Hostname: page-component-cb9f654ff-w5vf4 Total loading time: 0 Render date: 2025-08-23T11:18:16.578Z Has data issue: false hasContentIssue false

Searching for quasi periodic oscillations in optical and gamma-ray emissions and black hole mass estimation of blazar ON 246

Published online by Cambridge University Press:  28 July 2025

Ajay Sharma
Affiliation:
S. N. Bose National Centre for Basic Sciences, Kolkata, India
Raj Prince
Affiliation:
Department of Physics, Institute of Science, Banaras Hindu University, Varanasi, Uttar Pradesh, India
Debanjan Bose*
Affiliation:
Department of Physics, Central University of Kashmir, Ganderbal, India
*
Author for correspondence: Debanjan Bose, Email: debaice@gmail.co
Rights & Permissions [Opens in a new window]

Abstract

We report the detection of a potential quasi-periodic signal with a period of $\sim$2 yr in the blazar ON 246, based on Fermi-LAT ($\gamma$-rays) and ASAS-SN (optical) observations spanning 11.5 yr (MJD 55932–60081). We applied various techniques to investigate periodic signatures in the light curves, including the Lomb-Scargle periodogram (LSP), weighted wavelet Z-transform (WWZ), and REDFIT. The significance of the signals detected in LSP and WWZ was assessed using two independent approaches: Monte Carlo simulations and red noise modelling. Our analysis revealed a dominant peak in the $\gamma$-ray and optical light curves, with a significance level exceeding 3$\sigma$ in both LSP and WWZ, consistently persisting throughout the observation period. Additionally, the REDFIT analysis confirmed the presence of a quasi-periodic signal at $\sim$0.00134 day$^{-1}$ with a 99$\%$ confidence threshold. To explain the observed quasi-periodic variations in $\gamma$-ray and optical emissions, we explored various potential physical mechanisms. Our analysis suggests that the detected periodicity could originate from a supermassive binary black hole (SMBBH) system or the jet-induced orbital motion within such a system. Based on variability characteristics, we estimated the black hole mass of ON 246. The study suggests that the mass lies within the range of approximately $(0.142 - 8.22) \times 10^9$ M$_{\odot}$.

Information

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike licence (https://creativecommons.org/licenses/by-nc-sa/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the same Creative Commons licence is used to distribute the re-used or adapted article and the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Astronomical Society of Australia

1. Introduction

Active galactic nuclei (AGNs) are one of the most energetic astrophysical objects in the universe, powered by the accretion of matter of galaxy on supermassive black hole (SMBHs) with a mass in the range $10^6$ $10^{10}\,\textrm{M}_{\odot}$ (Sołtan Reference Sołtan1982). Blazars, a subset of radio-loud AGNs, are among the most luminous objects with the bolometic luminosity in the range of $10^{41}$ $10^{48}$ erg s $^{-1}$ . Blazars produce relativistic jets that are aligned within a few degrees ( $\lt5^{\circ}$ ) of our line of sight (Ghisellini et al. Reference Ghisellini, Padovani, Celotti and Maraschi1993; Megan & Padovani Reference Megan and Padovani1995; Blandford, Meier, & Readhead Reference Blandford, Meier and Readhead2019) and emit radiation over entire electromagnetic (EM) spectrum, from radio to very high energy (VHE; $\gt$ 100 GeV) $\gamma$ -rays (Megan & Padovani Reference Megan and Padovani1995; Ulrich, Maraschi, & Urry Reference Ulrich, Maraschi and Urry1997; Padovani Reference Padovani2017). Blazars are further classified into two subclasses: BL Lacertae (BL Lacs) and flat-spectrum radio quasars (FSRQs), based on the characteristics and strength of broad emission lines in their optical spectra. BL Lacs exhibit featureless nonthermal optical spectra (very weak or absence of the lines), while the FSRQs show bright and strong broad emission lines with EW $\gt 5 \ \mathring{A}$ in the rest frame (Giommi et al. Reference Giommi, Padovani, Polenta, Turriziani, D’Elia and Piranomonte2012).

Observational studies have shown that these sources exhibit rapid and large flux modulations across the entire band of the EM spectrum from radio to VHE $\gamma$ -ray, jet-dominated nonthermal emission that leads to the double-humped spectral energy distribution (Ulrich, Maraschi, & Urry Reference Ulrich, Maraschi and Urry1997; Fossati et al. Reference Fossati, Maraschi, Celotti, Comastri and Ghisellini1998). The variation in emission provides valuable insights into various aspects of blazars, including structures, underlying emission mechanisms/processes, and physical parameters of SMBHs (Ulrich, Maraschi, & Urry Reference Ulrich, Maraschi and Urry1997; Gupta Reference Gupta2017). The observed flux variability timescale in all bands ranges from minutes to several years. Blazar’s central region is very compact and often difficult to resolve directly with current facilities. By analyzing the rapid variability with timescale from minutes to hours, one can constrain the emission-region sizes of these sources effectively, and utilising the simultaneous observations with theoretical models, we can constrain the physical parameters of jets (Blandford & McKee Reference Blandford and McKee1982; Tavecchio, Maraschi, & Ghisellini Reference Tavecchio, Maraschi and Ghisellini1998; Li et al. Reference Li, Xia, Liang, Liao and Fan2018; Pandey & Stalin Reference Pandey and Stalin2022).

Observations have shown that blazar’s flux variations are stochastic, nonlinear, and aperiodic in nature and well characterised by the simplest model of Continuous Autoregressive Moving Average [CARMA(p,q)] (Kelly, Bechtold, & Siemiginowska Reference Kelly, Bechtold and Siemiginowska2009), also known as red noise model, but a small percentage of sources in the entire blazar population exhibit regular variations in the light curves, such particular phenomenon is known as quasi-periodic oscillation (QPO), which appear to be rare in AGNs. Such kinds of regular variations have been observed across the entire EM spectrum with the diverse timescales, ranging from minutes through months to years (Urry et al. Reference Urry1993; Wagner & Witzel Reference Wagner and Witzel1995; Petry et al. Reference Petry2000; Katarzyński, Sol, & Kus 2001; Aleksić et al. 2011; Sandrinelli, Covino, & Treves Reference Sandrinelli, Covino and Treves2014; Carnerero et al. Reference Carnerero2017; Sarkar et al. Reference Sarkar2019; Raiteri et al. Reference Raiteri2021a; Sobolewska et al. Reference Sobolewska, Siemiginowska, Kelly and Nalewajko2014; Gupta et al. Reference Gupta2008; Gupta et al. Reference Gupta2019; Mao & Zhang Reference Mao and Zhang2024). The observed diverse timescale of QPOs may be associated with different underlying physical mechanisms and radiation processes in blazars. Intraday variability with timescales ranging from minutes to several hours may be originated via rotating inhomogeneous helical jet or current-driven kink instabilities (Raiteri et al. Reference Raiteri2021a,b; Jorstad et al. Reference Jorstad2022). A short-term variability with a timescale ranging from days to a few months is believed to originate from the helical motion of magnetised plasma blob within the jet or perturbation in accretion disk at the innermost stable circular orbit (Zhou et al. Reference Zhou, Wang, Chen, Wiita, Vadakkumthani, Morrell, Zhang and Zhang2018; Gupta et al. Reference Gupta2019; Sarkar et al. Reference Sarkar, Gupta, Chitnis and Wiita2021; Roy et al. Reference Roy, Sarkar, Chatterjee, Gupta, Chitnis and Wiita2022; Banerjee et al. Reference Banerjee, Sharma, Mandal, Kumar Das, Bhatta and Bose2023; Prince et al. Reference Prince2023; Sharma et al. Reference Sharma, Kamaram, Prince, Khatoon and Bose2024; Tantry et al. Reference Tantry, Sharma, Shah, Iqbal and Bose2025). A long-term variability of timescale from several months to years may be associated with supermassive binary black hole (SMBBHs; Begelman, Blandford, & Rees Reference Begelman, Blandford and Rees1980) systems or jet structures. This interpretation has been adopted in several studies, e.g., PG 1302-102 (Graham et al. Reference Graham2015), Sandrinelli et al. (Reference Sandrinelli, Covino, Dotti and Treves2016) included PKS 0537–441, OJ 287, 3C 379, PKS 1510–089, PKS 2005–489, and PKS 2155–304 in the study, PKS 0301-243 (Zhang et al. Reference Zhang, Yan, Zhou, Fan, Wang and Zhang2017), OT 081 (Li et al. Reference Li, Cai, Yang, Luo, Yan, He and Wang2021), PKS J0805–0111 (Ren et al. Reference Ren, Zhang, Zhang, Ding, Yang, Li, Yan and Xu2021b), PKS J2134–0153 (Ren et al. Reference Ren, Ding, Zhang, Xue, Zhang, Xiong, Li and Li2021a), S5 0716+714 (Haiyan et al. Reference Haiyan, Xiefei, Xiaopan, Na, Haitao, Yuhui, Li and Yan2023), PKS 1510-089 (Li et al. Reference Li2023). The two most promising candidates, PG 1553+113 and OJ 287, have been reported for hosting an SMBBH system (Sillanpaa et al. Reference Sillanpaa, Haarala, Valtonen, Sundelius and Byrd1988; Valtonen et al. Reference Valtonen, Lehto, Takalo and Sillanpää2011; Ackermann et al. Reference Ackermann2015; Tavani et al. Reference Tavani, Cavaliere, Munar-Adrover and Argan2018; Adhikari et al. Reference Adhikari, Penil, Westernacher-Schneider, Dominguez, Ajello, Buson, Rico and Zrake2024). The $\gamma$ -ray emission in blazars originates from their relativistic jets. Investigating quasi-periodic variations in $\gamma$ -ray emissions not only deepens our understanding of jet physics but also provides insights into particle acceleration mechanisms and jet dynamics. This has become possible due to the continuous monitoring capability of Fermi Large Area Telescope (Fermi-LAT). Leveraging long-term observations from Fermi-LAT, numerous strong QPOs in the $\gamma$ -ray band have been reported in the literature. Additionally, in recent decades, systematic searches have been performed for QPOs in different wavebands for a number of sources (Gierliński et al. 2008; Gupta et al. Reference Gupta2008; Gupta, Srivastava, & Wiita Reference Gupta, Srivastava and Wiita2008; Lachowicz et al. Reference Lachowicz, Gupta, Gaur and Wiita2009; King et al. Reference King2013; Alston et al. Reference Alston, Parker, Fabian and Kara2015; Bhatta et al. Reference Bhatta2016; Pan et al. Reference Pan, Yuan, Yao, Zhou, Liu, Zhou and Zhang2016; Bhatta & Dhital Reference Bhatta and Dhital2020; Peñil et al. Reference Peñil2020; Bhatta Reference Bhatta2021; Ren et al. Reference Ren, Zhang, Zhang, Ding, Yang, Li, Yan and Xu2021b; Yang et al. Reference Yang, Yan, Zhang, Dai and Zhang2021; Wang, Cai, & Fan Reference Wang, Cai and Fan2022; Gong et al. Reference Gong, Zhou, Yuan, Zhang, Yi and Fang2022b; Gong et al. Reference Gong, Tian, Zhou, Yi and Fang2023; Otero-Santos et al. Reference Otero-Santos, Peñil, Acosta-Pulido, Becerra González, Raiteri, Carnerero and Villata2023; Lu et al. Reference Lu, Sun, Fang, Wan and Gong2024; Ren, Sun, & Zhang Reference Ren, Sun and Zhang2024). In recent years, some QPO studies with high statistical significance have been reported in different EM bands. For instance, Tripathi et al. (Reference Tripathi, Gupta, Aller, Wiita, Bambi, Aller and Gu2021) reported in radio, Roy et al. (Reference Roy, Sarkar, Chatterjee, Gupta, Chitnis and Wiita2022) in optical, Smith, Oramas, & Perlman (Reference Smith, Oramas and Perlman2023), and Gupta et al. (Reference Gupta2019) in $\gamma$ -ray.

