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

How Far Is Quasar UV/Optical Variability from a Damped Random Walk at Low Frequency?

, , , and

Published 2017 October 3 © 2017. The American Astronomical Society. All rights reserved.
, , Citation Hengxiao Guo et al 2017 ApJ 847 132 DOI 10.3847/1538-4357/aa8d71

Download Article PDF
DownloadArticle ePub

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

0004-637X/847/2/132

Abstract

Studies have shown that UV/optical light curves of quasars can be described using the prevalent damped random walk (DRW) model, also known as the Ornstein–Uhlenbeck process. A white noise power spectral density (PSD) is expected at low frequency in this model; however, a direct observational constraint to the low-frequency PSD slope is difficult due to the limited lengths of the light curves available. Meanwhile, quasars show scatter in their DRW parameters that is too large to be attributed to uncertainties in the measurements and dependence on the variation of known physical factors. In this work we present simulations showing that, if the low-frequency PSD deviates from the DRW, the red noise leakage can naturally produce large scatter in the variation parameters measured from simulated light curves. The steeper the low-frequency PSD slope, the larger scatter we expect. Based on observations of SDSS Stripe 82 quasars, we find that the low-frequency PSD slope should be no steeper than −1.3. The actual slope could be flatter, which consequently requires that the quasar variabilities should be influenced by other unknown factors. We speculate that the magnetic field and/or metallicity could be such additional factors.

Export citation and abstract BibTeX RIS

1. Introduction

The ubiquitous aperiodic variabilities of quasars can be utilized to trace physical information at different scales, especially for the inner accretion disk around the supermassive black hole. With the coming era of time domain surveys (e.g., OGLE, Udalski et al. 1997; SDSS Stripe 82, Sesar et al. 2007; CSS, Drake et al. 2009; PTF, Law et al. 2009; TDSS, Morganson et al. 2015; Pan-STARRS, Kaiser et al. 2002; DES, Dark Energy Survey Collaboration et al. 2016; LSST, Ivezic et al. 2008), quasar variability is gradually attracting more attention.

Measuring the power spectral density (PSD) shapes is one of the major aims of variation studies. An early X-ray timing analysis of a Galactic black hole X-ray binary (BHB)—Cyg X-1—shows that the PSD in the low state is flat below the low-frequency "knee," follows a slope ∼−1 up to another high-frequency "knee," and then steepens to a slope ∼−2. In the high state, only one "knee" has been detected so far, and the slope below and above the break is −1 and around −2 (Belloni & Hasinger 1990; Reig et al. 2002; McHardy et al. 2006; Kelly et al. 2011). Most X-ray power spectra of active galactic nuclei (AGNs) are similar to those of high-state BHB systems, and always only one break is detected (Lawrence et al. 1987; McHardy & Czerny 1987; Papadakis et al. 2002; Markowitz et al. 2003): the slope above the break frequency is suggested to be around −2, and flattens to a shallower slope −1 at low frequencies (Uttley et al. 2002; McHardy et al. 2004; Arévalo et al. 2008). The intrinsic mechanism of the break frequencies is likely to be the characteristic timescale (e.g., dynamical or thermal timescale) ending at the edge of the inner disk region (McHardy 2010).

In UV/optical bands, the story is different as the characteristic break frequency appears much lower than that in X-ray. Recently, Kelly et al. (2009) proposed that quasar optical light curves can be well modeled by the damped random walk (DRW) model, also known as the Ornstein–Uhlenbeck (OU) process (Uhlenbeck & Ornstein 1930), where a self-correcting term is added to a divergent random walk model to push any deviations back toward the mean value. The DRW light curves can be converted to a PSD with slopes ${\alpha }_{l}=0$ and ${\alpha }_{h}=-2$ at low and high frequencies, respectively, and the break frequency can be associated with the thermal timescale in the accretion disk (Kelly et al. 2009; Sun et al. 2015). Various investigations confirm that this model can describe not only the UV/optical light curves of quasars (Kozłowski et al. 2010; MacLeod et al. 2010; Zu et al. 2011), but also those of blazars (Ruan et al. 2012; Sobolewska et al. 2014). Subsequently, the DRW process has been widely adopted in the literature to interpret quasar variabilities (e.g., Andrae et al. 2013; Kelly et al. 2014; Caplar et al. 2016; Charisi et al. 2016; Kozłowski 2016a, 2016b, 2017a, 2017b; Kozłowski et al. 2016). Note that Kepler results with ∼30 min sampling revealed steeper power spectral slopes of −3 at very high frequency, which significantly deviate from the DRW model (Mushotzky et al. 2011; Kasliwal et al. 2015). Zu et al. (2013) concluded that, on a very short timescale (less than a few months) there is a deviation from the DRW model; on timescales from several months to a few years, the DRW model is very consistent with the observed light curves, as seen from SDSS Stripe 82 data (MacLeod et al. 2010, 2012) and well sampled OGLE data (Zu et al. 2011, 2013).

