, 0.12, 0.12 0.5, 1.9, 1.9
2010 Jun 01* 10, 10, 10 0.05 0.3
2011 Apr 12 10, 10, 10 2.31 15
2011 Apr 13 12, 12, 12 3.32 19
2012 Feb 24 4, 0, 0 0.12 3
2012 Feb 25 6, 0, 0 0.21 3
2012 Feb 26 4, 0, 0 0.2 4
2012 Feb 27 5, 5, 5 0.26 4
2012 Feb 28 5, 5, 5 0.26 4
2012 Apr 02 5, 5, 5 0.26 4
2012 Apr 03 6, 0, 5 0.26 3
2012 Apr 28 4, 0, 4 0.13 3
2012 Apr 29 4, 0, 4 0.15 3
2012 May 01 5, 5, 5 0.43 7
2012 May 02 5, 5, 5 0.54 8
2012 May 11 5, 5, 5 0.4 6
2012 May 13 5, 5, 5 0.44 7
2012 Dec 19 4, 18, 4 0.36, 1.98, 0.36 7, 2.2, 7
2012 Dec 20 4, 18, 4 0.27, 0.71, 0.27 5.4, 1.7, 5.4
2012 Dec 21 4, 64, 4 0.27, 1.98, 0.27 5.4, 1.7, 5.4
2013 Apr 01 16, 16, 15 3.6 14
2013 Apr 04 20, 20, 20 4.8 14
2014 Apr 19 13, 13, 13 4.02 19
2014 Apr 20 15, 13, 10 4.36 19
2014 Apr 21 80, 80, 80 3.52 3
2014 Apr 22 13, 11, 12 3.74 19
2014 Apr 23 16, 16, 12 3.21 13
2014 Apr 24 16, 16, 14 3.24 13
2014 Apr 25 25, 25, 18 5.19 13
2014 Apr 26 13, 13, 14 2.8 13
2015 Apr 08 26, 26, 26 4.34 11
2015 Apr 12 17, 17, 18 2.93 11
2015 Apr 13 23, 24, 23 4.63 11
2015 Apr 14 22, 22, 21 4.79 11
2015 Apr 15 27, 27, 24 4.51 11
2015 May 17 20, 18, 20 1.85 5.8
2015 May 19 13, 14, 15 3.41 8
2016 Mar 29 20, 20, 20 1.35 4.3
2016 Mar 30 52, 50, 52 3.8 4.3
2016 Mar 31 54, 52, 54 5.8 4.3
2016 Apr 23 54, 55, 50 4 4.3
2016 Apr 24 93, 94, 92 6.7 4.3
2016 Apr 25 96, 96, 95 6.8 4.3
2016 Apr 26 97, 97, 96 6.9 4.3
2016 May 07 12, 12, 12 2 4.9
2016 May 08 16, 16, 16 1.2 4.9
2016 May 09 31, 31, 31 3.4 5
2016 May 10 16, 19, 18 2 5.9
2016 May 27 37, 37, 37 1.9 3

Note. The "*" stands for non-quasi-simultaneous measurements in the V, R, and I bands (observed in the first mode).

Download table as:  ASCIITypeset images: 1 2

After flat-field and bias corrections, aperture photometry was performed using the APPHOT task of IRAF.10 The aperture radius of 2 × FWHM was selected, considering the best S/N ratio. The finding chart of 3C 273 was obtained from the Web page Finding Charts for AGN.11 Besides the instrumental magnitude of 3C 273, we measured the instrumental magnitudes of four comparison stars (C, D, E, and G in the finding chart) on the same field. We chose C and D as comparison stars because C was the brightest comparison star and the differential magnitude between C and D had the smallest variations among the four comparison stars. Following Zhang et al. (2004, 2008), Fan et al. (2014), and Bai et al. (1998), the source magnitude was given as the average of the values derived with respect to the two comparison stars ($\tfrac{{m}_{C}+{m}_{D}}{2}$, mC and mD are the blazar magnitudes obtained from standard star C and from standard star D, respectively). The magnitudes of the comparison stars in the field of 3C 273 were taken from Smith et al. (1985). The rms errors of the photometry on a certain night are calculated from the two comparison stars, star C and star D, in the usual way:

Equation (1)

where ${m}_{i}={({m}_{{\rm{C}}}-{m}_{{\rm{D}}})}_{i}$ is the differential magnitude of stars C and D, while $\overline{m}=\overline{{m}_{C}-{m}_{D}}$ is the averaged differential magnitude over the entire data set, and N is the number of observations on a given night. In order to further quantify the reliability of the variability, the value of Sx can be calculated as (e.g., Hu et al. 2014)

Equation (2)

where mi and $\overline{m}$ are the same as in Equation (1).

The variability amplitude (Amp) can be calculated by (Heidt & Wagner 1996)

Equation (3)

where Amax and Amin are the maximum and minimum magnitudes, respectively, of the light curve for the night being considered, and σ is the rms error.

The duty cycle (DC) is calculated as (Romero et al. 1999; Stalin et al. 2009; Hu et al. 2014)

Equation (4)

where ${\rm{\Delta }}{T}_{i}={\rm{\Delta }}{T}_{i,\mathrm{obs}}{(1+z)}^{-1}$, z is the redshift of the object, and ${\rm{\Delta }}{T}_{i,\mathrm{obs}}$ is the duration of the monitoring session of the ith night. Note that since for a given source the monitoring durations on different nights were not always equal, the computation of DC has been weighted by the actual monitoring duration ${\rm{\Delta }}{T}_{i}$ on the ith night. Ni will be set to 1 if IDV is detected, otherwise Ni = 0 (Goyal et al. 2013).

The actual number of observations for 3C 273 is 105 nights, obtaining 1901 I-band, 1707 R-band, and 1544 V-band data points. The results of the observations are given in Tables 24 for filters I, R, and V.

Table 2.  I-band Data

Date(UT) MJD Magnitude σ
2005 Apr 05 53464.698 12.053 0.014
2005 Apr 05 53464.717 12.074 0.014
2005 Apr 05 53464.731 12.084 0.014
2005 Apr 05 53464.811 12.055 0.014
2005 Apr 06 53465.685 12.057 0.011

Note. Column (1) is the universal time (UT) of the observation, column (2) is the corresponding modified Julian day (MJD), column (3) is the magnitude, and column (4) is the rms error.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

Table 3.  R-band Data

Date(UT) MJD Magnitude σ
2005 Apr 05 53464.703 12.538 0.059
2005 Apr 05 53464.719 12.514 0.059
2005 Apr 05 53464.733 12.529 0.059
2005 Apr 05 53464.815 12.517 0.059
2005 Apr 06 53465.687 12.530 0.009

Note. The meanings of the columns are the same as in Table 2.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

Table 4.  V-band Data

Date(UT) MJD Magnitude σ
2005 Apr 05 53464.709 12.659 0.040
2005 Apr 05 53464.723 12.687 0.040
2005 Apr 05 53464.737 12.655 0.040
2005 Apr 05 53464.819 12.686 0.040
2005 Apr 06 53465.688 12.695 0.015

Note. The meanings of the columns are the same as in Table 2.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

3. Results

3.1. Variability

A number of statistical tests have been proposed to prove the validity of IDV reports. Normally, the C-test was often used to analyze the IDV. However, de Diego (2010) has found that the C-test is not a reliable methodology, because it does not have a Gaussian distribution and its criterion is too conservative. At present, the optical IDV is often analyzed by the F-test, χ2-test, modified C-test, and one-way analysis of variance (ANOVA; de Diego 2010; Joshi et al. 2011; Goyal et al. 2012, 2013; Hu et al. 2014; Agarwal & Gupta 2015; Dai et al. 2015). Romero et al. (1999) introduced the variability parameter, C, as the average value between C1 and C2:

Equation (5)

where (BL-StarA), (BL-StarB), and (StarA-StarB) are the differential instrumental magnitudes of the blazar and comparison star A, the blazar and comparison star B, and comparison stars A and B. σ is the standard deviation of the differential instrumental magnitudes. The adopted variability criterion requires C ≥ 2.576, which corresponds to a 99% confidence level. Despite the very common use of C-statistics, de Diego (2010) has pointed out that it has severe problems.