Additionally, some reported QPOs have also been interpreted by other geometrical models, such as pulsating accretion flow instability, jet precession, and Lense-Thirring precession of accretion disks (Romero et al. Reference Romero, Chajet, Abraham and Fan2000; Rieger Reference Rieger2005; Stella & Vietri Reference Stella and Vietri1997). Apart from the long-term persistent QPOs, several transient QPOs have also been reported (Zhou et al. Reference Zhou, Wang, Chen, Wiita, Vadakkumthani, Morrell, Zhang and Zhang2018; Benkhali et al. Reference Benkhali, Hofmann, Rieger and Chakraborty2020; Peñil et al. Reference Peñil2020; Das et al. Reference Das, Prince, Gupta and Kushwaha2023; Prince et al. Reference Prince2023; Ren, Cerruti, & Sahakyan Reference Ren, Cerruti and Sahakyan2023; Sharma et al. Reference Sharma, Kamaram, Prince, Khatoon and Bose2024). The physical mechanisms of transient QPOs have been attributed to the orbiting hotspots on the disks, or close to the innermost stable circular orbits, magnetic reconnection within the jet, and helical orbital motion of blobs in the jet under the influence of magnetic field (Zhang & Bao Reference Zhang and Bao1990; Mangalam & Wiita Reference Mangalam and Wiita1993; Gupta et al. Reference Gupta2008; Gupta, Srivastava, & Wiita Reference Gupta, Srivastava and Wiita2008; Gupta et al. Reference Gupta2019; Huang et al. Reference Huang, Wang, Wang and Wang2013; Mohan & Mangalam Reference Mohan and Mangalam2015). Thus, QPO studies play a crucial role in understanding the origin of such variations, the underlying radiation mechanisms, and the physical properties of SMBH systems.

The mass of a black hole is one of the most fundamental parameters as it plays a key role in shaping its emission properties and evolutionary behaviour. In literature, several methods have been proposed to estimate the black hole mass: (1) the reverberation mapping technique (Kaspi et al. Reference Kaspi, Smith, Netzer, Maoz, Jannuzi and Giveon2000), (2) single-epoch spectral measurements/broad line width technique (Vestergaard Reference Vestergaard2002), (3) the gas and stellar dynamics technique (Genzel et al. Reference Genzel, Eckart, Ott and Eisenhauer1997), and (4) the variability timescale technique (Fan, Xie, & Bacon Reference Fan, Xie and Bacon1999; Cheng, Fan, & Zhang Reference Cheng, Fan and Zhang1999; Fan Reference Fan2005; Fan et al. Reference Fan2009; Yang & Fan Reference Yang and Fan2010; Liu & Bai Reference Liu and Bai2015) and reference therein. Pei et al. (Reference Pei, Fan, Yang, Huang and Li2022) derived the black hole mass of the blazar ON 246 to be $\sim8.08 \times 10^7 \,{\rm M}_{\odot}$ based on certain assumptions, including the variability timescale of $\sim$ 1 day and a low Doppler factor of 0.48.

The strong radio source S3 1227+25 (Pauliny-Toth et al. Reference Pauliny-Toth, Kellermann, Davis, Fomalont and Shaffer1972) at R.A.=187 $^{\circ}$ .560, decl.=25 $^{\circ}$ .298, also known as ON 246 (Dixon & Kraus Reference Dixon and Kraus1968), was first identified as a BL Lac candidate based on the correlation study between the ROSAT all-Sky Survey and the Hamburg Quasar Survey (Bade, Fink, & Engels Reference Bade, Fink and Engels1994). Several studies have been carried out to classify this source based on the synchrotron peak frequency ( $\nu_{peak}$ ). The observed $\nu_{peak}$ values are $10^{14.11}$ Hz (Wu, Gu, & Jiang Reference Wu, Gu and Jiang2009), $10^{14.41\pm 0.13}$ Hz (Fan et al. Reference Fan2016), and $10^{14.91}$ Hz (Ackermann et al. Reference Ackermann2015). These studies indicate that this source lies near the boundary between intermediate synchrotron peak blazar (IBL) and high synchrotron peak blazar (HBL) in the classification scheme. VERITAS (Acharyya et al. Reference Acharyya2023) has detected this source in VHE band during MJD 57158-57160 and the average integral flux of $\left (4.51 \pm 0.44\right)\times 10^{-11} \ \mathrm{cm^{-2}} \ \mathrm{s^{-1}}$ above 0.12 TeV was reported. This source is one of the IBL objects detected so far in the VHE band (Benbow et al. Reference Benbow2017; Paiano et al. Reference Paiano, Landoni, Falomo, Treves, Scarpa and Righi2017). Kharb, Gabuzda, & Shastri (Reference Kharb, Gabuzda and Shastri2008) resolved a parsec-scale core with radio observations. For the scientific study, we adopted the redshift $(z) = 0.325$ from Acharyya et al. (Reference Acharyya2023) in this work.

Figure 1. The figure presents the $\gamma$ -ray and optical light curves observed between MJD 55900 and 60150. The top panel shows the 10-day binned $\gamma$ -ray flux (blue points), with the Bayesian block (BB) representation overlaid as a black curve for illustrative purposes only. The bottom panel shows the ASAS-SN optical light curve (green) with the corresponding BB representation (black curve). The grey horizontal lines in both panels indicate the mean $\gamma$ -ray flux and optical magnitude, respectively.

The paper is structured as follows: Section 2 covers multi-wavelength observations and reduction techniques. In Section 3, we explore quasi-periodicity analysis using different methodologies, including the Lomb-Scargle Periodogram (LSP), Weighted Wavelet Z-transform (WWZ), and REDFIT. Section 4 focuses on Gaussian process modelling with a damped random walk model. In Section 5, we assess the significance of quasi-periodic signals (QPOs) using two independent approaches, including Monte Carlo simulations and red noise modelling of light curves. Section 8 presents the findings of our QPO study, while Section 9 presents an interpretation of the observed QPO across multiple bands and wraps up with a conclusion.

Figure 2. The $\gamma$ -ray light curve is analysed using the Lomb-Scargle Periodogram (LSP) and Weighted Wavelet Z-transform (WWZ) methods. The top panel shows the local significance of the detected peak at $\sim$ 0.00134 day $^{-1}$ in $\gamma$ -ray LSP is exceeding 99.73 $\%$ . The bottom panels display the WWZ map (left) and average wavelet power (right). The observed local significance of the detected peak at $\sim$ 0.00132 day $^{-1}$ in avg. wavelet has a significance level of $99.73\%$ .

2. Multi-wavelength observations

2.1 Fermi-LAT observation

The Fermi Gamma-ray Space Telescope, launched by NASA on June 11, 2008, onboard two instruments: the Large Area Telescope (LAT) and the Gamma-ray Burst Monitor (GBM). Together, they enable comprehensive gamma-ray observations across a wide energy range, from a few keV to 500 GeV. The Fermi-LAT, a pair-conversion gamma-ray detector, is designed to explore high-energy gamma rays from $\sim$ 20 MeV to 500 GeV. It provides a wide field of view ( $\gt$ 2 sr), covering about 20 $\%$ of the entire sky. Since its launch, Fermi-LAT has conducted all-sky surveys every three hours, providing near-continuous observations of $\gamma$ -ray emissions from astrophysical sources (Atwood et al. Reference Atwood2009).