However, on very long timescales (more than a few years), it is yet unclear whether there is significant deviation from the DRW, i.e., whether the very low-frequency PSD has a slope different from 0. The difficulty mainly comes from the limited lengths of the available light curves, which are not sufficient to cover the flat part of the PSD. Furthermore, Emmanoulopoulos et al. (2010) pointed out that spurious breaks often emerge in the structure function of almost all light curves even though they may contain no intrinsic characteristic timescale.5 There are a few AGNs which have light curves spanning as long as several decades or even a century, including NGC 4151 (Guo et al. 2014) and Mrk 421 (Chen et al. 2014), suggesting a non-flat PSD at very low frequency, but with poor photometric data obtained at early times. Additionally, we note that some quasars show dramatic variations in magnitude on very long timescales, including changing-look quasars (e.g., Osterbrock 1977; Cohen et al. 1986; LaMassa et al. 2015; MacLeod et al. 2016; McElroy et al. 2016; Ruan et al. 2016). These suggest that it is unlikely that the variation amplitude of quasars simply ceases to rise at very long timescales, as predicted with the simple DRW process. Statistical constraints to the very low-frequency PSD slopes of quasars are still lacking.

Meanwhile, which physical parameters determine quasar variability is still a riddle. Previous studies have suggested that quasar variability amplitude increases with decreasing luminosity, rest frame wavelength, and Eddington ratio, and increases with increasing black hole mass, while the correlation with redshift is very weak (Wills et al. 1993; Giveon et al. 1999; Vanden Berk et al. 2004; Wold et al. 2007). A fundamental question is: do quasars with the same physical parameters as mentioned above have identical variation properties?

MacLeod et al. (2008) showed that the structure functions of quasars for fixed luminosity, rest frame wavelength, and timescale show very large scatter. Bauer et al. (2011) and MacLeod et al. (2010, 2012) further explored this problem and found that quasars at fixed physical parameters show too large scatter in $\hat{{\sigma }}\,(\hat{{\sigma }}={{\rm{S}}{\rm{F}}}_{{\rm{\infty }}}/\sqrt{{\tau }}$, where ${\mathrm{SF}}_{\infty }$ and τ are the two key parameters of the DRW model) to be attributed to sparse data sampling and photometric uncertainties. They suggested the variability properties of quasars have intrinsic scatter driven by unknown factors, in additional to the aforementioned physical parameters.

However, the red noise leakage could also add a significant scatter in the observed variation parameters of quasars. Red noise leakage refers to the variability power transferring from low to high frequencies by the lobes of the window function, i.e., short light curves display a long-term trend of the variation. It distorts the measurement of PSD and variation amplitude using light curves with limited length. The significance of such leakage depends both on the PSD slope and the length of the observed light curves (Vaughan et al. 2003). Conversely, the observed scatter in the variation parameters can be used to independently constrain the quasar PSD slope at very low frequency, extending beyond the coverage of the available light curves.

In this paper, we focus on the variation parameter $\hat{\sigma }$, which can be much better constrained with observed light curves compared with ${\mathrm{SF}}_{\infty }$ and τ, to restrain the quasar PSD profile at very low frequency. In Section 2, we build a sample of quasars from SDSS Stripe 82, for which SDSS r-band light curves are modeled using the DRW process to derive their $\hat{\sigma }$. We show that the scatter in the measured $\hat{\sigma }$ of the sample (after correcting its dependence to known physical parameters, including luminosity, black hole mass, redshift, etc.) is too large to be recovered with simulated DRW light curves (applying the same data sampling and photometric uncertainties from the real data). In Section 3, we simulate light curves using non-DRW models with non-flat, low-frequency PSD slopes. We show that red noise leakage is able to reproduce the observed scatter, which in turn yields that the low-frequency slope in the PSD should be no steeper than −1.3. A discussion and conclusions are given in Section 4.

2. The Scatter of Quasar Variability Parameters

2.1. SDSS Stripe 82 Quasar Sample and DRW Fitting

SDSS Stripe 82, lying along the celestial equator in the southern Galactic hemisphere, covers an area of ∼290 square degrees, and has been scanned ∼60 times in the ugriz bands by the SDSS imaging survey. MacLeod et al. (2012) presented recalibrated ∼10 year long five-band light curves for 9275 spectroscopically confirmed quasars in the field.6 Measurements of black hole mass, absolute magnitude (K-corrected) and bolometric luminosity for most of them are available, taken from Shen et al. (2008). With these long light curves of such a large sample of quasars, we can extensively study the scatter of quasar variability parameters.

We model the SDSS light curves using the DRW process. The DRW as a stochastic process is well described by the exponential covariance matrix of the signal, with its structure function expressed as (e.g., Hughes et al. 1992; MacLeod et al. 2010)

Equation (1)

with ${\mathrm{SF}}_{\infty }=\sqrt{2}\sigma $, τ the characteristic timescale, and ${\rm{\Delta }}t$ the time lag. Clearly,

Equation (2)

Equation (3)

In reality, measuring τ and ${\mathrm{SF}}_{\infty }$ is often difficult due to the limited light curve length. Instead, $\hat{\sigma }$ (=${\mathrm{SF}}_{\infty }/\sqrt{\tau }$) can be measured with much better accuracy, even when τ and ${\mathrm{SF}}_{\infty }$ are poorly constrained, and can be used as a good proxy to evaluate the properties of quasar variabilities. According to the continuous-time first-order autoregressive process (CAR(1), Kelly et al. 2009), the PSD of a DRW process is given by

Equation (4)

which infers that $P(f)\propto {f}^{-2}$ when $f\gt {(2\pi \tau )}^{-1}$ and P(f) yields a constant when $f\lt {(2\pi \tau )}^{-1}$. In our paper, the nomenclature "DRW" only refers to the CAR(1) process, and "non-DRW" models refer to several special non-CAR(1) models, which are included in a more general model (continuous-time autoregressive moving average, CARMA; Kelly et al. 2014).