The F-test is thought to be a proper statistics to quantify variability (e.g., de Diego 2010; Joshi et al. 2011; Goyal et al. 2012; Hu et al. 2014; Agarwal & Gupta 2015; Xiong et al. 2016). The F value is calculated as

Equation (6)

where Var(BL-StarA), Var(BL-StarB), and Var(StarA-StarB) are the variances of the differential instrumental magnitudes. The F value from the average of F1 and F2 is compared with the critical F value, ${F}_{{\nu }_{{bl}},{\nu }_{* }}^{\alpha }$, where νbl and ν* are the number of degrees of freedom for the blazar and comparison star, respectively ($\nu =N-1$), and α is the significance level set as 0.01 (2.6σ). If the average F value is larger than the critical value, the blazar is variable at a confidence level of 99%. Other alternatives to the standard F-test are ANOVA or the χ2 test (e.g., de Diego 2010). de Diego (2010) showed that ANOVA is a powerful and robust estimator for microvariations. We use ANOVA in our analysis because it does not rely on error measurement but derives the expected variance from subsamples of the data. Considering the time of exposure, we bin the data in groups of three or five observations (see Xiong et al. 2016 and de Diego 2010 for detail). This method is used only for light curves with more than 20 observations on a given night; if the measurements in the last group are less than three or five, then it is merged with the previous group. The critical value of ANOVA can be obtained by ${F}_{{\nu }_{1},{\nu }_{2}}^{\alpha }$ in the F-statistics, where ${\nu }_{1}=k-1$ (k is the number of groups), ${\nu }_{2}=N-k$ (N is the number of measurements), and α is the significance level (Hu et al. 2014). When analyzing, we remove the outliers, where Sx is more than 3σ.

The blazar is considered variable (V) if the light curve satisfies the two criteria of the F-test and ANOVA. The blazar is considered probably variable (PV) if only one of the above two criteria is satisfied. The blazar is considered non-variable (N) if none of the criteria are met. We only analyze the data with exposure times of more than 1 minute because much shorter exposure times produce lower quality data and their use makes the analysis of the detection of IDVs much more difficult. The results of our analysis of the IDV are shown in Table 5. From Table 5, it can be seen that there is IDV found on five nights (I band on 2013 April 04, 2015 April 15, 2010 March 19, and 2009 January 17, and V band on 2016 April 25; Figure 1). As an illustration, the specific variabilities on two nights are presented. On 2013 April 04, the largest magnitude change of the I band is ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 055$ in 175 minutes from MJD = 56386.621 to MJD = 56386.742 corresponding to the variability amplitude Amp = 5.39%. On 2015 April 15, 3C 273 brightens by ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 038$ in 52 minutes from the beginning of MJD = 57127.656 to MJD = 57127.692, and then fades by ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 029$ in 41 minutes from MJD = 57127.692 to MJD = 57127.721. From Table 5 and Figure 1, we can see that the variability amplitudes on these nights are close to 5%. For the R band from 2014 April 25 and the I band from 2015 May 17, the light curves meet the criterion of the F-test, but do not reach the critical value of ANOVA. Though we use different bin sizes (N = 2–6), the two nights still do not reach the critical value of ANOVA. It is can be seen that the light curves of the two nights have large variation with Amp > 30% (Table 5 and Figure 1). The ANOVA test is used to compare the means of a number of samples. For the light curves of the two nights, when 3C 273 flares or darkens, there are only a few data points in the corresponding time interval, which reduces the variations between groups, and causes the ANOVA test to not detect the IDV (see de Diego 2010 for details). Therefore, according to results from the F-test, the two nights are detected as IDVs with a large variability amplitude. On 2014 April 25, 3C 273 brightens by ${\rm{\Delta }}R=0\buildrel{\rm{m}}\over{.} 27$ in 51 minutes from the beginning of MJD = 56772.668 to MJD = 56772.703, and then quickly fades by ${\rm{\Delta }}R=0\buildrel{\rm{m}}\over{.} 27$ in 26 minutes from MJD = 56772.703 to MJD = 56772.721. On 2015 May 17, 3C 273 brightens by ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 27$ in 18 minutes from the beginning of MJD = 57159.651 to MJD = 57159.663, and then quickly fades by ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 42$ in 5.8 minutes from MJD = 57159.663 to MJD = 57159.667 and again brightens by ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 22$ in 5.9 minutes from the beginning of MJD = 57159.667 to MJD = 57159.671. The outliers (Sx > 3σ) only appear in two nights (2015 April 15 and 2016 April 25) and never exceed two data points for per night.

Figure 1.

Figure 1. Light curves of the IDV for 3C 273. The black squares and lines are the light curves for 3C 273. The red circles and lines are the variations of SI, SR, and SV. The light curves of SI, SR, and SV are offset to avoid their eclipsing with the light curves of 3C 273.