We collected Fermi-LAT data of blazar ON 246 during the period 2012 January 6 (MJD 55932) to 2023 May 17 (MJD 60081). During the data download procedure, we chose the energy range of 0.1–300 GeV with Pass8 class events (evclass==128, evtype==3) recommended by the Fermi-LAT collaboration from a region of interest (ROI) with a radius of $10^{\circ}$ centered at the source (R.A.=187.559, Dec = 25.302). The analysis of $\gamma$ -rays was performed following the standard procedures for point-source analysis using the Fermi Science Tools package (v11r05p3), provided by the Fermi Science Support Center. To minimise contamination from the Earth’s limb, a zenith angle cut of $\gt 90^\circ$ was applied. The good time interval (GTI) data was extracted using standard filtering with the following criteria: (DATA_QUAL > 0) && (LAT_CONFIG == 1). This ensures that only high-quality observations are included. We used GTLTCUBE and GTEXPOSURE tools to calculate the integrated livetime as a function of sky position and off-axis angle and exposure, respectively. To model the galactic and extragalactic diffuse background emissions, we used models gll_iem_v07.fitsFootnote a and , respectively. Further, we used the make4FGLxml.py script to create the source model XML file, which contains the information about the source location and the best prediction of spectral form. The unbinned likelihood analysis was performed with GTLIKE tool (Cash Reference Cash1979; Mattox et al. Reference Mattox1996) using the XML spectrum file, and the instrumental response function (IRF) ‘P8R3_SOURCE_V3’ was adopted to get the final source spectrum. To find the significance of the source of interest, we used GTTSMAP tool to calculate the test statistics (TS), which is defined as TS = 2 $\Delta$ log(likelihood) = -2log( $\frac{L}{L_0}$ ), where L and $L_0$ are the maximum likelihood of the model with and without a point source at the target location, respectively. The significance of finding the source at the specified position is assessed by the TS value with TS $\sim \sigma^2$ (Mattox et al. Reference Mattox1996).

We adopted a criterion with TS $\ge$ 9 for data points in the light curve, and a 10-day binned light curve is generated using Fermipy.Footnote b The resulting $\gamma$ -ray light curve is shown in Figure 1.

2.2 ASAS-SN

All-Sky Automated Survey for Supernovae (ASAS-SN; Shappee et al. Reference Shappee2014; Kochanek et al. Reference Kochanek2017) is a global network of 24 telescopes that has been continuously scanning the extragalactic sky since 2012. ASAS-SN’s limiting magnitude of $V\sim$ 16.5–17.5 and $g\sim$ 17.5–18.5 depending on lunation. ASAS-SN camera’s field of view (FOV) is 4.5 deg $^2$ , and the pixel scale and full-width and half maxima (FWHM) are $8^{''}.0$ and $\sim$ 2 pixels, respectively. For this study, we collected both bands’ observations through the ASAS-SN Sky Patrol (V2.0Footnote c ; Shappee et al. Reference Shappee2014; Hart et al. Reference Hart2023).

3. Periodicity search

We adopted various methodologies in search of a potential periodic signal in the $\gamma$ -ray and optical light curves of blazar ON 246. Figure 1 illustrates the 10-day binned $\gamma$ -ray light curve along with optimal Bayesian Block representation (top panel) and an ASAS-SN optical light curve in the bottom panel.

The utilised methodologies include the Lomb-Scargle periodogram (LSP), Weighted Wavelet Z-Transform (WWZ), and a first-order autoregressive model $\left(AR(1)\right)$ in this QPO investigation. A detailed description and the observed findings from all the utilised methods are given in the following Section 3.1, 3.2.

3.1 Lomb-Scargle Periodogram

The Lomb-Scargle periodogram (LSP) (Lomb Reference Lomb1976; Scargle Reference Scargle1979) is one of the most widely used approaches in the literature to identify any potential periodic signal in a time series. In which a sinusoidal function fits the time series using the least square method. This approach is capable of handling the non-uniform sampling in the time series data efficiently by reducing the impact of noise and gaps and providing a precise measurement of the identified periodicity. In this study, we used the GLSP package to compute the Lomb-Scargle (LS) power. The expression of LS power is given as VanderPlas (2018):

(1) \begin{equation}\begin{split}P_{LS}(f) = \frac{1}{2} \bigg[ &\frac{\left(\sum_{i=1}^{N} x_i \cos(2\pi f (t_i - \tau))\right)^2}{\sum_{i=1}^{N} \cos^2(2\pi f (t_i - \tau))} \\[4pt] &+ \frac{\left(\sum_{i=1}^N x_i \sin(2\pi f (t_i - \tau))\right)^2}{\sum_{i=1}^N \sin^2(2\pi f (t_i - \tau))} \bigg]\end{split}\end{equation}

where $\tau$ is specified for each f to ensure time-shift invariance:

(2) \begin{equation} \tau = \frac{1}{4 \pi f} \tan^{-1} \left( \frac{\sum_{i=1}^N \sin\left( 4 \pi f t_i \right)}{\sum_{i=1}^N \cos\left( 4 \pi f t_i \right)} \right)\end{equation}

where we selected the minimum frequency $\left( f_{\min} \right)$ and maximum frequency $\left( f_{\max} \right)$ in temporal frequency range as 1/T and 1/2 $\Delta T$ , respectively, and here T and $\Delta T$ represent the total observation time frame and the time difference between two consecutive points in the light curve, respectively.

The LSP analysis reveals prominent peaks at frequencies of $\sim$ 0.00134 day−1 ( $746\pm68$ days) in the $\gamma$ -ray LSP (see Figure 2) and $\sim$ 0.00132 day−1 ( $757\pm106$ days) in the optical LSP (see Figure 3). Both peaks have a local significance level exceeding 99.73 $\%$ . The uncertainty on the observed period is estimated by fitting a Gaussian function to the dominant LSP peak, and the obtained half-width and half maxima (HWHM) value is used as an uncertainty on period. The distribution of LS power as a function of frequency is given in Figure 2.

Figure 3. The detected QPO signals in the optical emissions from ON 246. The top panel shows the LSP with a dominant peak at $\sim$ 0.00132 day $^{-1}$ has a local significance level exceeding 99.73 $\%$ . The bottom panels display the wavelet map (bottom left panel) and avg. wavelet power at frequency of $\sim0.00131$ day $^{-1}$ with a significance level greater than 99.73 $\%$ .

3.2 Weighted Wavelet Z-transform

In contrast to the LSP approach, the Weighted Wavelet Z-transform (WWZ) (Foster Reference Foster1996) emerges as a powerful, robust, and widely used method in astronomical studies to identify any potential periodic pattern in irregularly sampled light curves. The WWZ method incorporates wavelet analysis, enhancing the LSP’s capabilities by providing better localisation of periodic signals in both temporal and spectral space. In studying the evolution of a periodic signal over time, this approach emerges as a powerful tool, enabling us to identify and characterise the nature of a periodic signal.

In this study, we adopted the abbreviated Morlet kernel that has the following functional form (Grossmann & Morlet Reference Grossmann and Morlet1984):

(3) \begin{equation} f[\omega (t - \tau)] = \exp[i \omega (t - \tau) - c \omega^2 (t - \tau)^2]\end{equation}

and the corresponding WWZ map is given by,

(4) \begin{equation} W[\omega, \tau \;:\; x(t)] = \omega^{1/2} \int x(t)f^* [\omega(t - \tau)] dt\end{equation}

where $f^*$ is the complex conjugate of the wavelet kernel f, $\omega$ is the frequency, and $\tau$ is the time-shift. This kernel acts as a windowed DFT, where the size of the window is determined by both the parameters $\omega$ and a constant c. The resulting WWZ map offers a notable advantage; it not only identifies dominant periodicities but also provides insights into their duration over time.

Figure 4. Analysis of the light curves, left panel represent the REDFIT curve of $\gamma$ -ray emissions and right panel exhibit the REDFIT curve of optical emissions, using the AR(1) process with the REDFIT software. The red noise-corrected power spectrum (black) is presented alongside theoretical (blue) and average AR(1) (cyan) spectra. The significance levels of 99 $\%$ , 95 $\%$ , and 90 $\%$ are indicated in red, green, and brown, respectively.

In this study, we used publicly available python codeFootnote d (Aydin Reference Aydin2017) to generate the WWZ map. The observed power concentration is located around 0.00132 day $^{-1}$ ( $757\pm80$ days) in the WWZ map utilising $\gamma$ -ray emissions (see Figure 2) and at $\sim$ 0.00131 day $^{-1}$ ( $763\pm102$ days) in optical WWZ map (see Figure 3). In both cases, the observed local significance surpasses 99.73 $\%$ . The uncertainty on the period was estimated using the method as described in Section 3.1.

3.3 REDFIT

The light curves of AGNs are typically unevenly sampled, of finite duration, and predominantly influenced by red noise, which arises from stochastic processes occurring in the accretion disk or jet. Red noise spectra are characteristic of autoregressive processes, where current activity is related to past behaviour. The emissions from AGNs are effectively modeled using a first-order autoregressive (AR1) process. To model such behaviour, the software programme REDFIT, developed by Schulz & Mudelsee (Reference Schulz and Mudelsee2002), is specifically designed to analyse the stochastic nature of AGNs dominated by red noise. This software fits the light curve AR(1) process, where the current emission ( $r_t$ ) depends linearly on the previous emission ( $r_{t - 1}$ ) and a random error term ( $\epsilon_t$ ). The AR(1) process is defined as:

(5) \begin{equation} r(t_i)=A_i r(t_{i-1}) + \epsilon(t_i)\end{equation}

where $r(t_i)$ is the flux value at time $t_i$ and $A_i = \exp\left( \left[ \frac{t_{i-1} - t_i}{\tau}\right] \right) \in [0,1]$ , A is the average autocorrelation coefficient computed from mean of the sampling intervals, $\tau$ is the time-scale of autoregressive process, and $\epsilon$ is a Gaussian-distributed random variable with zero mean and variance of unit. The power spectrum corresponding to the AR(1) process is given by

(6) \begin{equation} G_{rr}(f_i) = G_0 \frac{1 - A^2}{1 - 2 A \cos\left( \frac{\pi f_i}{f_{Nyq}} \right) + A^2}\end{equation}

where $G_0$ is the average spectral amplitude, $f_i$ are the frequencies, and $f_{Nyq}$ is representing the Nyquist frequency.