In this work, to make a direct comparison with the results of MacLeod et al. (2012), we elect to use the same r-band light curves, which have the best photometries among the five bands for the quasars. We also adopt similar sample selections: (1) only quasars with ≥40 epochs (to ensure reasonable DRW fitting) and measurements of black hole mass and absolute magnitude are retained; (2) ${\rm{\Delta }}{L}_{\mathrm{noise}}=\mathrm{ln}({L}_{\mathrm{best}}/{L}_{\mathrm{noise}})\gt 2$ and ${\rm{\Delta }}{L}_{\infty }=\mathrm{ln}({L}_{\mathrm{best}}/{L}_{\infty })\gt 0.05$, where ${\rm{\Delta }}{L}_{\mathrm{noise}}$, ${\rm{\Delta }}{L}_{\infty }$, ${\rm{\Delta }}{L}_{\mathrm{best}}$ are the likelihoods for τ = 0, $\tau =\infty $, and the best-fit τ (Kozłowski et al. 2010; Zu et al. 2013). The former equation is equivalent to setting a noise limit to request the obtained characteristic timescale larger than the average cadence. The latter eliminates the fitting results that cannot distinguish the best τ and $\tau =\infty $ based on the insufficient light curve length.

We fit SDSS r-band light curves with the Javelin code7 (Zu et al. 2011) to measure the damping timescale τ and asymptotic magnitude ${\mathrm{SF}}_{\infty }$ 8 (see the upper panel of Figure 1 for a demonstration). According to the simulations of Kozłowski (2017a), only if the characteristic timescale (in the observed frame, the same hereafter unless otherwise stated) is less than ∼10% of the light curve length, the input DRW parameters ${\tau }_{\mathrm{in}}$ can be reasonably recovered by DRW fitting, although biased, depending on different priors and fitting method. To address this issue, we only include quasars at lower redshift ($z\lt 1.2$) and with low black hole mass (Log ${M}_{\mathrm{BH}}\lt 9$, see Figure 2) for which the damping timescale is considerably small compared with the light curve length (e.g., Kozłowski 2016b). The final sample includes 1678 quasars. The median value of their observed ${\tau }_{\mathrm{obs}}$ is 301 days, and the typical length of the Stripe 82 light curves is ∼2800 days. The ratio of the typical damping timescale to the light curve duration is 11%.

Figure 1.

Figure 1. Upper panel: r-band light curve of an SDSS Stripe 82 quasar, fitted using the DRW process. Lower panel: using the DRW parameters obtained in the upper panel (shown in the lower left corner), we make an artificial DRW light curve (gray dots). A mock light curve is produced by resampling the simulated DRW light curve and adding photometric errors (red dots and error bars). The best-fit DRW parameter to the mock light curve is also given. The solid black lines in both panels are the DRW models with 1σ "error snakes" (black dashed lines).

Standard image High-resolution image
Figure 2.

Figure 2. Color coding distribution of redshift, black hole mass, and observed ${\tau }_{\mathrm{obs}}$ for all Stripe 82 quasars. Quasars with small ${\tau }_{\mathrm{obs}}$ are mainly located at the lower redshift and black hole mass corner. Only quasars within the dotted box are included in this study, for which τ can be relatively reliably measured with the SDSS Stripe 82 light curves.

Standard image High-resolution image

2.2. The Scatter of $\hat{\sigma }$

In Figure 3 we plot the distributions of the measured ${\tau }_{\mathrm{obs}}$, ${\mathrm{SF}}_{\infty ,\mathrm{ratio}}$, ${\hat{\sigma }}_{\mathrm{ratio}}$, and Kratio for our quasar sample, where $K=\tau \sqrt{{\mathrm{SF}}_{\infty }}$ is a variable orthogonal to $\hat{\sigma }$ in log space. All of them show clear scatters, but slightly smaller than those of MacLeod et al. (2010) as we only include quasars with small black hole mass and redshift, for which τ, ${\mathrm{SF}}_{\infty }$ can be more reliably constrained.

Figure 3.

Figure 3. Distributions of ${\tau }_{\mathrm{ratio}}$, SFratio, ${\hat{\sigma }}_{\mathrm{ratio}}$, Kratio. Solid gray lines plot the scatter of the observed sample. The residual scatters after taking account of dependence to physical factors (Equation (5)) are plotted as blue dashed lines. The uncertainties of DRW fitting (due to the limited length of the light curves, sparse sampling, and photometric errors) are plotted as red dashed lines. The widths of the distributions are measured as inter-quartile ranges, and σ are given in the upper left corners.