Standard image High-resolution image

Table 5.  Results of the IDV Observations of 3C 273

Date Band N T(hr) F FC(99) FA FA(99) V/N A(%) Ave(mag)
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11)
2007 Mar 27 V 40 1.75 0.76 2.16 1.76 3.26 N 6.24 12.66
2007 Mar 28 I 20 2.48 1.03 3.03 4.24 4.89 N 1.65 12.00
2007 Apr 15 I 40 2.89 0.45 2.16 5.67 3.26 PV 9.53 11.98
2007 May 08 I 20 1.59 0.68 3.03 0.66 4.89 N 2.71 12.03
2007 May 10 I 40 1.48 0.52 2.16 0.96 3.26 N 3.03 12.01
2008 Jan 11 I 30 0.78 0.85 2.42 4.56 3.90 PV 1.15 12.02
2008 Jan 12 I 20 0.48 0.63 3.03 1.34 5.29 N 1.18 12.05
2008 Jan 12 R 20 0.63 0.76 3.03 0.08 5.29 N 1.22 12.45
2009 Jan 17 I 20 2.39 3.93 3.03 21.15 4.70 V 6.99 12.19
2009 May 11 I 75 2.41 0.41 1.75 4.11 2.44 PV 1.50 12.20
2009 May 11 V 20 1.17 0.87 3.03 10.72 5.29 PV 3.50 12.80
2010 Mar 19 I 70 2.36 3.74 1.75 109.00 2.47 V 4.70 12.15
2010 Mar 19 R 20 0.76 0.57 3.03 15.62 5.29 PV 1.77 12.57
2010 Apr 04 I 30 5.21 0.66 2.42 5.04 3.46 PV 2.66 12.22
2010 Apr 04 R 30 5.23 0.61 2.42 2.95 3.90 N 0.85 12.66
2010 Apr 04 V 30 5.25 0.68 2.42 3.39 3.46 N 1.60 12.80
2010 Apr 05 I 30 5.04 1.00 2.42 13.57 3.46 PV 1.73 12.21
2010 Apr 05 R 30 5.04 0.93 2.42 3.00 3.46 N 1.83 12.64
2010 Apr 05 V 30 5.04 0.55 2.42 2.06 3.46 N 0.59 12.79
2012 Dec 21 R 64 1.98 0.97 1.82 4.69 2.35 PV 1.54 12.56
2013 Apr 04 I 20 4.80 7.35 3.03 9.03 4.70 V 5.39 12.16
2013 Apr 04 R 20 4.80 1.53 3.03 4.97 4.70 PV 3.78 12.59
2013 Apr 04 V 20 4.81 0.86 3.03 2.34 4.70 N 2.75 12.72
2014 Apr 21 V 80 3.52 0.94 1.69 0.75 2.13 N 20.23 12.81
2014 Apr 25 I 25 5.19 1.07 2.66 3.02 3.93 N 6.91 12.24
2014 Apr 25 R 25 5.19 7.65 2.66 2.42 3.93 V 30.06 12.66
2015 Apr 08 I 26 4.34 1.06 2.60 1.27 3.84 N 4.39 12.40
2015 Apr 08 R 26 4.34 3.26 2.60 1.76 3.84 PV 6.48 12.85
2015 Apr 08 V 26 4.34 3.33 2.60 0.22 3.84 PV 6.88 12.99
2015 Apr 13 I 23 4.58 2.57 2.79 1.08 4.20 N 9.04 12.40
2015 Apr 13 R 24 4.60 5.15 2.79 0.53 4.02 PV 15.14 12.86
2015 Apr 13 V 23 4.63 1.51 2.78 1.47 4.20 N 8.04 12.99
2015 Apr 14 I 22 4.79 2.93 2.86 1.25 4.30 PV 15.87 12.40
2015 Apr 14 R 20 4.62 3.90 2.86 0.13 4.30 PV 14.90 12.89
2015 Apr 14 V 20 4.62 4.74 2.94 1.43 4.46 PV 24.33 13.00
2015 Apr 15 I 27 4.51 3.58 2.55 8.30 3.70 V 5.72 12.41
2015 Apr 15 R 26 4.51 0.63 2.55 1.88 3.70 N 3.16 12.87
2015 Apr 15 V 24 4.51 2.04 2.72 3.91 4.03 N 3.69 13.00
2015 May 17 I 20 1.85 20.68 3.03 2.02 4.62 V 41.93 12.38
2016 Mar 29 V 20 1.35 5.06 3.03 1.80 4.70 PV 4.69 12.83
2016 Mar 30 V 52 3.76 0.64 1.95 0.85 2.57 N 3.40 12.85
2016 Mar 31 V 54 5.77 2.02 1.92 1.67 2.51 PV 20.83 12.83
2016 Apr 23 V 50 3.83 0.56 1.95 4.91 2.62 PV 6.31 12.87
2016 Apr 24 V 92 6.74 0.84 1.63 1.44 2.04 N 9.87 12.85
2016 Apr 25 V 95 6.74 2.27 1.63 4.18 2.00 V 3.45 12.84
2016 Apr 26 V 96 6.87 1.26 1.62 2.30 2.00 PV 4.07 12.84
2016 May 09 V 31 3.40 0.79 2.39 0.28 3.40 N 3.51 12.86
2016 May 27 V 37 1.86 0.45 2.21 8.85 3.06 PV 1.48 12.85

Note. Column 1 is the date of the observation, column 2 the observed band, column 3 the number of data points, column 4 the time spans, column 5 the average F value, column 6 the critical F value with 99% confidence level, column 7 the F value of ANOVA, column 8 the critical F value of ANOVA with 99% confidence level, column 9 the variability status (V: variable, PV: probably variable, N: non-variable), and columns 10–11 the variability amplitude and daily average magnitudes, respectively.

Download table as:  ASCIITypeset image

To sum up, IDV is detected on seven nights. For the seven nights, we also check the color variations on intraday timescales. However, the results from three statistical tests do not show the corresponding color variations on intraday timescales. There are 14 nights detected as PV, excluding the seven nights. When estimating the variability amplitude (Amp), we only consider the night detected as "V" and "PV". The distributions of the variability amplitude from different bands are given in Figure 2, which shows that the variability amplitudes for most nights are less than 10%, and for four nights, more than 20% (also see Table 5). The correlations between the variability amplitudes and the source average brightness are shown in Figure 3. The results from the analysis of the Spearman rank indicate that there are no significant correlations between variability amplitude and brightness (significance level P > 0.05, N = 11, 7, 9 for the I, R, and V bands). However, there are not enough data to support the results. Making use of Equation (4), we calculate the DC of the IDV. The value of the DC is 10.84% for the V case and 53.65% for the PV+V cases. When considering the nights with time spans >4 hr, the value of the DC is 14.17% for the V case and 55.24% for the PV+V cases.

Figure 2.

Figure 2. IDV amplitude distributions. The black solid lines stand for nights detected as a V case, the red dashed lines for nights detected as a PV case.

Standard image High-resolution image
Figure 3.

Figure 3. Variability amplitude vs. the average brightness. The blue circles, red circles, and black squares stand for the V, R, and I bands, respectively.

Standard image High-resolution image

Long-term light curves and color index variations are given in Figure 4. From Figure 4, we can see that on the whole, the quasar becomes dark from 2005 to 2015 but also bright between different years; in 2015, the quasar reaches a low flux state; compared with 2015, the quasar in 2016 has a tendency to brighten; there is a color index variability for long timescales. Over the 12 years, the overall magnitude and color index variabilities are ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 67$, ${\rm{\Delta }}R=0\buildrel{\rm{m}}\over{.} 72$, ${\rm{\Delta }}V=0\buildrel{\rm{m}}\over{.} 68$, and ${\rm{\Delta }}(V-R)=0\buildrel{\rm{m}}\over{.} 25$, respectively. The magnitude distributions in the V, R, and I bands are 13fm10 < V < 12fm42, 13fm09 < R < 12fm37, and 12fm59 < I < 11fm93, respectively. The average values of the magnitude and color index are $\langle I\rangle =12\buildrel{\rm{m}}\over{.} 178\pm 0\buildrel{\rm{m}}\over{.} 118$, $\langle R\rangle =12\buildrel{\rm{m}}\over{.} 645\,\pm 0\buildrel{\rm{m}}\over{.} 107$, $\langle V\rangle \,=12\buildrel{\rm{m}}\over{.} 791\pm 0\buildrel{\rm{m}}\over{.} 106$, and $\langle V-R\rangle =0\buildrel{\rm{m}}\over{.} 126\pm 0\buildrel{\rm{m}}\over{.} 023$, respectively.

Figure 4.

Figure 4. Long-term light curves of 3C 273 in the V, R, and I bands and the color index VR.