In our study, we used the publicly available REDFITFootnote e code to analysis the light curve. In this method, the Nyquist frequency is defined as $f_{Nyq} = H_{fac}/ (2 \Delta t)$ , where the factor $H_{fac}$ is introduced to prevent the noisy high-frequency end of the spectrum from influencing the fit, as described by equation (7). The REDFIT analysis detected prominent peaks at frequencies of $\sim$ 0.00128 day−1 ( $781\pm160$ days) in the $\gamma$ -ray light curve (see the left panel of Figure 4) and $\sim$ 0.00132 day−1 ( $757\pm160$ days) in the optical light curve (see the right panel of Figure 4). The uncertainties in the periods were estimated using the methodology described in Section 3.1.

Figure 5. The figure displays the posterior probability distributions of the DRW model parameters, obtained from the $\gamma$ -ray light curve (left panel) and the ASAS-SN light curve (right panel).

4. Gaussian process modelling

The observed variability in AGN is inherently stochastic. The AGN light curves can be well described by the stochastic processes, also known as Continuous Time Autoregressive Moving Average [CARMA(p, q)] processes (Kelly et al. Reference Kelly, Becker, Sobolewska, Siemiginowska and Uttley2014), defined as the solutions to the stochastic differential equation:

(7) \begin{equation}\begin{split}\frac{d^p y(t)}{dt^p} + \alpha_{p-1}\frac{d^{p-1}y(t)}{dt^{p-1}}+\ldots+\alpha_0 y(t) =\\\beta_q \frac{d^q \epsilon(t)}{dt^q}+\beta_{q-1}\frac{d^{q-1}\epsilon(t)}{dt^{q-1}}+\ldots+\beta_0 \epsilon(t),\end{split}\end{equation}

where y(t) represents a time series, $\epsilon (t)$ is a continuous time white noise process, and $\alpha_*$ and $\beta_*$ are the coefficients of autoregressive (AR) and moving average (MA) models, respectively. Here, p and q are the order parameters of AR and MA models, respectively.

The simplest model is a continuous autoregressive [CAR(1)] model, also known as the Ornstein-Uhlenbeck process. It is a popular red noise model (Kelly, Bechtold, & Siemiginowska Reference Kelly, Bechtold and Siemiginowska2009; Kozłowski et al. Reference Kozłowski2009; MacLeod et al. Reference MacLeod2012; Ruan et al. Reference Ruan2012; Zu et al. Reference Zu, Kochanek and Udalski2013; Moreno et al. Reference Moreno, Vogeley, Richards and Yu2019; Burke et al. Reference Burke2021; Zhang, Yan, & Zhang Reference Zhang, Yan and Zhang2022, Reference Zhang, Yan and Zhang2023; Sharma et al. Reference Sharma, Kamaram, Prince, Khatoon and Bose2024; Zhang, Yang, & Dai Reference Zhang, Yang and Dai2024; Sharma, Prince, & Bose Reference Sharma, Prince and Bose2024), usually referred to as the damped random walk (DRW) model, described by the following differential equation:

(8) \begin{equation} \left[ \frac{d}{dt} + \frac{1}{\tau_{DRW}} \right] y(t) = \sigma_{DRW} \epsilon(t)\end{equation}

where $\tau_{DRW}$ and $\sigma_{DRW}$ are the characteristic damping time-scale and amplitude of the DRW process, respectively. The mathematical form of the covariance function of the DRW model is defined as

(9) \begin{equation} k(t_{nm}) = a \cdot \exp(\!-\!t_{nm} \, c),\end{equation}

where $t_{nm} = | t_n -t_m|$ denotes the time lag between measurements m and n, with $a = 2 \sigma_{DRW}^2$ and $c = \frac{1}{\tau_{DRW}}$ . The power spectral density (PSD) of the DRW model is defined as:

(10) \begin{equation} S(\omega) = \sqrt{\frac{2}{\pi}} \frac{a}{c} \frac{1}{1 + (\frac{\omega}{c})^2}\end{equation}

The DRW PSD has a form of Broken Power Law (BPL), where the broken frequency $f_b$ corresponds to the characteristic damping timescale $\tau_{DRW} = \frac{1}{2\pi f_b}$ .

In the best-fit parameters estimation of the DRW model for both light curves, we employed the Markov chain Monte Carlo (MCMC) algorithm provided by the emceeFootnote f package (Foreman-Mackey et al. Reference Foreman-Mackey, Hogg, Lang and Goodman2013). For the modelling, we employed the EzTaoFootnote g package, which is built on top of celerite.Footnote h In this study, we generated the distributions of the posterior parameters by running 10 000 steps as burn-in and 20 000 as burn-out, as illustrated in Figure 5. The resulting best-fitted $\gamma$ -ray and optical light curves are presented in Figures 6 and 7. The autocorrelation functions of the standardised residuals and their squares indicate that the DRW model effectively captures the characteristic variability in both light curves. The corresponding DRW power spectral densities are displayed in Figure 8, where the shaded regions represent areas of unreliability due to the finite duration and sampling cadence of the light curves.

Figure 6. The celerite modelling with DRW model was performed using the 10-day binned $\gamma$ -ray light curve of blazar ON 246 over 4 000 days from the time stamp MJD 55932. In this figure, the top panel depicts the $\gamma$ -ray flux points with their uncertainties, along with the best-fitting profile of the DRW model in blue, including the 1 $\sigma$ confidence interval. The bottom panels represent the autocorrelation functions (ACFs) of the standardised residuals (bottom left) and the squared of standardised residuals (bottom right), respectively, along with 95 $\%$ confidence intervals of the white noise.

Figure 7. This figure demonstrates the modelling of the ASAS-SN light curve of ON 246 with the DRW model. The top panel shows the best-fitting profile of the DRW modelling along with 1 $\sigma$ confidence interval. The bottom panels represent the autocorrelation functions as described in Figure 6.

Figure 8. This figure presents the DRW PSDs obtained from $\gamma$ -ray and ASAS-SN observations, along with their 1 $\sigma$ confidence intervals. The two shaded regions highlight biased regions due to observational limitations, such as finite length and the mean cadence of the light curve. The regions with orange star symbols represent the invalid areas in $\gamma$ -ray PSD, while the blue hatch line regions in PSD indicate limitations imposed by the ASAS-SN light curve’s duration and mean cadence.

5. Probe the significance

AGN emissions exhibit stochastic variability and are well characterised by red noise. The combination of red noise, characteristic nature, and uneven sampling in the light curve can lead to spurious peaks in the periodogram. Therefore, it is crucial to evaluate the significance of any periodic signals detected in the light curves. In the estimation of significance, we employed two different approaches.

Figure 9. The Lomb-Scargle periodograms of the $\gamma$ -ray (left panel) and ASAS-SN (right panel) observations of the blazar ON 246 are shown. The figure presents the LSPs of the original light curves (black) and spectral windows (pink). The significance levels (blue) of the dominant observed peaks in both periodograms are estimated using DRW synthetic light curves. The detected peaks at 0.00105 day $^{-1}$ ( $\sim$ 950 days) and 0.00105 day $^{-1}$ ( $\sim$ 950 days) in $\gamma$ -ray and ASAS-SN observations exceed 3 $\sigma$ and 4 $\sigma$ significance levels, respectively. The shaded regions in both figures represent the invalid areas estimated using the mean cadence and baseline of the light curves. Logarithmic versions of the periodograms are provided in the sub-figures.

The periodogram is usually represented as a power spectral density (PSD) of a form $P(\nu) \sim A \nu^{-\beta}$ , where the temporal frequency is represented by $\nu$ and $\beta \gt 0$ represents the spectral slope. In the first approach, to calculate the statistical significance of the periodic feature, we employed the approach developed by Emmanoulopoulos, McHardy, & Papadakis (Reference Emmanoulopoulos, McHardy and Papadakis2013). The methodology involved modelling the observed PSD using a power-law model. For that, we employed DELightcurveSimulation Footnote i code, which involved randomising the amplitude and phase of the Fourier components, each mimicking the characteristics of the original data, including the best-fitting power-law model slope and similar properties in terms of flux distribution. We performed a Monte Carlo simulation, generating $5\times10^4$ synthetic light curves for each case. The simulations were based on the best-fitting power-law model slopes ( $\beta$ = 0.48 $\pm$ 0.16 for $\gamma$ -rays light curve and $\beta$ = 1 $\pm$ 0.16 for ASAS-SN light curve) utilising their flux distribution properties, respectively. Each simulated light curve for both cases mimics the underlying properties of the original light curves. After that, we generated the Lomb-Scargle periodogram of each simulated light curve as we did for both original light curves. To estimate the significance level of the dominant peak in original LSPs, we calculated the 84th, 97.5th, 99.85th, and 99.995th percentiles of the 50 000 simulated LSP for each frequency value, which correspond to the 1 $\sigma$ , 2 $\sigma$ , 3 $\sigma$ , and 4 $\sigma$ significance levels. In this first approach, the significance level of the dominant peak in $\gamma$ -ray LSP surpasses the 3 $\sigma$ and just touches the 3 $\sigma$ significance level in ASAS-SN LSP, as shown in Figures 2 and 3.