Standard image High-resolution image

We revisited the empirical relations given by MacLeod et al. (2010) between the measured DRW parameters (τ, ${\mathrm{SF}}_{\infty }$ and $\hat{\sigma }$), and physical parameters including the i-band absolute magnitude (${M}_{i}$), rest frame wavelength (${\lambda }_{\mathrm{RF}}$), black hole mass (${M}_{\mathrm{BH}}$), and redshift (z). Following their method, we first determined the rest wavelength dependence of the variability parameters (τ, ${\mathrm{SF}}_{\infty }$, and $\hat{\sigma }$) using various SDSS bands, since each individual quasar has other physical parameters fixed. Note we only utilize the $u,g,r$ bands here, and exclude the $i,z$ bands since the weaker intrinsic variabilities and longer intrinsic damping timescales of the latter make the measurements of τ and ${\mathrm{SF}}_{\infty }$ less reliable than in the bluer bands. We then use power laws to fit other physical parameters simultaneously with fixed wavelength coefficients for r band:

Equation (5)

Using our redefined quasar sample, we obtain A = −0.50, B = −0.48, C = 0.13, and D = 0.18 for $f={\mathrm{SF}}_{\infty }$ (mag), A = 2.1, B = 0.66, C = −0.19, and D = 0.14 for f = τ (days, in the rest frame), and A = −1.72, B = −0.86, C = 0.22, and D = 0.12 for $f=\hat{\sigma }$ ($\mathrm{mag}/{\mathrm{day}}^{1/2}$, in the rest frame). All E are fixed to 0, assuming the variability has no evolution with redshift. The typical errors of these coefficients are 0.05. The correlations imply that not only the long/short-term variability but also the damping timescale strongly depend on the rest frame wavelength, luminosity, and black hole mass. Note in Equation (5) that the dependence on ${M}_{i}$ and ${M}_{\mathrm{BH}}$ can be coupled as both quantities correlate with redshift. Caution is thus needed when comparing the coefficients with those reported in other studies (e.g., Kozłowski 2016b, 2017b). Obtaining decoupled coefficients requires samples spanning larger ranges in luminosity and ${M}_{\mathrm{BH}}$, and is beyond the scope of this work. For this study the coefficients we obtained using our sample are sufficient to derive the residual scatters in the DRW parameters (excluding the dependence on the physical parameters listed in Equation (5), see next paragraph).

We then obtain the expected ${\tau }_{\exp }$, ${\mathrm{SF}}_{\infty ,\exp }$, ${\hat{\sigma }}_{\exp }$, and Kexp for each quasar based on the empirical relations. In Figure 3, we plot the distribution of ${\tau }_{\mathrm{obs}}/{\tau }_{\exp }$, ${\mathrm{SF}}_{\infty ,\mathrm{obs}}/{\mathrm{SF}}_{\infty ,\exp }$, ${\hat{\sigma }}_{\mathrm{obs}}/{\hat{\sigma }}_{\exp }$, and Kobs/Kexp respectively, where we can see that the residual scatters (the blue dotted line refers to the scatter after excluding the dependence on the physical parameters listed in Equation (5)), in all four quantities are still prominent. Note that the uncertainties in the measurement of black hole mass contribute little to these residual scatters. For instance, based on Equation (5), a 0.3 dex uncertainty in BH mass can produce a scatter of 0.036 in log($\hat{\sigma }$), which is only 7% of the residual scatter of $\hat{\sigma }$ (${0.036}^{2}/{0.14}^{2}$).

The residual scatters can be at least partially attributed to the uncertainties in the measurement of the DRW parameters, due to limited light curve length, sparse time sampling, and photometric errors. Following MacLeod et al. (2010, 2012), we perform Monte Carlo simulations to quantify such uncertainties. Using the observed ${\tau }_{\mathrm{obs}}$ and ${{\rm{SF}}}_{{\rm{\infty }},{\rm{obs}}}$, we generate a 100 year long artificial DRW light curve spaced every 1 day for each quasar after eliminating the burn-in part (Kelly et al. 2009). A 10 year segment is randomly cut to cross-match with the observed light curve, with the Stripe 82 time sampling and photometric errors imposed at the same time (see the lower panel of Figure 1). We generate one artificial light curve for each quasar in our sample, then fit the artificial light curves again to obtain the output DRW parameter. The ratios of the output to input DRW parameters are over-plotted in Figure 3, from which we can see that the uncertainties in the measurement of the DRW parameters can only account for 41%, 68%, 33%, and 44% of the residual scatter (${\sigma }_{\mathrm{red}}^{2}/{\sigma }_{\mathrm{blue}}^{2}$), for τ, ${\mathrm{SF}}_{\infty }$, $\hat{\sigma }$, and K, respectively. Below we focus on the $\hat{\sigma }$ for which DRW fitting to the simulated light curves recovers the input values with the smallest scatter (red line in the lower left panel of Figure 3), and the largest fraction of the residual scatter (67%) remains unaccounted for.

To better demonstrate how the input the DRW parameters were recovered through fitting the artificial light curves, in Figure 4 we plot the contour distributions of output τ and ${\mathrm{SF}}_{\infty }$ for four different pairs of input values. The input values of τ are 100, 300, 1000, and 3000 days, and the input ${\mathrm{SF}}_{\infty }$ was adjusted to keep $\hat{\sigma }$ unchanged. For each pair of input τ and ${\mathrm{SF}}_{\infty }$, we generate ∼2000 artificial light curves and measure output values for each of them. The contour plots show that the output τ and ${\mathrm{SF}}_{\infty }$ (for each pair of input values) are clearly coupled, following the direction of constant $\hat{\sigma }$. This indicates that the input $\hat{\sigma }$ can be recovered with remarkably high accuracy and small scatter, comparing with τ and ${\mathrm{SF}}_{\infty }$ (see also the red lines in Figure 3). This holds even for those light curves with intrinsic large τ (∼1000 days), for which the ${\tau }_{\mathrm{out}}$ and ${\mathrm{SF}}_{\infty ,\mathrm{out}}$ are poorly constrained with huge scatters. This is also one of the key reasons that we opted to focus on $\hat{\sigma }$ (but not ${\tau }_{\mathrm{out}}$ or ${\mathrm{SF}}_{\infty }$) in this work. Offsets of the median output values from input ones are seen for τ, ${\mathrm{SF}}_{\infty }$ and $\hat{\sigma }$. Such offsets are dependent on the sampling and priors adopted in the DRW fitting. Correcting such offsets will not change the results we present in this work.