Standard image High-resolution image

3.2. Cross-correlation Analysis and Variability Timescales

Using the z-transformed discrete correlation function (ZDCF; Alexander 1997), we perform the interband correlation analysis and search for possible interband time delay. The ZDCF estimates the cross-correlation function (CCF) in the case of non-uniformly sampled light curves, and deals with sparsely and unequally sampled light curves better than both the interpolated cross-correlation function (ICCF) and the discrete correlation function (DCF; Edelson et al. 1996; Alexander 1997; Giveon et al. 1999; Roy et al. 2000). The ZDCF attempts to correct the biases that affect the original DCF by using equal-population binning. The ZDCF involves three steps (Alexander 1997; Oscoz et al. 2001). (i) All possible pairs of observations are sorted according to their time lag, and binned into equal-population bins of at least 11 pairs. Multiple occurrences of the same point in a bin are discarded so that each point appears only once per bin. (ii) Each bin is assigned its mean time lag and the intervals above and below the mean that contain 1σ of the points each. (iii) The correlation coefficients of the bins are calculated and z-transformed. The error is calculated in z-space and transformed back to r-space. The time lag corresponding to the maximum value of the ZDCF is assumed as the time delay between both components. The ZDCF code of Alexander (1997) can automatically set how many bins are given and used to calculate the interband correlation and the ACF. For each correlation, a Gaussian fitting is made to find the central ZDCF points. The time where the Gaussian profile peaks denotes the lag between the correlated light curves (Wu et al. 2012). The results that can determine the time delay are displayed in Figure 5. From Figure 5 and the results of the Gaussian fitting, the corresponding time lags are −0.002 ± 0.001, 0.002 ± 0.002, −0.007 ± 0.003, 0.003 ± 0.002, 0.011 ± 0.004, 0.014 ±0.003, 0.026 ± 0.006, and −0.005 ± 0.004, respectively. Considering the time resolutions, we do not find significant time lags between the V-band magnitude and I-band magnitude. For the rest of the nights, there are not enough data points or good Gaussian profiles to determine time lags.

Figure 5.

Figure 5.  ZDCF plots between the V and I bands. The curves show Gaussian fittings to the points, and their peaks are marked with the vertical lines.

Standard image High-resolution image

The autocorrelation function (ACF) is defined by

Equation (7)

where the brackets denote a time average. The ACF measures the correlation of the light curve with itself, shifted in time, as a function of the time lag τ (Giveon et al. 1999). The zero-crossing time is the shortest time it takes the ACF to fall to zero (Alexander 1997). If there is an underlying signal in the light curve, with a typical timescale, then the width of the ACF peak near zero time lag will be proportional to this timescale (Giveon et al. 1999; Liu et al. 2008). The zero-crossing time is a well-defined quantity and used as a characteristic variability timescale (e.g., Netzer et al. 1996; Alexander 1997; Giveon et al. 1999; Liu et al. 2008; Liao et al. 2014). The width of the ACF may be related to the characteristic size scale of the corresponding emission region (Abdo et al. 2010a; Chatterjee et al. 2012). Another function used in variability studies to estimate the variability timescales is the first-order structure function (SF; e.g., Trevese et al. 1949) defined by

Equation (8)

There is a simple relation between the ACF and the SF,

Equation (9)

where V is the variance of the light curve (Giveon et al. 1999). We therefore perform only an ACF analysis on our light curves. The ACF was estimated by ZDCF. We only analyze the nights detected as IDV. The results used to estimate the characteristic variability timescales are given in Figure 6. Following Giveon et al. (1999), Liu et al. (2008), and Liao et al. (2014), we use a least-squares procedure to fit a fifth-order polynomial to the ACF, with the constraint that the ACF(τ = 0) = 1. The fitting results show that the detected variability timescales are 0.107, 0.023, 0.005, and 0.028 days for the I band on 2013 April 04, R band on 2014 April 25, and I bands on 2015 May 17 and 2015 April 15. In Section 3.1, we have given the timescales corresponding to the change of brightness on these nights (also see Figure 1). By comparison, we obtain that the variability timescales from the ACF analysis are consistent with the timescales corresponding to the change of brightness.

Figure 6.

Figure 6. Results of the ACF analysis. The red dashed line is a fifth-order polynomial least-squares fit.

Standard image High-resolution image

3.3. Correlation between Magnitude and Color Index

In this section, we analyze the relationships between the magnitude and color index for intraday timescales, short-term timescales, and whole time-series data sets. The whole time-series data sets include data from 2005 to 2016, but exclude the data from 2006 to 2008 due to non-quasi-simultaneous measurements. For the color index, we correct for the Galactic extinction. The correction factors of the Galactic extinction are from Schlafly & Finkbeiner (2011). We concentrate on the VR index and V magnitude because the VR index versus V magnitude is frequently studied. When exploring the relationships, we only analyze the data with more than nine color indices obtained on intraday timescales and quasi-simultaneous measurements. The results of correlations between the VR index and V magnitude are given in Table 6. As an example, Figure 7 shows the correlations between the VR index and V magnitude on intraday timescales. The results from Table 6 show that most of the nights have strong correlations between the VR index and V magnitude on intraday timescales, and the rest of the nights have moderate or weak correlations. Therefore, a bluer-when-brighter (BWB) chromatic trend is dominant for 3C 273 on intraday timescales. The BWB trend exists for short-term timescales and intermediate-term timescales but different timescales have different correlations between the VR index and V magnitude (Table 6). The correlation between the VR index and V magnitude in the whole time-series data sets is shown in Figure 8. The result of the correlation analysis from Table 6 shows that there is no BWB trend in the whole time-series data sets. We have the opportunity to explore the relationship between the BWB trend and the length of timescales in the V band because of long-term optical monitoring. We use the coefficient of correlation from error-weighted linear regression analysis to indicate the intensity of the BWB trend and the sum of the monitoring time spans on intraday timescales to indicate length of timescales. Analysis of the Spearman rank shows that there is significant anticorrelation between the BWB trend and length of timescales (r = −0.491, P = 0.002; see Figure 9).

Figure 7.

Figure 7. Correlations between the VR index and V magnitude. The red solid lines are the results of linear regression analysis.

Standard image High-resolution image
Figure 8.

Figure 8. Correlation between the VR index and V magnitude for our whole time-series data sets.

Standard image High-resolution image
Figure 9.

Figure 9. Coefficient of correlation vs. length of timescales in the V band. The red empty circles and black filled circles stand for intraday and longer timescales, respectively.

Standard image High-resolution image

Table 6.  Results of Error-Weighted Linear Regression Analysis

Date(UT) N r P
2016 Mar 29 20 0.55 0.01
2016 Mar 30 50 0.42 0.002
2016 Mar 31 51 0.36 0.01
2016 Apr 23 50 0.24 0.09
2016 Apr 24 90 0.31 0.003
2016 Apr 25 95 0.35 <0.0001
2016 Apr 26 96 0.39 <0.0001
2016 May 09 31 0.52 0.003
2016 May 10 18 0.63 0.005
2016 Mar 29–2016 May 10 501 0.39 <0.0001
2015 Apr 15 24 0.29 0.16
2015 Apr 14 20 0.37 0.11
2015 Apr 13 23 0.58 0.004
2015 Apr 12 17 0.26 0.32
2015 Apr 08 26 0.67 0.00017
2015 May 17 17 0.44 0.08
2015 May 19 12 0.72 0.009
2015 Apr 08–2015 May 19 139 0.39 <0.0001
2014 Apr 19 11 0.64 0.03
2014 Apr 20 10 0.98 <0.0001
2014 Apr 21 79 0.82 <0.0001
2014 Apr 22 10 0.88 0.0008
2014 Apr 23 12 0.6 0.04
2014 Apr 24 14 0.59 0.025
2014 Apr 25 17 0.35 0.16
2014 Apr 26 12 0.84 0.0006
2014 Apr 19–2014 Apr 26 165 0.61 <0.0001
2013 Apr 01 15 0.3 0.28
2013 Apr 04 20 0.47 0.036
2013 Apr 01–2013 Apr 04 35 0.26 0.13
2012 Feb 27–2012 Dec 21 47 0.5 0.0003
2010 Apr 04 30 0.57 0.001
2010 Apr 04–2010 May 03 65 0.45 0.0002
2009 Apr 15 32 0.66 <0.0001
2009 Apr 14 24 0.68 0.0003
2009 Mar 26–2009 Dec 14 84 0.87 <0.0001
2005 Apr 06 15 0.82 0.0002
2005 Apr 05–2005 Apr 06 19 0.89 <0.0001
All data 1082 0.04 0.18

Note. N is the number of data, r is the coefficient of correlation, and P is the chance probability.