In addition to modelling the PSDs with a simple power-law, we also applied a bending power-law model, as detailed in Appendix A. The prominent peak observed in the $\gamma$ -ray LSP at $\sim 0.00134 \ \mathrm{day^{-1}}$ exhibits a significance of approximately $3\sigma$ , while, the dominant peak at $\sim 0.00132 \ \mathrm{day^{-1}}$ in the optical LSP surpasses a significance threshold of $3.48\sigma$ .

In the second approach, considering that AGN variability is stochastic and well characterised by a red noise process, we employed the Damped Random Walk (DRW) model to describe both light curves and determine the optimal model parameters, including the amplitude and damping timescale. Using the EzTao package, we simulated 10 000 synthetic light curves with a sampling rate consistent with the real observations. After generating these synthetic light curves, we computed Lomb-Scargle periodograms (LSPs) for each, following the same procedure as for the original $\gamma$ -ray and ASAS-SN light curves. The significance levels were then estimated using the same methodology as described earlier. From this analysis, the peaks at 0.00134 day $^{-1}$ in $\gamma$ -ray LSP, while the peak at 0.00132 day $^{-1}$ in the optical LSP surpasses the 4 $\sigma$ threshold. Additionally, we calculated the spectral window periodogram by constructing a light curve with a total number of time stamps ten times larger than the original within the observed temporal frame. In this synthetic light curve, the time stamps matching the original observations were assigned a value of one, while all others were set to zero. Further, we applied the LSP method to generate the periodogram, as shown in pink in Figure 9. The QPO findings are summarised in Table 1.

Table 1. Summary of the quasi-periodic signal detection using three different methodologies. Column (1) lists the 4FGL name of the blazar ON 246, while Column (2) specifies the observational waveband. Columns (3) and (4) present the QPO frequencies obtained from the Lomb-Scargle periodogram and Weighted Wavelet Z-transform methods, along with their uncertainties. The local significance of each detected QPO signal is provided in parentheses next to the corresponding frequency value. Column (5) presents the estimated local significance level of the QPO in LSP based on DRW-modeled mock light curves, and Column (6) provides the QPO frequency and significance level derived from the REDFIT analysis.

Note: The QPO peaks were fitted using a Gaussian function, and the uncertainties correspond to the half-width at half-maximum (HWHM).

6. Gamma-ray/optical cross correlations

We investigate the possible time lags between the $\gamma$ -ray and optical light curves of ON 246 utilising the interpolated cross-correlation function (ICCF: Peterson et al. Reference Peterson, Wanders, Horne, Collier, Alexander, Kaspi and Maoz1998, Reference Peterson2004), which is one of the commonly used methods in the time-series analysis of AGNs. As we know, AGN light curves generally are not regularly sampled in time but are discretely sampled N times at time stamps $t_i$ , where $t_{i+1} - t_i = \Delta t$ for all values $1 \le i \le N-1$ . ICCF method emerges as a powerful technique to estimate the time-leg between two-time series. The method uses the linear interpolation method to deal with unevenly sampled light curves and calculate the cross-correlation coefficient as a function of the time lag for two light curves:

(11) \begin{equation} F_{CCF}(\tau)=\frac{1}{N} \sum_{i=1}^N \frac{\left[ L(t_i) - \bar{L} \right] \left[ C(t_i - \tau) - \bar{C} \right]}{\sigma_L \sigma_C}\end{equation}

where N is the number of data points in the light curves L and C. Each light curve has a corresponding mean value ( $\bar{L}$ and $\bar{C}$ ) and uncertainty ( $\sigma_L$ and $\sigma_C$ ).

The ICCF is evaluated for a time lag ( $\tau$ ) in a range [-1 000, 1 000] days with searching step $\Delta \tau$ , which should be smaller than the median sampling time of the light curves. We adopted $\Delta \tau$ = 7 days and used the public PYTHON version of the ICCF, PYCCF (Sun, Grier, & Peterson Reference Sun, Grier and Peterson2018) in this study. As applying the ICCF to the light curves, a strong peak is obtained, and its corresponding centroid is calculated using the ICCF for the time lags around the peak. We adopted the centroid of the CCF ( $\tau_{cent}$ ) using only time lags with $r\gt 0.8 r_{\max}$ , where $r_{\max}$ is the peak value of the CCF. The 1 $\sigma$ confidence on the time lag is estimated using a model-independent Monte Carlo method. We found a lag of $0.736_{-2.73}^{+3.13}$ day with the $\gamma$ -ray leading the optical emissions. In addition to cross-correlation centroid distribution (CCCD), we also estimated the cross-correlation peak distribution (CCPD; see the right panel of Figure 10). To access the significance, we also calculated the ICCF between the $\gamma$ -ray and DRW mock optical light curves (see Section 5); the observed findings are shown in the left panel of Figure 10. After simulating the 10 000 mock ICCFs between $\gamma$ -ray and optical mock light curves, we calculated the significance level of the ICCF peak observed in correlation analysis between the observed $\gamma$ -ray and ASAS-SN light curves. In the left panel of Figure 10, a red dashed curve represents the 99.9999 $\%$ , corresponding to 4 $\sigma$ , significance level. The observed finding indicates a significant correlation with almost zero-day lag between $\gamma$ -ray and optical emissions, suggesting a common origin of them (Cohen et al. Reference Cohen, Romani, Filippenko, Bradley Cenko, Lott, Zheng and Li2014).

Figure 10. Cross-correlation analysis between $\gamma$ -ray and optical flux variations. The left panel shows ICCF (solid black curve) with a significance level (dash red curve) of 4 $\sigma$ . The cross-correlation centroid distribution (CCCD) in orange and cross-correlation peak distribution (CCPD) in blue are given in the right panel, where vertical dashes in orange and blue represent the median of CCCD and CCPD, respectively.

7. Black hole mass estimation

The mass of the central black hole in an AGN is one of the most crucial parameters for understanding its central engine. Several methods have been developed to estimate black hole mass, including stellar and gas kinematics and reverberation mapping (Gupta, Srivastava, & Wiita Reference Gupta, Srivastava and Wiita2008). The stellar and gas kinematics approach requires high spatial resolution spectroscopic observations to resolve the gravitational influence of the black hole. Reverberation mapping relies on detecting high-ionisation emission lines from regions close to the black hole. In addition to the black hole mass estimation methods mentioned above, the variability timescales can also serve as a useful tool for determining the mass of an AGN’s black hole. The timescales of high-amplitude variations may be linked to the black hole mass, offering valuable insights into the central engine of AGNs. The shortest variability timescale provides a constraint on the size of the emitting region in blazars. In relativistic jets, the emitting region is often simplified as a ‘blob’ with a characteristic size D. This size can be constrained by the relation:

(12) \begin{equation} D \lesssim \frac{\delta \Delta t_{{obs}} c}{1+z}\end{equation}

where $\delta$ is the Doppler factor of the relativistic jet, $\Delta t_{obs}$ is the observed variability timescale, z is the redshift of the blazar, and c represents the speed of light.

Several studies (Hartman Reference Hartman, Richard Miller, Webb and Noble1996; Ghisellini & Madau Reference Ghisellini and Madau1996; Xie et al. Reference Xie, Bai, Zhang and Fan1998; Yang & Fan Reference Yang and Fan2010) have suggested that $\gamma$ -ray emissions originate from a region located a few hundred Schwarzschild radii ( $r_g\equiv GM_* /c^2$ ), where G and $M_*$ are the gravitational constant and black hole mass, respectively. The emission region is typically constrained to a size of $r \lt 200 r_g $ . Yang & Fan (Reference Yang and Fan2010) provided a relation to estimate the lower limit of the black hole mass in terms of $\gamma$ -ray luminosity, redshift, and variability timescale, which is defined as

(13) \begin{equation} \frac{\mathrm{M}}{ \ \ \mathrm{M_{\odot}}} \ge 1\times10^3 \ \frac{\Delta \mathrm{ t_{obs}}}{\mathrm{1+z}} \ \left[\frac{L_{\gamma}\mathrm{(1+z)}}{6.3\times 10^{40} \Delta \mathrm{ t_{obs}}}\right]^{\frac{1}{4+\alpha_{\gamma}}}\end{equation}

The $\gamma$ -ray luminosity is defined as

(14) \begin{equation} L_{\gamma} = 4 \pi d_L^2 (1+z)^{\alpha_{\gamma}-2} \times f\end{equation}

where f is the total $\gamma$ -ray flux calculated using equations (19) or (20), $(1+z)^{\alpha_{\gamma} - 2}$ represents a K-correction factor, and $d_L$ is the luminosity distance value estimated following the expression given as

(15) \begin{equation} d_L = \frac{c}{H_0} \int_1^{1+z} \ \frac{1}{\sqrt{\Omega_M x^3 \ + \ 1 \ - \ \Omega_M }} \ dx\end{equation}

Using the $\Lambda-\mathrm{CDM}$ model with $\Omega_M$ , we calculated $d_L$ , where $H_0$ is a hubble constant ( $H_0=73\,{\rm km}\,{\rm s}^{-1}\,{\rm Mpc}^{-1}$ ).