Figure 4.

Figure 4. Contour distributions of output $\tau \mbox{--}{\mathrm{SF}}_{\infty }$ for four pairs of input values (with constant input $\hat{\sigma }$). The contours (with red, blue, green, and cyan corresponding to ${\tau }_{\mathrm{in}}=100$, 300, 1000, and 3000 days respectively) show that output τ and ${\mathrm{SF}}_{\infty }$ are tightly coupled, along the direction of constant $\hat{\sigma }$ (gray dashed lines). The input and median output values are marked with filled circles and stars. With increasing input τ, the scatters of output τ and ${\mathrm{SF}}_{\infty }$ increase significantly, but that of $\hat{\sigma }$ shows little change. Deviations of the output median values from the input ones (dependent on the sampling and priors adopted in the DRW fitting) are seen for τ, ${\mathrm{SF}}_{\infty }$, and $\hat{\sigma }$. Correcting such small offsets will not alter the conclusions presented in this work.

Standard image High-resolution image

The intrinsic τ for some of the quasars in our sample may still have been significantly underestimated due to the limited length of the light curves. We perform further simulations to check whether such effect may alter our results. We enlarge the input τ value for each source in our sample by a factor of 2–10, and adjust ${\mathrm{SF}}_{\infty }$ accordingly to keep input $\hat{\sigma }$ unchanged. However, we see no significant difference (<5%) in the scatter of ${\hat{\sigma }}_{\mathrm{out}}/{\hat{\sigma }}_{\mathrm{in}}$, demonstrating that our results are not affected by possible underestimations of τ in our sample. It can also be seen in Figure 4 that, although larger input τ yield much stronger scatter in ${\tau }_{\mathrm{out}}$, the scatter of ${\hat{\sigma }}_{\mathrm{out}}$ appears insensitive to input τ. We also double-check this issue by dividing our sample into two equal-size subsamples according to the BH mass. We find that adopting only the lower-mass sample ($\mathrm{Log}\ {M}_{\mathrm{BH}}\lesssim 8.5$), for which the intrinsic τ should be even smaller, does not alter the results presented above.

3. Can Non-DRW Processes Account for the Residual Scatter?

In Section 2 we showed that the scatter in quasar variability parameter $\hat{\sigma }$ (modeled with the DRW) is too large to be attributed to uncertainties in DRW parameter measurements (due to limited light curve length, sparse sampling, and photometric errors). Monte Carlo simulations show that such an effect can only account for 33% of the residual scatter in $\hat{\sigma }$ (the observed scatter after excluding its dependence on the physical parameters in Equation (5)). What is the origin of the remaining dominant fraction of the scatter in $\hat{\sigma }$?

MacLeod et al. (2010) suggested that this could be due to intrinsic scatter in quasar variability controlled by some unknown factors, that is, for quasars with the same BH mass, luminosity, redshift, and at the same wavelength, their variation properties could be distinct. While it is yet unknown what extra factors control quasar variability, in this work we point out that red noise leakage, if quasar variability deviates from the DRW at very low frequency, could produce extra scatter in the observed variability parameters. Consequently, we develop an approach to place an independent constraint on the low-frequency end PSD slope of quasars.

3.1. Simulating Non-DRW Variation

Non-DRW light curves are simulated from a broken power-law PSD with the public Python code pyLCSIM,9 based on the algorithm of Timmer & Koenig (1995). Two key variables are needed in this approach: one is the break frequency ${(2\pi \tau )}^{-1}$, where τ is the damping timescale; the other is the rms variability amplitude ($\mathrm{rms}=\sigma ={\mathrm{SF}}_{\infty }/\sqrt{2}$ for DRW). As a starting point, we use the observed ${(2\pi {\tau }_{\mathrm{obs}})}^{-1}$ and ${\mathrm{SF}}_{\infty ,\mathrm{obs}}/\sqrt{2}$ (where ${\tau }_{\mathrm{obs}}$ and ${\mathrm{SF}}_{\infty ,\mathrm{obs}}$ are the best-fit DRW parameters) of each quasar as input values to generate 100 year long artificial light curves spaced every 1 day. The slopes of the broken power-law model can be arbitrarily decided. We investigate four non-DRW models comparing with the standard DRW model: setting low-frequency slopes ${\alpha }_{l}=-1$ (model A), −1.9 (model B) with fixed high-frequency slopes at ${\alpha }_{h}=-2$; and ${\alpha }_{h}=-1.8$ (model C), −3 (model D) with fixed ${\alpha }_{l}=0$ (see Figure 5).

Figure 5.

Figure 5. Demonstration of DRW and non-DRW models.