Download table as:  ASCIITypeset image

3.4. Periodicity Analysis

In order to search for periodicity, well-covered long-term light curves are required. Soldi et al. (2008) presented an update of 3C 273's database hosted by the ISDC.12 3C 273's database, which was first published by Turler et al. (1999), is one of the most complete multiwavelength databases currently available for the quasar. The time spans of the optical V-band database are from 1968 to 2005. However, there are no V-band data from 1998 to 1999. We also compiled V-band data in 1998–1999 from Fan et al. (2014). Our data from 2005 to 2016 are updates and supplementary to the above data. For the data from Fan et al. (2014) and the present work, we convert magnitudes to fluxes in the Geneva photometric system as described in Turler et al. (1999), without any additional correction (Vcorr = 0) because the data from Fan et al. (2014) and the present work used standard Johnson photometric system while Soldi et al. (2008) and Turler et al. (1999) used the Geneva photometric system. For data from Soldi et al. (2008), we consider data of Flag = 0 in order to use only the best quality data and exclude the data with the influence of strong synchrotron flares (Flag = 1). Combining all data, we search the periodic signals of 3C 273 in 48 year time spans. The long-term light curves (bin = 1 day) are given in Figure 10. When there is more than one data on intraday timescales, we use the standard deviation as the error.

Figure 10.

Figure 10. Long-term light curves from 1968 to 2016. The red circles are our latest data, and the black squares are data from Soldi et al. (2008) and Fan et al. (2014).

Standard image High-resolution image

For the periodicity analysis of AGNs, unevenly sampled data, frequency-dependent red noise, flares of high activity, and total monitoring time need to be considered (e.g., Schulz & Statteger 1997; Schulz & Mudelsee 2002; Vaughan 2005, 2010; Fan et al. 2014; Sandrinelli et al. 2016). For the data, the strong synchrotron flares are excluded. The time spans from the data are 48 years. Therefore, when analyzing periodicity, we mainly consider unevenly sampled data and red noise. In addition, in order to reduce false periodicity produced by short timescale variability, we pay more attention to periodicity in year timescales. In order to increase the reliability of the periodicity analysis, we use three methods (REDFIT: Schulz & Mudelsee 2002; Lomb–Scargle method: Lomb 1976 and Scargle 1982; Jurkevich method: Jurkevich 1971) to explore periodicity. The periodicity can be considered valid only if three methods have consistent results. Schulz & Mudelsee (2002) presented a computer program (REDFIT3.8e13 ) estimating red-noise spectra directly from unevenly spaced time series by fitting a first-order autoregressive (AR1) process. The program can be used to test if peaks in the spectrum of a time series are significant against the red-noise background from an AR1 process, and removes the bias of this Fourier transform for unevenly spaced data by correcting for the effect of correlation between Lomb–Scargle Fourier components. However, this program has two underlying assumptions: (i) the noise background recorded in a time series can indeed be approximated by an AR1 process; (ii) the distribution of data points along the time axis is not too clustered. For our data, the results of the non-parametric runs test indicate that the spectrum is consistent with an AR1 model (rtest = 139 falls inside the 98% acceptance region [129, 170]; see Schulz & Mudelsee 2002 and usage38e for further details). Figure 10 also shows that the distribution of data points along the time axis is not too clustered. So, it is appropriate to use REDFIT3.8e for periodicity analysis. The Lomb–Scargle method is commonly used to detect periodicity in unevenly sampled time series. The periodogram is a function of circular frequency ω, and is defined by the formula (Li et al. 2016)

Equation (10)

where X(ti) (i = 0, 1..., N0) is a time series. τ is calculated by the equation

Equation (11)

where ω = 2πν. Thus, the periodogram is a function of frequency ν. For a true signal X(ti), the power in PX(ω) would present a peak, or the power of a purely noise signal would be an exponential distribution. For a power level z, the False Alarm Probability (FAP) is calculated by (Scargle 1982; Press et al. 1994, p. 569; Li et al. 2016)

Equation (12)

where N is the number of data points. The Jurkevich method is an alternative and powerful method that is insensitive to the mean shape of the periodicity, less sensitive than Fourier analysis to uneven sampling. The Jurkevich method involves testing a run of trial periods around which the data is folded. The data is binned in phases around each trial period, and the total variance for all the points in the bin computed. A good period will give a much reduced variance (${V}_{{\rm{m}}}^{2}$). No firm rule exists for assessing the significance of a minimum in the ${V}_{{\rm{m}}}^{2}$. But a good guide is the fractional reduction of the variance $\left(f=\tfrac{1-{V}_{{\rm{m}}}^{2}}{{V}_{{\rm{m}}}^{2}}\right)$. Generally, a value f ≥ 0.5 (${V}_{{\rm{m}}}^{2}\leqslant 0.667$) implies that there is a very strong periodicity in the data (Jurkevich 1971; Kidger et al. 1992). We use half of the FWHM of the peak as errors of the periodicity. For REDFIT, when estimating errors of the periodicity, we choose the minimum of the peak height above the average power level of the red noise in the vicinity of the investigated period. The results of the periodicity analysis are given in Figure 11. The results of REDIFIT show that there is a broad peak (the center frequency Vpeak = 2.55 × 10−4 day−1) with 90% significance levels and above the red-noise power level. The corresponding periodicity is 3918 ± 1112 days in year timescales. The results of the Lomb–Scargle method show that there are three significant peaks, P1 (4961 ± 1072 days), P2 (3848 ± 471 days), and P3 (2793 ± 325 days) above 0.01 FAP levels. The results of the Jurkevich method show that from 2749 (±36) days to 5594 (±44) days, there are some strong periodicity signals. The strong periodicity signal of 3950 ± 14 days also appears in the results of the Jurkevich method. We do not consider periods >5840 days because the periods (considered) are limited to three times periods less than the total time spans (48 years). From the above results, we should note that the results from the REDFIT and Lomb–Scargle methods have large errors, and that for the Jurkevich method, there are some possible periods with range from 2749 days to 5594 days. Therefore, within the errors of the periods, the periods estimated by the three methods have consistent results. A possible quasi-periodicity of P = 3918 ± 1112 days is found.

Figure 11.