The $\gamma$ -ray photon number per unit energy can be defined as a power-law distribution

(16) \begin{equation} \frac{dN}{dE} = N_0 E^{-\alpha_{\gamma}}\end{equation}

where $\alpha_{\gamma}$ represents the photon spectral index and $N_0$ is the normalisation constant. $N_0$ can be estimated by integrating equation (16)

(17) \begin{equation} N_0 = N_{(E_L - E_U)} \left( \frac{1}{E_L} - \frac{1}{E_U}\right)\!, \quad \mathrm{if} \ \ \alpha_{\gamma} = 2;\end{equation}
(18) \begin{equation} N_0 = N_{(E_L - E_U)} . \frac{1 - \alpha_{\gamma}}{E_U^{1 - \alpha_{\gamma}} - E_L^{1 - \alpha_{\gamma}}}, \quad \mathrm{if} \ \ \alpha_{\gamma} \neq 2\end{equation}

where $N_{(E_L - E_U)}$ is the integral photons in units of photon cm−2 s−1 in the energy range $E_L - E_U$ , $E_L$ and $E_U$ represent the lower and upper energy limits respectively. In this study, we set $E_L$ = 100 MeV and $E_U$ = 300 GeV. Thus, the total $\gamma$ -ray flux can be calculated by $f = \int_{E_L}^{E_U} \ EdN$ , which elaborated as

(19) \begin{equation} f = N_{(E_L - E_U)} \left( \frac{1}{E_L} - \frac{1}{E_U}\right)ln \frac{E_U}{E_L}, \quad \mathrm{if} \ \ \alpha_{\gamma} = 2; \end{equation}
(20) \begin{equation} f = N_{(E_L - E_U)} \frac{1 - \alpha_{\gamma}}{2 - \alpha_{\gamma}} \ \frac{\left( E_U^{2-\alpha_{\gamma}} \ - E_L^{2-\alpha_{\gamma}} \right)}{E_U^{1-\alpha_{\gamma}} \ - E_L^{1-\alpha_{\gamma}}}, \quad \mathrm{otherwise} \end{equation}

in units of GeV cm−2 s $^{-1}$ .

Acharyya et al. (Reference Acharyya2023) reported the spectral parameters of power-law distribution equation (16), including $N_0 = (6.52 \pm 1.12 )\times 10^{-11} \ \mathrm{cm^{-2}} \ \mathrm{s^{-1}} \ \mathrm{MeV^{-1}}$ and $\alpha_{\gamma} = 1.79 \pm 0.14$ . The estimated $\gamma$ -ray luminosity is $\sim 1.18\times 10^{47}\,{\rm erg}\,{\rm s}^{-1}$ and the mass of the black hole to be $M_* \gt 1.42 \times 10^8\,{\rm M}_{\odot}$ using the equations (14) and (13), respectively.

Additionally, Liu & Bai (Reference Liu and Bai2015) proposed a model to determine the upper limit of the black hole mass based on variability studies, which is briefly discussed here. In this model, a blob in the jet-production region initially has a size of $D_0$ and expands to $D_R$ at a distance $R_{jet}$ from the central black hole. As the blob propagates outward along the jet, it expands with an average velocity $\left( \bar{v}_{exp} \right)$ . Since the expansion velocity is non-relativistic ( $\bar{v}_{exp} \lt\lt \bar{v}_{jet}$ , where $\bar{v}_{jet}$ is the relativistic jet velocity), we have the condition $D_0 \lesssim D_R$ . Following equation (12), we can express this relationship as:

(21) \begin{equation} D_0 \lesssim D_R \leqslant \frac{\delta \Delta t_{\mathrm{min}}^{obs}}{1+z} c\end{equation}

where $\Delta t_{\min}^{obs}$ is the observed minimum timescale of variability. Relativistic jets are believed to originate from the inner regions of the accretion disk, close to the central black hole (Blandford & Znajek Reference Blandford and Znajek1977; Blandford & Payne Reference Blandford and Payne1982; Meier, Koide, & Uchida Reference Meier, Koide and Uchida2001). The inner radius of the disk is typically assumed to be near the marginally stable orbit, also known as the innermost stable circular orbit (ISCO). The ISCO radius ( $r_{ms}$ ) is expressed in terms of the gravitational radius ( $r_g$ ) and the dimensionless spin parameter $j = J/J_{\max}$ , where the maximum angular momentum is given by $J_{\max}=GM_*^2 / c$ , as defined in Equation (5) of Liu & Bai (Reference Liu and Bai2015). For a Schwarzschild black hole ( $j=0$ , non-rotating), the ISCO is located at $r_{ms}=6r_g$ , where $r_g=GM_* / c^2$ . The corresponding size of the emitting region is $D_{ms}=12r_g$ .

In the case of Kerr black hole (j = 1, maximally rotating), for prograde orbits, the emitting region is $D_e = 2r_e = 4r_g$ , where $r_e$ represents the equatorial boundary of the ergosphere. These estimated sizes are consistent with findings from general relativistic magnetohydrodynamic (GRMHD) simulations (Meier, Koide, & Uchida Reference Meier, Koide and Uchida2001). Thus, the size of the blob spans from $D_0=4-12 r_g$ for $0\leqslant j \lesssim 1$ . Following equation (13), we have (as given in equations 6a and 6b in Liu & Bai Reference Liu and Bai2015)

(22) \begin{align} M_{{var}}^{{Sch}} &\lesssim 1.695 \times 10^4 \frac{\delta \Delta t_{\mathrm{min}}^{obs}}{1+z} {\rm M}_{\odot}, \quad (j \sim 0) \end{align}
(23) \begin{align} M_{{var}}^{{Kerr}} &\lesssim 5.086 \times 10^4 \frac{\delta \Delta t_{\mathrm{min}}^{obs}}{1+z} {\rm M}_{\odot}, \quad (j \sim 1) \end{align}

where $\Delta t_\mathrm{min}^{obs}$ is in units of seconds.

Acharyya et al. (Reference Acharyya2023) reported the fasted observed variability timescale in $\gamma$ -rays of 6.2 $\pm$ 0.9 hr around MJD 57178. We utilised this timescale to constrain the black hole mass. The calculated black hole masses are $M_{var}^{Sch} \lesssim2.74 \times 10^9\,{\rm M}_{\odot}$ and $M_{var}^{Kerr} \lesssim8.22\times 10^9\,{\rm M}_{\odot}$ . In the calculation, we adopted the Doppler factor value, $\delta=9.6$ (Zhou et al. Reference Zhou, Zheng, Zhu and Kang2021; Acharyya et al. Reference Acharyya2023). The uncertainties in the estimated black hole masses were determined based on the errors in the variability timescale. However, the lack of precise information regarding the exact location of the emission region relative to the central black hole could introduce additional uncertainties, potentially increasing the error bars on the mass estimates.

By combining the estimates of the lower and upper limits of the black hole mass, we obtain a mass range of $(0.142 \lt M_* \lt 8.22)\times 10^9\,{\rm M}_{\odot} $ . Notably, the lower limit derived from the variability study is higher than the estimate reported by Pei et al. (Reference Pei, Fan, Yang, Huang and Li2022).

In recent years, several studies have demonstrated a correlation between variability timescales and black hole mass. Burke et al. (Reference Burke2021) identified a relationship between optical timescales in accretion disks and black hole masses, spanning the entire mass range of SMBH and even extending to stellar-mass BH systems. This correlation has been further explored across different EM bands, including $\gamma$ -rays (Ryan et al. Reference Ryan, Siemiginowska, Sobolewska and Grindlay2019; Zhang, Yan, & Zhang Reference Zhang, Yan and Zhang2022; Zhang, Yan, & Zhang Reference Zhang, Yan and Zhang2023; Sharma et al. Reference Sharma, Kamaram, Prince, Khatoon and Bose2024; Zhang, Yang, & Dai Reference Zhang, Yang and Dai2024; Sharma, Prince, & Bose Reference Sharma, Prince and Bose2024), X-rays (Zhang, Yang, & Dai Reference Zhang, Yang and Dai2024), and sub-millimeter wavelengths (Chen et al. Reference Chen, Bower, Dexter, Markoff, Ridenour, Gurwell, Rao and Wallström2023). Notably, the observed $\gamma$ -ray variability timescales in AGNs have been found to overlap with those in the optical band (Burke et al. Reference Burke2021), suggesting a link between non-thermal jet emissions and thermal emissions from accretion disks. Thus, by utilising the relationship between the rest-frame variability timescale $\left(\tau_{rest}^{Damping}\right)$ and BH mass $\left(M_{BH}\right)$ as established by Burke et al. (Reference Burke2021), $\tau_{rest}^{damping}=107_{-12}^{+12} \left(\frac{M_{BH}}{10^8\,{\rm M}_{\odot}}\right)^{0.38_{-0.04}^{+0.05}} \ \mathrm{days}$ , the mass of the SMBH can be estimated.

In our study, we estimated the mass of the SMBH to be approximately $(7.3 \pm 6) \times 10^9\,{\rm M}_{\odot}$ based on a rest-frame damping timescale of $\sim$ 547 days observed in $\gamma$ -ray emission of the blazar ON 246. While the uncertainty in the SMBH mass estimation is considerably high, the derived mass value is less and close to the upper limit of $M_{var}^{Kerr}$ and higher than the $M_{var}^{Sch}$ , as calculated using the model proposed by Liu & Bai (Reference Liu and Bai2015). Further, the SMBH mass of ON 246, estimated using the optical timescale derived from DRW modelling, is $(1.56 \pm 0.66)\times 10^8\,{\rm M}_{\odot}$ .

The estimated black hole mass of ON 246, derived from variability characteristics in both energy bands, falls within the range of $(0.142 - 8.22) \times 10^9\,{\rm M}_{\odot}$ . The estimated mass also lies in the range (10 $^8$ –10 $^9$ M $_{\odot}$ ) of mass of SMBH for FSRQ as derived in various studies (Ghisellini & Tavecchio Reference Ghisellini and Tavecchio2008; Castignani et al. Reference Castignani, Haardt, Lapi, De Zotti, Celotti and Danese2013; Xiong & Zhang Reference Xiong and Zhang2014; Paliya et al. Reference Paliya, Domnguez, Ajello, Olmo-Garca and Hartmann2021; Zhang et al. Reference Zhang, Yang and Dai2024) using various sample of FSRQs. Using the gamma-ray variability time, Pei et al. (Reference Pei, Fan, Yang, Huang and Li2022) have also estimated the mass of the black hole, but the value is close to $\sim$ 8.08 $\times$ 10 $^{7}$ M $_{\odot}$ , which is much smaller than the value estimated above. This discrepancy is because in Pei et al. (Reference Pei, Fan, Yang, Huang and Li2022), authors have used a fixed variability time of 1 day and the Doppler factor ( $\delta$ ) as a very small value of 0.48.