Standard image High-resolution image

Figure 6 briefly illustrates the deviation of non-DRW models from the DRW one. The first three panels show three 100 year light curves produced by three models with the same input values (rms = 0.1 mag, τ = 10 days ), with ${\alpha }_{h}$ fixed at −2 and slopes ${\alpha }_{l}=0$ (DRW), −1 (model A), and −1.9 (model B), respectively. In the last panel, the corresponding power density spectra are plotted. Note the pyLCSIM code generates light curves with rms (measured through the whole light curve) simply fixed at the input value. Short segments of such light curves have smaller rms for red noise PSD, but with dispersion in rms, reflecting red noise leakage. We find that selecting a 10 year segment out of a 100 year long simulated light curve is sufficient to reproduce the effect of red noise leakage (for the PSD we chose), and simulating even longer light curves is not needed. For the DRW process, the variation is white noise at timescales longer than τ, thus there is no red noise leakage. On the other hand, utilizing the DRW model to fit the red noise PSD will misestimate the τ and ${\mathrm{SF}}_{\infty }$, as shown in the last panel of Figure 6 (for the case of ${\alpha }_{l}=-1$).

Figure 6.

Figure 6. Examples of our simulated DRW and non-DRW light curves (100 year long, with ${\mathrm{SF}}_{\infty ,\mathrm{in}}=0.14$ mag and ${\tau }_{\mathrm{in}}=10$ days). Their corresponding power density spectra are plotted in the last panel (solid lines), where the colored dots are the predicted power spectra with random phase and Poisson noise. The red dashed line shows the systematic bias in fitting non-DRW variation (red solid line) with DRW. The two vertical black dashed lines mark the break frequencies of the red dashed and solid line, and arrows infer the biases (${\rm{\Delta }}\tau $ and ${\rm{\Delta }}\mathrm{SF}$). The PSD unit in the last panel is arbitrary.

Standard image High-resolution image

In Figure 7 we demonstrate the discrepancies of the output τ and ${\mathrm{SF}}_{\infty }$ from the input ones used to simulate artificial light curves based on various models. For each set of input parameters (shown in Figure 7 with fixed τ = 300 days and ${\mathrm{SF}}_{\infty }$ changing from 0.01 to 0.64 mag, or fixed ${\mathrm{SF}}_{\infty }=0.18$ mag and τ changing from 50 to 3200 days), we generate 10,000 artificial resampled light curves and plot the averaged output parameters versus the input ones. From Figure 7 we see that for the DRW model, the input parameters are well recovered (with only slight offsets), except that ${\tau }_{\mathrm{out}}$ is significantly underestimated at ${\tau }_{\mathrm{in}}$ > 1000 days (due to the limited length of the light curves), consistent with the results of Kozłowski (2017a). For the non-DRW models, only models C and D yield ${\mathrm{SF}}_{\infty ,\mathrm{out}}$ consistent with input values. Significant deviations from input τ are seen for all non-DRW models. For model B, particularly, the output τ barely correlates with the input value. This is because the PSD break in model B is so weak that modeling it with the DRW process somehow makes no sense (imaging modeling pure red noise PSD slope of −2 with a DRW process).

Figure 7.

Figure 7. Best-fit DRW parameters of the simulated light curves vs. the input values. In the left panel, input τ is fixed at 300 days with ${\mathrm{SF}}_{\infty }/\sqrt{2}$ ranging from 0.01 to 0.64 mag (${\mathrm{SF}}_{\infty }=\sqrt{2}\sigma $), while in the the right panel ${\mathrm{SF}}_{\infty }$ is set to 0.18 mag with τ ranging from 50 to 3200 days.

Standard image High-resolution image

To correct the aforementioned biases (mismatches between input and output τ and ${\mathrm{SF}}_{\infty }$), we perform simulations for a grid of input parameters (τ ranging from 1 to 105 days and ${\mathrm{SF}}_{\infty }$ from 0.01 to 0.8 $\mathrm{mag}$). The output parameters (averaged over a large number of simulations with the same input parameters) build a ${\tau }_{\mathrm{out}}\mbox{--}{\mathrm{SF}}_{\infty ,\mathrm{out}}$ surface. Finding the position of each quasar on the ${\tau }_{\mathrm{out}}\mbox{--}{\mathrm{SF}}_{\infty ,\mathrm{out}}$ plane, we derive the corresponding ${\tau }_{\mathrm{in}}$ and ${\mathrm{SF}}_{\infty ,\mathrm{in}}$ to be input for simulations.

3.2. Outputs of Non-DRW Models

In this section we examine whether non-DRW models could account for the residual scatter in $\hat{\sigma }$. Comparing with the DRW, models A and B (deviation from the DRW at low frequency: ${\alpha }_{l}=-1$, −1.9 instead of 0, with ${\alpha }_{h}$ fixed at −2) produce obviously larger scatters in ${\hat{\sigma }}_{\mathrm{out}}/{\hat{\sigma }}_{\mathrm{in}}$ (Figure 8). This is because of red noise leakage which refers to the variability power transferred from low to high frequencies (Uttley et al. 2002; Zhu & Xue 2016). This effect is more significant for PSDs with steeper slopes at low frequencies. An offset from unity of the ${\hat{\sigma }}_{\mathrm{out}}$/${\hat{\sigma }}_{\mathrm{in}}$ distribution is seen for model B. As explained in Section 3.1, this is because the PSD break in model B is so weak that modeling it with the DRW yields an output τ insensitive to the input value, thus our corrections in Section 3.1 are not sufficient. Nevertheless, this does not affect the major conclusions of this work.

Figure 8.

Figure 8. Epected scatter in $\hat{\sigma }$ from our non-DRW models (A and B), compared with that from the DRW model.