Figure 11. Results of the periodicity analysis. The upper, middle and bottom panels are the results of the REDFIT, Lomb–Scargle method, and Jurkevich method respectively. For the upper panel, the black line is bias-corrected power spectra. Curves starting from the bottom are the theoretical red-noise spectrum, 90%, 95% and 99% χ2 significance levels. The amplified subregion is located at the upper right of the panel. The marked periodicity is 3918 ± 1112 days. For the middle panel, the red line indicates 0.01 FAP levels. The P1, P2, and P3 peaks are 4961 ± 1072, 3848 ± 471, and 2793 ± 325 days. For the bottom panel, The amplified subregion is located at the lower left of the panel. The marked periods are 2749 ± 36, 3950 ± 14, and 5594 ± 44 days. The green dashed line stands for ${V}_{{\rm{m}}}^{2}=0.667.$

Standard image High-resolution image

4. Discussion and Conclusions

4.1. Variability

Long-term multiband observations of the quasar 3C 273 were performed from 2005 to 2016 in the V, R, and I bands. The magnitude distributions in the V, R, and I bands are 13fm10 < V < 12fm42, 13fm09 < R < 12fm37, and 12fm59 < I < 11fm93, respectively. The average values of the magnitude are $\langle I\rangle =12\buildrel{\rm{m}}\over{.} 178\pm 0\buildrel{\rm{m}}\over{.} 118$, $\langle R\rangle =12\buildrel{\rm{m}}\over{.} 645\pm 0\buildrel{\rm{m}}\over{.} 107$, and $\langle V\rangle =12\buildrel{\rm{m}}\over{.} 791\pm 0\buildrel{\rm{m}}\over{.} 106$, respectively. Toone (2004) observed this source during the period of 1980–2004 and presented the yearly averaged light curve. The magnitude distributions in the V band from Toone (2004) are 13fm12–12fm5 with $\langle V\rangle =12\buildrel{\rm{m}}\over{.} 86\pm 0\buildrel{\rm{m}}\over{.} 13$. The magnitude distributions in 2000–2008 from Fan et al. (2009) are 13fm567–12fm204, 13fm313–12fm014, and 12fm669–11fm628 for the V, R, and I bands respectively. The magnitude distributions in 1998–2008 from Fan et al. (2014) are 12fm9–12fm5, 12fm8–12fm3 and 12fm2–11fm85 for V, R and I bands respectively. The magnitude distributions in 2003–2005 from Dai et al. (2009) are 12fm824–12fm663, 12fm714–12fm338 and 12fm318–12fm055 in V, R and I bands with the mean magnitudes 12fm698, 12fm441, and 12fm139, respectively. The magnitude distributions in 1996 from Xie et al. (1999) are 13fm9–12fm86, 12fm98–12fm6, and 12fm47–12fm14 in the V, R, and I bands, respectively. Ikejiri et al. (2011) showed that the V-band magnitude distributions between 2008 and 2010 are 12fm75–12fm51. From the above references, we obtain that the maximum and minimum magnitudes for this source data are 13fm9–12fm2, 13fm31–12fm01, and 12fm67–11fm63 for the V, R, and I bands respectively. Through comparisons, we can find that our distributive ranges of magnitude in the V, R, and I bands are within the ranges of the maximum and minimum magnitudes from previous observations. From Figure 10, we also get the same conclusion. Moreover, our long-term multiband observations are updates and supplementary to the quasar data. During our observation, on the whole the quasar becomes dark from 2005 to 2015 but also bright between different years. In 2015, the quasar reaches a low flux state. Compared with 2015, the quasar in 2016 has a tendency to brighten. Over the 12 years, the overall magnitude variabilities are ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 67$, ${\rm{\Delta }}R=0\buildrel{\rm{m}}\over{.} 72$, and ${\rm{\Delta }}V=0\buildrel{\rm{m}}\over{.} 68$, respectively. These results indicate that unlike other blazars, 3C 273 shows moderate magnitude variations on a timescale of many years.

In order to accurately determine the optical IDV, we use two statistical tests to analyze the observation data. The blazar is considered variable only if the light curve satisfies the two criteria of the F-test and ANOVA. During the analysis, we remove the outliers, that is, Sx is more than 3σ. Through these tests and processing, we attain the reliable result that IDV is found on five nights. On 2013 April 04, the largest magnitude change of the I band is ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 055$ in 175 minutes, corresponding to the variability amplitude Amp = 5.39%. The variability amplitudes for the rest of the four nights are close to 5%. Rani et al. (2011) presented two nights detected as an IDV (Amp ∼ 5%) using the C- and F-tests. For the R band from 2014 April 25 and the I bands from 2015 May 17, the light curves meet the criteria of the F-test, but do not reach the critical value of ANOVA. We still consider the two nights detected as do IDV because the light curves of the two nights have large variations with Amp > 30% and ANOVA easily misses this kind of IDV since there are only a few data points in the time interval of flaring or darkening (see Section 3.1). On 2015 May 17, 3C 273 quickly faded by ${\rm{\Delta }}I=0\buildrel{\rm{m}}\over{.} 42$ in 5.8 minutes. Fan et al. (2009) noted that the quasar showed IDVs of 0.55 mag in 13 minutes and 0.81 mag in 116 minutes. The results we also obtain show that the variability amplitudes for most nights are less than 10% and for four nights, more than 20%. The value of the DC is 10.84% for the V case and 53.65% for the PV+V cases. Extensive IDV studies of different subclasses of AGNs revealed that the occurrence of IDV in a blazar observed on a timescale of <6 hr is ∼60%–65%; if the blazar is observed more than 6 hr, then the possibility of IDV detection is 80%–85% (Gupta & Joshi 2005; Rani et al. 2011). Then it is quite likely that the value of the DC is related to the time spans of observation. When considering the nights with time spans >4 hr, the value of the DC is 14.17% for the V case and 55.24% for the PV+V cases. Therefore, the quasar has low variability amplitudes and DC, which are different from other FSRQs. However, we still need more nights with longer time spans to confirm the results further. Although our results show that there are no significant correlations between the variability amplitude and brightness, we do not consider this result reliable because the result is not supported by enough data.

We notice that the variability amplitudes for the non-variable nights are non-zero and greater than 5% or even 10% (Table 5). The variability amplitudes from three nights with exposure time ∼1 minute (V bands on 2007 March 27, 2014 April 21, and 2016 April 24) are more than 5%. If we choose the bin = 3 minutes, the variability amplitudes decrease to 2.4%, 6.17%, and 3.48% respectively. Gopal-Krishna et al. (2003) and Goyal et al. (2013) have found that IRAF and DAOPHOT produce nominal errors that are too small by factors of ∼1.5. It is possible that the non-zero variability amplitudes for the non-variable nights are due to very small values of σ in Equation (3). Assuming variability amplitudes below 1% for the non-variable nights, the underestimated factors range from 1 to 4.7, with average values ∼2. If the original variability amplitudes from Table 5 for the non-variable nights are below 1%, the underestimated factors are set to 1, i.e., the value of σ is not underestimated. For the nights detected as the V case, using the underestimated factors ∼2, we recalculate the variability amplitudes. The results show that compared with the original variability amplitudes from Table 5, the recalculated variability amplitudes do not change significantly. Therefore, for some nights, the uncertainties in Tables 24 are underestimated by factors of ∼2. On the other hand, the ANOVA (such as the I band on 2015 May 17 or nights of low sampling rate) and standard F-test could fail to detect IDV (Joshi et al. 2011). This may also partly explain the problem.

From the results of the ACF, the detected characteristic variability timescales are 0.107, 0.023, 0.005, and 0.028 days for the I band on 2013 April 04, the R band on 2014 April 25, and the I bands on 2015 May 17 and 2015 April 15, respectively. Given the minimum values of the lags that can be measured, we cannot claim 0.023, 0.005, and 0.028 days as the significance of the timescales. Since the flares are essentially random and only one or two are seen on 2013 April 04, the variability timescale of 0.107 days (154 minutes) is considered as a possible variability timescale. We also use a least-squares procedure to fit fourth-order and sixth-order polynomials to the ACF on 2013 April 04. The fitting results of different order polynomials show that the differences in the variability timescales are within 2 minutes.