8. Result and discussion

In this study, we investigated quasi-periodic signals in the $\gamma$ -ray and optical emissions of the blazar ON 246 (4FGL J1230.2+2517) over the period of 11.6 yr, from MJD 55932 to 60081, employing three different methodologies as outlined in Section 3. Our analysis reveals a distinct quasi-periodic signal in the $\gamma$ -ray emissions of the blazar ON 246, with periods of approximately 746 days, 757 days, and 781 days, as identified using the Lomb-Scargle Periodogram (LSP), Wavelet Weighted Z-transform (WWZ), and REDFIT analysis, respectively. The significance of the detected periodicity was assessed through two independent approaches: Monte Carlo simulations and stochastic modelling using the damped random walk (DRW) model. The dominant peaks detected in both LSP and WWZ have a local significance level above 3 $\sigma$ , based on Monte Carlo simulations, and exceed 3 $\sigma$ when evaluated using DRW modelling. The uncertainties in the observed periods were estimated as the half-width at half-maximum (HWHM) by fitting the peak profiles with a Gaussian function. The QPO detected in $\gamma$ -ray emissions is further corroborated by an independent analysis of the optical light curve. Using the same methodologies, we searched for oscillatory signals in the optical data and assessed their significance. A similar periodicities of $\sim$ 757, $\sim$ 763, and $\sim$ 757 days were found in LSP, WWZ, and REDFIT analyses, respectively, with a significance level exceeding 3 $\sigma$ , reinforcing the presence of the detected QPO. Additionally, the phase-folded light curves of Fermi-LAT and ASAS-SN were fitted with sinusoidal functions corresponding to frequencies of $\sim 0.00134 \ \mathrm{day^{-1}}$ and $\sim 0.00132 \ \mathrm{day^{-1}}$ , respectively, as illustrated in Figure 11.

Figure 11. The folded Fermi-LAT and ASAS-SN light curves of ON 246 with a period of 746 and 757 days are shown in the top and bottom panels, respectively. The dashed blue line represents the mean value, and the sine functions (red) with frequencies of 0.00134 and 0.00132 day $^{-1}$ were fitted to the folded $\gamma$ -ray and optical light curves, respectively. Two full period cycles are shown for better clarity.

In addition, we also employed the interpolation cross-correlation function (ICCF) to analyse the correlation between $\gamma$ -ray and optical emissions. As display in the Figure 10, it can be seen that the correlation coefficient is maximum at near zero time-lag, with $\tau_{cent}=0.736_{-2.73}^{+3.13}$ d, have a significance level exceeding 4 $\sigma$ , indicating that the variability features between these two bands are similar in nature and is believed to be originated by the lepton single-zone scenario of blazar emission (Cohen et al. Reference Cohen, Romani, Filippenko, Bradley Cenko, Lott, Zheng and Li2014). Studying correlations across multiple wavebands is crucial for gaining deeper insights into the emission mechanisms of these systems. A strong correlation between low-energy and high-energy emissions can be well explained by the leptonic model. In this scenario, low-energy radiation originates from synchrotron emission, while the same seed photons undergo synchrotron self-Compton (SSC) and inverse Compton scattering to produce high-energy radiation, leading to a significant correlation between the two energy bands (Maraschi, Ghisellini, & Celotti Reference Maraschi, Ghisellini and Celotti1992; Giommi et al. Reference Giommi1999; Tagliaferri et al. Reference Tagliaferri, Ravasio, Ghisellini, Tavecchio, Giommi, Massaro, Nesci and Tosti2003; Zheng et al. Reference Zheng, Zhang, Huang and Kang2013; Liao et al. Reference Liao, Bai, Liu, Weng, Chen and Li2014; Li et al. Reference Li, Jiang, Guo, Chen and Yi2016).

Conversely, when high-energy emission is generated through the external Compton process, which involves seed photons originating outside the jet (Malmrose et al. Reference Malmrose, Marscher, Jorstad, Nikutta and Elitzur2011; Liao et al. Reference Liao, Bai, Liu, Weng, Chen and Li2014), the correlation between low-energy and high-energy emissions tends to weaken or become insignificant. Our findings reveal a strong correlation between the optical and $\gamma$ -ray emissions of the blazar ON 246, supporting the leptonic model’s predictions. Acharyya et al. (Reference Acharyya2023) observed a strong correlation between MeV and GeV emissions with a peak at zero time-lag, suggesting the same origin, and also found the positive time-lags with radio and optical emissions.

In this study, we employed three different methods to estimate the black hole (BH) mass of the blazar ON 246 based on its variability characteristics. First, we determined the lower and upper limits of the BH mass using the models proposed by Yang & Fan (Reference Yang and Fan2010) and Liu & Bai (Reference Liu and Bai2015), respectively. The minimum variability timescale observed in the $\gamma$ -ray band yielded a lower mass limit of $M_* \gt 0.142 \times 10^9\,{\rm M}_{\odot}$ . For the upper limit, we obtained $M_* \lt 2.74 \times 10^9\,{\rm M}_{\odot}$ for a Schwarzschild black hole, and $M_* \lt 8.22 \times 10^9\,{\rm M}_{\odot}$ for a maximally rotating Kerr black hole. In a third approach, we utilised the damping timescales of $\gamma$ -ray and optical light curves, applying an empirical correlation between the rest-frame damping timescale ( $\tau_{{rest}}^{\textrm{Damping}}$ ) and the BH mass, as established by Burke et al. (Reference Burke2021). This method yielded BH mass estimates of $(7.3 \pm 6) \times 10^9\,{\rm M}_{\odot}$ from $\gamma$ -ray variability and $(1.56 \pm 0.66) \times 10^8\,{\rm M}_{\odot}$ from optical variability. Combining these results, we constrain the BH mass of ON 246 to lie within the range of approximately $(0.142 - 8.22) \times 10^9\,{\rm M}_{\odot}$ .

8.1 Potential mechanisms for QPO

Several physical models have been proposed in the literature to explain the phenomenon of quasi-periodic oscillations (QPOs) in blazars. These include scenarios involving supermassive binary black hole (SMBBH) systems, precession or helical motion of relativistic jets (geometric effects), and instabilities in the accretion flow within the disk. A more detailed discussion on each of these scenarios is given below.

The supermassive binary black hole (SMBBH) scenario provides a plausible explanation for long-term flux modulations observed in blazars (Sillanpaa et al. Reference Sillanpaa, Haarala, Valtonen, Sundelius and Byrd1988; Xie et al. Reference Xie, Yi, Li, Zhou and Chen2008; Valtonen et al. Reference Valtonen2008; Villforth et al. Reference Villforth2010; Graham et al. Reference Graham2015; Wang, Yin, & Xiang Reference Wang, Yin and Xiang2017; Gong et al. Reference Gong, Zhou, Yuan, Zhang, Yi and Fang2022b, 2024). Several sources have been identified as potential candidates exhibiting long-term periodic flux modulations across multiple wavebands. Notable examples include a $\sim$ 12-yr QPO in OJ 287 (Sillanpaa et al. Reference Sillanpaa, Haarala, Valtonen, Sundelius and Byrd1988; Valtonen et al. Reference Valtonen2008; Villforth et al. Reference Villforth2010; Sandrinelli et al. Reference Sandrinelli, Covino, Dotti and Treves2016), a $2.18 \pm 0.08$ -yr periodicity in PG 1335 + 113 (Ackermann et al. Reference Ackermann2015), a $1.84 \pm 0.1$ -yr period in PKS 1510-089 (Xie et al. Reference Xie, Yi, Li, Zhou and Chen2008), and a 3-yr periodicity in 3C 66A (Otero-Santos et al. Reference Otero-Santos2020). Several other sources are also considered potential candidates exhibiting long-term periodic flux variations consistent with the SMBBH scenario. These periodic modulations in blazars are often attributed to the orbital motion of the secondary black hole within the binary system.

Rieger (Reference Rieger2004) investigated the potential geometrical origins of periodicity in blazar-type sources, proposing that periodic variations in emission can result from three possible mechanisms, including orbital motion in a binary black hole system, jet precession, and intrinsic jet rotation. The precessional-driven ballistic motion is unlikely to produce observable periods shorter than several decades, While the orbital motion in a close SMBBH system produces a period of $P_{obs}\gtrsim$ 10 days and Newtonian-driven precession in a close SMBBH can be a possible mechanisms of periodicity $P_{obs} \gtrsim 1$ yr. Therefore, the observed periodic flux modulations in emission from these sources canbe reasonably explained by orbital-driven (nonballistic) helical motion in a close SMBBH system.