Standard image High-resolution image

We repeat our simulations for different low-frequency slopes. In Figure 9, we plot the reproduced scatter in ${\hat{\sigma }}_{\mathrm{out}}$/${\hat{\sigma }}_{\mathrm{in}}$ versus the low-frequency PSD slope, where we see a clear trend that a steeper lower-frequency PSD yields a stronger scatter in $\hat{\sigma }$. Simulating and modeling quasar light curves with the DRW provides a much lower scatter in $\hat{\sigma }$ ($0.08\ \mathrm{mag}/{\mathrm{yr}}^{1/2}$) as compared to empirically measured values (0.14 $\mathrm{mag}/{\mathrm{yr}}^{1/2}$). A non-DRW model with low-frequency slope ${\alpha }_{l}\simeq -1.3$ appears just sufficient to explain the observed residual scatter in $\hat{\sigma }$.

Figure 9.

Figure 9. Expected scatter in $\hat{\sigma }$ decreases with increasing ${\alpha }_{l}$ (0, DRW, −1, Model A, and −1.9 Model B). The horizontal red dashed line marks the observed residual scatter in the SDSS Stripe 82 quasars (see the blue line in the lower left panel of Figure 3). The different models are marked. The DRW model accounts for a much lower scatter in $\hat{\sigma }(0.08\,\mathrm{mag}/{\mathrm{yr}}^{1/2}$) as compared to the empirically measured values ($0.14\,\mathrm{mag}/{\mathrm{yr}}^{1/2}$, red line). The vertical black dashed line suggests the lower limit of ${\alpha }_{l}$.

Standard image High-resolution image

When modeling the SDSS stripe 82 light curves, we obtained a runaway fraction (RF) of ${\mathrm{RF}}_{\mathrm{obs}}=19$%, i.e., for 19% of the light curves τ cannot be well constrained (output τ hits the boundary at τ = 105 days). The RF was also measured for our simulations: ${11}_{-2}^{+2}$% for the DRW model, ${16}_{-2}^{+2}$% for model A, and ${40}_{-3}^{+3}$% for model B, ${6}_{-1}^{+1}$% for model C and ${28}_{-2}^{+2}$% for model D. The errors in RF are measured as scatters from different sets of simulations. The observed RF appears more consistent with model A, also suggesting the low-frequency PSD slope is likely ${\alpha }_{l}\sim -1$, deviating from the DRW.

For model C and D (${\alpha }_{h}=-1.8$, −3 instead of −2, with ${\alpha }_{l}$ fixed at 0), as there is no red noise leakage; both models yield ${\hat{\sigma }}_{\mathrm{out}}/{\hat{\sigma }}_{\mathrm{in}}$ scatter similar to that of the DRW model (see Figure 10).

Figure 10.

Figure 10. Same as Figure 8, but for models C and D.

Standard image High-resolution image

4. Discussion and Conclusions

While direct measurement of the low-frequency PSD slope is difficult for quasars (which require much longer light curves), our simulations in this work yield a lower limit (below the break, ${\alpha }_{l}\gt -1.3$). The analyses of the RF (the fraction of quasar light curves for which the characteristic timescale τ cannot be well constrained) also suggest a red noise low-frequency PSD slope (${\alpha }_{l}\sim -1.0$, also see MacLeod et al. 2010),10 instead of white noise predicted by the DRW model. Similarly, low-frequency PSD slopes of −1.0 to −1.2 (on average, with considerably large scatter) were obtained with Pan-STARRS1 light curves (Simm et al. 2016). Deviation from the DRW at low frequency can also be seen in the structure function plots presented in MacLeod et al. (2012), which include Palomar archive data to extend the line base up to 50 years for SDSS quasars. A red noise low-frequency PSD also appears consistent with that seen in X-ray (see Zheng et al. 2017, in preparation, for X-ray PSDs obtained based on CDF-S light curves spanning over 16 years). We note a second break is needed at even lower frequency as the total variability power needs to stay finite.

Long-term variation in quasars could be due to changes of global accretion rate (e.g., Li & Cao 2008), or thermal fluctuation of the disk (see Dexter & Agol 2011; Cai et al. 2016, for a promising inhomogeneous disk model).

The low-frequency PSD, deviating from the single DRW process, could be due to the combination of different DRW processes with various damping timescales (Kelly et al. 2011; Dobrotka et al. 2017). Additionally, based on fluctuations of the accretion rate, Lyubarskii (1997) proposed an accretion disk model with an α parameter fluctuating at different radii, which indicates a roughly flicker-noise-type PSD ($P\propto {f}^{-1\sim -2}$) with a characteristic timescale approximating the order of the local viscous timescales. Moreover, to account for the intriguing PSD features in Cyg X-1, Takeuchi et al. (1995) proposed a cellular-automaton model for the accretion disk based on the concept of self-organized criticality. In this model, mass accretion takes place in the form of avalanches only when the local mass density exceeds a critical value, which can produce a ${f}^{-1.5}$-like PSD.

Note that the low-frequency slope of −1.3 we obtained in this work is a lower limit. If the PSD low-frequency slope is actually flatter, an interesting consequence is that there should be other factors (in additional to those aforementioned) which control the variation amplitudes of quasars. We speculate that such factors may include metallicity and magnetic field. Recent 3D radiation magnetohydrodynamic simulations suggest that an iron opacity bump may have a strong impact on the thermal stability of an accretion disk (Jiang et al. 2016), pointing to a possible link between metallicity and disk stability. Furthermore, the radiative cooling rate increases with increasing metallicity in hot plasma (Boehringer & Hensler 1989); thus inhomogeneous thermal instability may reduce with increasing metallicity in quasars. However, direct observational evidence of such a link is still absent.