For blazars, the intrinsic origin of IDV is relativistic jet activities or accretion disk instabilities (Marscher & Gear 1985; Chakrabarti & Wiita 1993; Mangalam & Wiita 1993; Marscher et al. 2008). Extrinsic mechanisms include interstellar scintillation and gravitational microlensing. Interstellar scintillation causes radio variations at low frequencies (Heeschen et al. 1987) and gravitational microlensing is important on timescales of weeks to months (Agarwal & Gupta 2015). Therefore, we do not consider extrinsic mechanisms to explain the origin of the IDV. For 3C 273 in the optical band, there is the blue-bump emission. Paltani et al. (1998) presented two possible variable components (B and R) to explain the blue-bump emission. The R component could be synchrotron emission not directly related to the radio–millimeter jet (Soldi et al. 2008). The B component could be explained by reprocessing on an accretion disk (Paltani et al. 1998). The results from Paltani et al. (1998) showed that the B component has a maximum timescale of variability of about 2 years (a mean of 0.46 years) and the R component has a longer timescale of variability. Thus, the two variable components cannot explain the origin of the IDV. In the outburst state, the origin of the IDV can be attributed to the shock-in-jet model—relativistic shocks propagating through a relativistic jet of plasma (Marscher & Gear 1985; Marscher et al. 2008). The observed radiation is predominantly non-thermal Doppler-boosted jet emission. The flare arises as disturbances in the flow of a jet cause a shock wave to propagate along the jet (Marscher & Gear 1985). Models based on the instabilities on the accretion disk could give rise to IDV when the blazar is in the low state, because in the low state, any contribution from the jets is weak (Rani et al. 2011). For the quasar, the instabilities on the accretion disk could yield IDV in the modest state because it has an obvious blue-bump emission. From Figure 10, we see that on 2013 April 04 and 2014 April 25, the quasar has the mean flux level, but on 2015 May 17 it  has a low flux level. On 2015 May 17, a large variability amplitude (41.93%) in 5.8 minutes is found. Although the quasar is in a low flux level on this night, it is impossible to use disk-based models to explain the type of variability, because for luminous quasars with black hole masses in excess of 108 M, even the fastest disk-based timescale can be a few hours (Joshi et al. 2011). This type of variability is likely to have an association with jet activities. The shocks propagate down relativistic jets, sweeping emitting regions. If the emitting regions have large intrinsic changes (magnetic field, particle velocity/distribution, a large number of new particles injected), then we could see a large flare on a very short variability timescale. Another possible explanation is the Doppler-factor change of the emitting region. The helical jet structure (Gopal-Krishna & Wiita 1992) may cause the Doppler-factor change on a very short variability timescale. On 2014 April 25, the large intrinsic changes in emitting regions are likely to explain the IDV. On 2013 April 04, the small variability amplitude (∼5%) in 154 minutes can be explained by turbulence behind an outgoing shock along the jet (Agarwal & Gupta 2015) or  hot spots or disturbances in or above accretion disks (e.g., Chakrabarti & Wiita 1993; Mangalam & Wiita 1993; Gaur et al. 2012). The interpretations of IDV for the other nights are similar to that on 2013 April 04. If the observed 154 minutes indicate an innermost stable orbital period from the accretion disk, an upper limit can be obtained for the mass of the central black hole (e.g., Xie et al. 2004b; Fan et al. 2014; Dai et al. 2015). From Fan (2005) and Fan et al. (2014), the innermost stable orbit depends on the black hole and the accretion disk, and $r\leqslant c{\rm{\Delta }}{t}_{\min }/(1+z)$, and then the upper limits of the black hole masses are (i) ${M}_{8}\leqslant 1.2\times \tfrac{{\rm{\Delta }}{t}_{\min }(\mathrm{hr})}{1+z}$ for a thin accretion disk surrounding a Schwarzschild black hole; (ii) ${M}_{8}\leqslant 1.8\times \tfrac{{\rm{\Delta }}{t}_{\min }(\mathrm{hr})}{1+z}$ for a thick accretion disk surrounding a Schwarzschild black hole;  and (iii) ${M}_{8}\leqslant \left(\tfrac{7.3}{1+\sqrt{1-{a}^{2}}}\right)\left(\tfrac{{\rm{\Delta }}{t}_{\min }(\mathrm{hr})}{1+z}\right)$ for the case of a Kerr black hole, where M8 is the black hole mass in units of 108 M and a is an angular momentum parameter. Then we obtain that the upper limits of black hole mass are M8 ≤ 2.66 and M8 ≤ 3.99 for thin and thick accretion disks, and M8 ≤ 16.18 for the extreme Kerr black hole case (a = 1). The previous reverberation mapping results obtained M8 = 2.3–5.5 (Kaspi et al. 2000) and M8 = 65.9 (Paltani & Turler 2005). Therefore, based on the assumption that the observed 154 minutes indicate an innermost stable orbital period, our upper limits on the black hole masses are consistent with reverberation mapping results.

Fan et al. (2009) found that the quasar shows a time delay between the V and I bands. However, the results of our ZDCF analysis indicate that there are no significant time lags between the V-band magnitude and the I-band magnitude. The optical interband time delay supports the shock-in-jet model. For our results, these factors (e.g., sampling rate, time length of observation, flat light curve) can cause the significant time delay to not be found.

4.2. Relation of the Color and Magnitude

Over the 12 years, the overall color index variability is ${\rm{\Delta }}(V-R)=0\buildrel{\rm{m}}\over{.} 25$. The average value of the color index is $\langle V-R\rangle =0\buildrel{\rm{m}}\over{.} 126\pm 0\buildrel{\rm{m}}\over{.} 023$. From our results, the BWB chromatic trend is dominant for 3C 273 and appears at different flux levels for intraday timescales. The BWB trend exists for short-term timescales and intermediate-term timescales but different timescales have different correlations between the VR index and the V magnitude. There is no BWB trend for our whole time-series data sets (12 years). The results from Dai et al. (2009) implied that the quasar has the BWB trend on both intraday and long-term timescales (three years). Ikejiri et al. (2011) found that 3C 273 exhibits the BWB trend in their whole time-series data sets (three years). Fan et al. (2014) found that the spectrum becomes flat when the source becomes bright in the overall trend (about five years). However, there is a steeper-when-brighter trend in the details. When considering the V-band magnitude range from Dai et al. (2009), we still find the BWB chromatic trend in the long-term timescales for our data. When considering the V-band magnitude range from Ikejiri et al. (2011), we find a weak BWB chromatic trend for our data. The difference may be due to the different samples and different color indices—Ikejiri et al. (2011) used VJ while we used VR. From Figure 7 in Fan et al. (2014), there is a gap close to the V-band flux FV = 28 mJy. From our Figure 10, we find that there are some data close to the V-band flux FV = 28 mJy. Then if the new data are included in Figure 7 of Fan et al. (2014), the results of Fan et al. (2014) could change. In addition, we notice that compared with previous results, the time spans of our results are longer. The above discussions can explain the differences between our results and previous results in the long-term timescales.

The BWB behavior is most likely to support the shock-in-jet model. According to the shock-in-jet model, as the shock propagating down the jet strikes a region with a large electron population, radiations at different visible colors are produced at different distances behind the shocks. High-energy photons from the synchrotron mechanism typically emerge sooner and closer to the shock front than the lower frequency radiation, thus causing color variations (Agarwal & Gupta 2015). When we consider two distinct synchrotron components, the BWB trend could be explained if the flare component has a higher Vpeak than the underlying component (Ikejiri et al. 2011). For 3C 273, there is a weak host galaxy contribution, which is a redder non-variable component. Also, the Doppler-boosted jet emission almost invariably swamps the light from the host galaxy. Gravitational microlensing is important on weeks to months timescales and is achromatic. The optical emission may be contaminated by thermal emission from the accretion disk and the surrounding regions, especially for quasars with a blue bump. The quasar 3C 273 has a blue bump, which flattens the spectral slope in the optical region. Since the thermal contribution is larger in the blue region, the composite spectrum would be flatter than the non-thermal component. Then, when the object is brightening, the non-thermal component has a more dominant contribution to the total flux, and the composite spectrum steepens (Gu et al. 2006). Consequently, in the low flux level, the quasar could have a redder-when-brighter (RWB) trend. For a very low flux state where the thermal radiation of the disk dominates the total flux, the accretion disk model can explain the BWB trend (Li & Cao 2008; Gu & Li 2013; Liu et al. 2016). So the BWB trend on intraday timescales is most likely explained by the shock-in-jet model, and is also possibly due to two distinct synchrotron components or the accretion disk model. In addition, we find that there is a significant anticorrelation between the BWB trend and the length of timescales (Figure 9). Sun et al. (2014) found that the color variability of quasars is prominent on timescales as short as ∼10 days, but gradually reduces toward timescales of up to years, i.e., the variable emission on shorter timescales is bluer than that on longer timescales. They proposed the thermal accretion disk fluctuation model to explain the anticorrelation that fluctuations in the inner, hotter region of the disk are responsible for short-term variations, while longer-term and stronger variations are expected from the larger and cooler disk region. The BWB trend can also be interpreted in terms of two components: a "mildly chromatic" longer-term component and a "strongly chromatic" shorter-term one, which can be likely due to Doppler-factor variations on a convex spectrum and intrinsic phenomena, respectively (Villata et al. 2002, 2004; Gu et al. 2006). The long-term achromatic trend could be due to the superposition of different components (jet components, accretion disk components, gravitational microlensing effect; Fan et al. 2008; Ikejiri et al. 2011; Bonning et al. 2012; Agarwal & Gupta 2015; Gaur et al. 2015; Xiong et al. 2016).

4.3. Periodicity

In order to increase the reliability of the periodicity analysis, we appropriately dealt with all of these factors: the best quality data excluding the influence of strong synchrotron flares, three methods for periodicity analysis, unevenly sampled data, red noise, periodicity in year timescales, and periods three times less than total time spans. Converting magnitudes to fluxes in different photometric systems could bring uncertainties that can cause a small (false) variability in the light curve. In this case, if we search for periodicity on short or intraday timescales, the small variability could result in false periodicity. However, the small variability will not cause a significant impact on periodicity in year timescales because the periodicity mainly represents the total change in year timescales. So, analyzing the periodicity in year timescales can reduce the effect from the uncertainties. When analyzing V-band data, a possible quasi-periodicity of P = 3918 ± 1112 days is found. However, it is very possible that a non-periodic model could provide a better fit. Fan et al. (2014) used data over 100 years from the B band (from 1882 to 2012) to analyze the periods of the quasar. Their results obtained possible periods of P = 21.10 ± 0.14, 10.90 ± 0.14, 13.20 ± 0.09, 7.30 ± 0.10, 2.10 ± 0.06, and 0.68 ± 0.05 years for 3C 273. Smith & Hoffleit (1963) suggested a period of P = 12.7–15.2 years for 3C 273. Babadzhanyants & Belokon (1993) analyzed the B-band data of 1887–1991 and found a period of P = 13.4 years. Vol'vach et al. (2013) obtained periods of 11.7 ± 2.5, 7.2 ± 0.8, 4.9 ± 0.3, and 2.8 ± 0.3 years based on the optical B-band light curve of 1963–2011. Therefore, within the errors, the quasi-period of P = 3918 ± 1112 days is consistent with the previous works. For the long-term period of 3C 273, some possible interpretations are as follows: the orbit of a perturbing object, precession of jet, a rotating helical jet structure, and a binary black hole model (Lehto & Valtonen 1996; Villata & Raiteri 1999; Rani et al. 2009; Vol'vach et al. 2013; Fan et al. 2014; Marscher 2014; Sandrinelli et al. 2016).

5. Summary

We have monitored the quasar 3C 273 in the optical V, R, and I bands from 2005 to 2016. In total, the actual number of observations for the quasar is 105 nights, with 1901 I-band, 1707 R-band, and 1544 V-band data points, with rms error less than 0.06 mag. Our main results are the following.

(i) Intraday variability (IDV) is detected on seven nights. The variability amplitudes for most nights are less than 10%, and for four nights, more than 20%. When considering the nights with time spans >4 hr, the value of the duty cycle (DC) is 14.17%. Over the 12 years, on the whole, the quasar becomes dark from 2005 to 2015 but it is also bright between different years. In 2015, the quasar reaches a low flux state. Compared with 2015, in 2016 the quasar had a tendency to brighten. The overall magnitude and color index variabilities are ${\rm{\Delta }}I\,=0\buildrel{\rm{m}}\over{.} 67$, ${\rm{\Delta }}R=0\buildrel{\rm{m}}\over{.} 72$, ${\rm{\Delta }}V=0\buildrel{\rm{m}}\over{.} 68$, and ${\rm{\Delta }}(V-R)=0\buildrel{\rm{m}}\over{.} 25$, respectively.

(ii) The results of fitting the ACF show a possible variability timescale of 0.107 days (154 minutes). A large variability amplitude (41.93%) in 5.8 minutes is found. The type of variability is likely to have an association with jet activities. The light curve of our observation shows a small variability amplitude (∼5%) in 154 minutes, which can be explained by turbulence behind an outgoing shock along the jet or hot spots/disturbances in or above accretion disks. If the observed 154 minutes indicate an innermost stable orbital period, our upper limits of the black hole mass are consistent with reverberation mapping results.

(iii) The BWB chromatic trend is dominant for 3C 273 and appears at different flux levels for intraday timescales. The BWB trend exists for short-term timescales and intermediate-term timescales but different timescales have different correlations. There is no BWB trend for our whole time-series data sets. A significant anticorrelation between the BWB trend and length of timescales is found, which can be explained by the thermal accretion disk fluctuation model. The BWB behavior on intraday timescales is most likely to support the shock-in-jet model. The BWB trend can also be interpreted in terms of the two-component scenario.

(iv) By analyzing V-band data over 48 years, a possible quasi-periodicity of P = 3918 ± 1112 days is found.

(v) No significant time lag between the V and I bands is found on intraday timescales.

We sincerely thank the referee for valuable comments and suggestions. D.R.X. thanks Chuyuan Yang, Yonggang Zheng, Hongtao Liu, and Longhua Qin for useful discussions, and Xuliang Fan, Shaokun Li, Liang Chen, Nenghui Liao, and Jin Zhang for observations. We acknowledge the support of the staff of the Lijiang 2.4 m and Kunming 1 m telescopes. Funding for the two telescopes has been provided by the Chinese Academy of Sciences and the People's Government of Yunnan Province. This work is financially supported by the National Nature Science Foundation of China (11433004, 11663009, 11133006, 11673057, 11361140347, and U1531245), the Key Research Program of the Chinese Academy of Sciences (grant No. KJZD-EW-M06), the Strategic Priority Research Program "The emergence of Cosmological Structures" of the Chinese Academy of Sciences (grant No. XDB09000000), and the Key Laboratory of Yunnan Province University for Research in Astrophysics. D.R.X. acknowledges support from the Chinese Western Young Scholars Program and the "Light of West China" Program provided by CAS. M.F.G acknowledges the supports from the National Science Foundation of China (grant 11473054 and U1531245) and by the Science and Technology Commission of Shanghai Municipality (14ZR1447100). This research has made use of the NASA/IPAC Extragalactic Database (NED), that is operated by Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.

Footnotes

Please wait… references are loading.
10.3847/1538-4365/aa64d2