As regards the magnetic field, recent theoretical works have suggested that magnetically driven winds can greatly reduce the disk temperature and help the disk become more stable at a given accretion rate (e.g., Li & Begelman 2014). Moreover, there are studies claiming that optically thick, geometrically thin accretion disks with strong toroidal magnetic field are stable against thermal and viscous instabilities (Begelman & Pringle 2007; Oda et al. 2009). Sadowski (2016a, 2016b) also indicated that magnetic-pressure-dominated thin accretion disks can maintain thermal equilibrium, in contrast to thermally unstable disks dominated by radiation pressure. Therefore, a link between magnetic field strength and quasar variability is also expected.

The main conclusions of this work are as follows.

  • (1)  
    We model 1678 low-redshift and low black hole mass quasar light curves of SDSS Stripe 82 quasars using the DRW process. Quasars show significant residual scatter in DRW parameters, even after correcting for their dependence on known physical parameters, including luminosity, black hole mass, redshift, and rest frame wavelength.
  • (2)  
    With simulated DRW light curves, we show that accurately measuring τ and ${\mathrm{SF}}_{\infty }$ (two key DRW parameters) for an individual quasar using SDSS Stripe 82 light curves is difficult, with huge scatter and strong bias, for sources with intrinsically large τ (≳1000 days in the observed frame), consistent with previous studies. However, $\hat{\sigma }$ ($={\mathrm{SF}}_{\infty }/\sqrt{\tau }$) can be recovered with remarkably small scatter (for τ up to 3000 days, comparable to the length of the light curves).
  • (3)  
    The observed residual scatter in $\hat{\sigma }$ of our quasar sample is too large to be attributed to uncertainties in the DRW parameter measurements caused by limited light curve length, sparse sampling, and photometric errors.
  • (4)  
    We show using simulations that red noise leakage, if the PSD deviates from the DRW below the break frequency, can produce large residual scatter in $\hat{\sigma }$.A low-frequency PSD slope of −1.3 is required to match the observed scatter in $\hat{\sigma }$, indicating that the actual low-frequency slope should be no steeper than −1.3.
  • (5)  
    If the actual low-frequency PSD slope is flatter than −1.3, there must be other factors (in additional to the known ones) which affect the variation amplitudes of quasars, to explain the observed scatter in $\hat{\sigma }$. Two candidates are magnetic field and metallicity.

We especially thank the anonymous referee and AAS Journal statistician for their helpful comments and suggestions that have significantly improved the paper. We thank Chelsea Macleod, Ying Zu, Riccardo Campana, and Xinwu Cao for valuable discussions. This work is supported by the National Basic Research Program of China (973 program, grant No. 2015CB857005), CAS Frontier Science Key Research Program QYCDJ-SSW-SLH006, and National Science Foundation of China (grants No. 11233002, 11421303, and 11503024). J.X.W. is grateful for the support from the Chinese Top-notch Young Talents Program. Z.Y.C. acknowledges support from the Fundamental Research Funds for the Central Universities. M.Y.S. acknowledges support from the China Postdoctoral Science Foundation (grant No. 2016M600485) and the National Science Foundation of China (grant No. 11603022).

Funding for the Sloan Digital Sky Survey I and II has been provided by the Alfred P. Sloan Foundation, the U.S. Department of Energy Office of Science, and the Participating Institutions. SDSS-IV acknowledges support and resources from the Center for High-Performance Computing at the University of Utah. The SDSS web site is www.sdss.org.

SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions of the SDSS Collaboration including the Brazilian Participation Group, the Carnegie Institution for Science, Carnegie Mellon University, the Chilean Participation Group, the French Participation Group, Harvard-Smithsonian Center for Astrophysics, Instituto de Astrofísica de Canarias, The Johns Hopkins University, Kavli Institute for the Physics and Mathematics of the Universe (IPMU)/University of Tokyo, Lawrence Berkeley National Laboratory, Leibniz Institut für Astrophysik Potsdam (AIP), Max-Planck-Institut für Astronomie (MPIA Heidelberg), Max-Planck-Institut für Astrophysik (MPA Garching), Max-Planck-Institut für Extraterrestrische Physik (MPE), National Astronomical Observatories of China, New Mexico State University, New York University, University of Notre Dame, Observatário Nacional/MCTI, The Ohio State University, Pennsylvania State University, Shanghai Astronomical Observatory, United Kingdom Participation Group, Universidad Nacional Autónoma de México, University of Arizona, University of Colorado Boulder, University of Oxford, University of Portsmouth, University of Utah, University of Virginia, University of Washington, University of Wisconsin, Vanderbilt University, and Yale University.

Footnotes

  • This has little effect on statistical work.

  • Note we assume a uniform prior on the initial input parameters since this non-informative prior is more judicial for unknown parameters, while some other studies favor logarithmic (MacLeod et al. 2010) or log-uniform priors (Charisi et al. 2016).

  • 10 

    However, as we have shown in Section 3, the RF is also sensitive to the high-frequency PSD slope, thus the constraint to ${\alpha }_{l}$ based on RF analyses alone is rather loose. Contrarily, the high-frequency slope has little effect on the scatter of $\hat{\sigma }$.

Please wait… references are loading.
10.3847/1538-4357/aa8d71