In our study, the observed oscillation period is $P_{obs}$ = 746 days, which is significantly shorter than the actual physical period $P_d$ due to the light-travel time effect (Rieger Reference Rieger2004). The period is related by the equation, $P_d = \frac{P_{obs} \Gamma^2}{1+z}$ , where $\Gamma$ is the bulk Lorentz factor. To estimate $\Gamma$ , we used a relation from (Megan & Padovani Reference Megan and Padovani1995) and Li et al. (Reference Li, Qin, Gong, Liu, Guo, Gao, Yi and Liu2024), $\Gamma \le \frac{1}{2} \left(\delta + \frac{1}{\delta}\right)$ , which is basically gives the lower limit to the $\Gamma$ at a given value of $\delta$ . This condition is valid only for $\delta \gt1$ . Using a Doppler factor of $\delta$ = 9.6 (Zhou et al. Reference Zhou, Zheng, Zhu and Kang2021), we estimate the lower limit of the Lorentz factor as $\Gamma \geq 4.85$ . Consequently, the intrinsic period of the blazar ON 246 is $\sim$ 36.28 yr. Based on this corrected QPO period, the estimated mass of the primary black hole is

(24) \begin{equation} M \simeq P_{\mathrm{corrected, yr}}^{8/5} \ R^{3/5} \ 10^6\,{\rm M}_{\odot}\end{equation}

where $P_{\mathrm{corrected}}$ is the real physical period in units of years, R represents the mass ratio of the primary black hole to the secondary black hole, $R=\frac{M}{m}$ . Previous studies (Roland et al. Reference Roland, Britzen, Caproni, Fromm, Glück and Zensus2013; Yang et al. Reference Yang, Yan, Zhang, Dai and Zhang2021; Gong et al. Reference Gong, Zhou, Yuan, Zhang, Yi and Fang2022a) have suggested that the R lies within the ranges of 4–10.5, 1–100, and 10–100, respectively. In our analysis, we adopt R within the range of 1–100 to calculate the mass of the primary black hole. Based on this, the estimated mass is found to be in the range of $ M \sim 3.12\times 10^8 - 5 \times 10^9\,{\rm M}_{\odot} $ . The estimated BH mass range is comparable with the derived mass range in Section 7, indicating that the observed year-like QPO is likely caused by non-ballistic helical motion driven by the orbital dynamics of a close SMBBH system.

Additionally, instabilities in the accretion flow within the disk may also contribute to the flux modulations in blazars. In this scenario, oscillations in the innermost regions of the accretion disk or Kelvin-Helmholtz instabilities could lead to quasi-periodic plasma injections into the jet, as a result producing quasi-periodic oscillations in the jet emissions (Gupta, Srivastava, & Wiita Reference Gupta, Srivastava and Wiita2008; Wang et al. Reference Wang, An, Baan and Lu2014; Bhatta et al. Reference Bhatta2016; Sandrinelli et al. Reference Sandrinelli, Covino, Dotti and Treves2016; Tavani et al. Reference Tavani, Cavaliere, Munar-Adrover and Argan2018). The mass of the SMBH can be estimated using the following equation

(25) \begin{equation} M=\frac{3.23 \ \times \ 10^4 \ \delta \ P_{obs} }{\left(r^{3/2}+a\right) \left(1+z\right)}\,{\rm M}_{\odot}\end{equation}

where $P_{obs}$ is in units of seconds, $\delta$ is the Doppler factor, parameter r is the radius of this source zone in untis of $GM/c^2$ , parameter a is the spin parameter of SMBH, and z is redshift. In our study, using equation (17), the estimated mass of the SMBH is $1.02 \times 10^{12}\,{\rm M}_{\odot}$ for a Schwarzschild black hole with r = 6 and a = 0, while for a maximal Kerr black hole with r = 1.2 and a = 0.9982, the estimated mass of SMBH is $6.52 \times 10^{12}\,{\rm M}_{\odot}$ . The estimated masses of SMBH significantly exceed the black hole mass calculated in Section 7. Therefore, the observed flux modulations in both bands are unlikely to be associated with this scenario.

As previously discussed, blazar emission is primarily jet-dominated, and variability in jet emission can also arise due to geometric effects. One such scenario involves a plasma blob moving along a helical trajectory within the jet, causing observed emission variations (Villata & Raiteri Reference Villata and Raiteri1999; Rieger Reference Rieger2004; Li et al. Reference Li, Xie, Chen, Dai, Lei, Yi and Ren2009; Li et al. Reference Li, Chen, Yi, Jiang, Chen, Lü and Li2015; Mohan & Mangalam Reference Mohan and Mangalam2015; Sobacchi, Sormani, & Stamerra Reference Sobacchi, Sormani and Stamerra2017; Zhou et al. Reference Zhou, Wang, Chen, Wiita, Vadakkumthani, Morrell, Zhang and Zhang2018; Otero-Santos et al. Reference Otero-Santos2020; Gong et al. Reference Gong, Zhou, Yuan, Zhang, Yi and Fang2022b; Gong et al. Reference Gong, Tian, Zhou, Yi and Fang2023; Prince et al. Reference Prince2023; Sharma et al. Reference Sharma, Kamaram, Prince, Khatoon and Bose2024). In this geometrical model, emission is enhanced due to relativistic beaming, and as the blob follows a helical path within the jet, the viewing angle changes over time. Such helical jet trajectories can result from jet bending, wiggling motion, or the presence of helical magnetic fields within the jet, leading to periodic emission patterns. Consequently, the blob’s motion causes a continuous variation in the viewing angle relative to the observer’s line of sight, which can be characterised as

(26) \begin{equation} \mathrm{cos} \ \theta_{{obs}}(t)= \mathrm{sin} \phi \ \mathrm{sin} \psi \ \mathrm{cos} (2\pi t / P_{{obs}}) \ + \ \mathrm{cos} \phi \ \mathrm{cos} \psi\end{equation}

where $\phi$ represents the pitch angle between the blob motion and the jet axis, $\psi$ is the viewing angle or inclination angle between the observer’s line of sight and jet axis, and $P_{{obs}}$ represents the observed period of oscillations in the light curve. Due to the helical motion of emitting region in the jet, the Doppler factor undergoes temporal variations, defined as

(27) \begin{equation} \delta = \frac{1}{\Gamma \left(1 \ -\ \beta \ \mathrm{cos} \theta_{{obs}}(t)\right)}\end{equation}

where $\Gamma = \frac{1}{\sqrt{1 - \beta^2}}$ represents the bulk Lorentz factor and $\beta = v_{\mathrm{jet}/c}$ . Following this, the observed flux is defined as $F_{\nu} \propto \delta^3$ . The relationship between the observed period and rest frame period of blob can be defined by the following expression:

(28) \begin{equation} P_{{rest}}=\frac{P_{{obs}}}{1 \ - \ \beta \ {\cos} \psi \ \mathrm{cos} \phi}\end{equation}

The blazer ON 246 is BL Lac type, therefore, we considered typical values of $\phi=2^{\circ}, \ \psi = 5^{\circ}$ (Sobacchi, Sormani, & Stamerra Reference Sobacchi, Sormani and Stamerra2017; Zhou et al. Reference Zhou, Wang, Chen, Wiita, Vadakkumthani, Morrell, Zhang and Zhang2018; Sarkar et al. Reference Sarkar2019; Prince et al. Reference Prince2023), and $\Gamma=4.85$ as estimated above. The period of oscillations in the rest frame to be $P_{{rest}} \sim$ 78.85 yr.

As discussed earlier, Rieger (Reference Rieger2004) outlined three potential mechanisms, one of which suggests that if the observed periodicity exceeds 1 yr, it can be reasonably attributed to the helical motion of the jet driven by the orbital motion of an SMBBH system. Using the expression in equation (16) and periodicity in rest frame, we estimate the mass of the primary black hole to be $M \sim 1.083 \times 10^9, \ 4.314 \times 10^9, \ 1.71 \times 10^{10}\,{\rm M}_{\odot}$ for $R=1,10$ , and 100, respectively. The estimated primary black hole mass in SMBBH for $R=1,10$ is consistent with the derived mass value utilising different approaches as described in Section 7. Therefore, our comprehensive analysis suggests that the observed periodicity is most likely attributed to the jet being driven by the orbital motion of the SMBBH system.

9. Conclusion

In this study, we investigated the $\gamma$ -ray and optical emission of the blazar ON 246 over the period of 11.6 yr (MJD 55932–60081). The key findings of our analysis are summarised as follows:

  • We identified a potential quasi-periodic signal of approximately 746 days in the $\gamma$ -ray and optical emissions of blazar ON 246 with a significance level exceeding 3 $\sigma$ .

  • Cross-correlation analysis reveals a strong correlation between $\gamma$ -ray and optical emissions, indicating a common origin for both. Additionally, we assessed the significance of the peak in the cross-correlation plot, finding it to be at the 4 $\sigma$ level.

  • We estimated the black hole mass range of ON 246 to be $(0.142 - 8.22) \times 10^9\,{\rm M}_{\odot}$ based on the shortest variability timescale observed in the $\gamma$ -ray band. Furthermore, using the rest-frame damping timescales in the $\gamma$ -ray and optical emissions, the black hole masses were found to be $(7.3 \pm 6) \times 10^9\,{\rm M}_{\odot}$ and $(1.56 \pm 0.66) \times 10^8\,{\rm M}_{\odot}$ , respectively. Both estimated values fall within the derived black hole mass range.

  • To explain the observed QPO, we explored various scenarios that could potentially account for the flux modulation with a period of $\sim$ 746 days in both light curves. Our analysis suggests that a non-ballistic jet driven by the orbital motion of a close supermassive binary black hole (SMBBH) system is a plausible explanation for the long-term periodic variations observed in the light curves.

Acknowledgments

A. Sharma is grateful to Prof. Sakuntala Chatterjee at S.N. Bose National Centre for Basic Sciences for providing the necessary support to conduct this research. RP acknowledges the support of the IoE seed grant from BHU.

Data availability statement

This research utilises publicly available data of ON 246 obtained from the Fermi-LAT data server provided by NASA Goddard Space Flight Center (GSFC): https://fermi.gsfc.nasa.gov/ssc/data/access/ and ASAS-SN observations from https://asas-sn.osu.edu.

Appendix A. Smooth bending power-law model

In addition to the power-law and stochastic (DRW) model, we assessed the local significance of QPOs in both gamma-ray and optical light curves by generating 50 000 synthetic light curves based on the smoothly bending power-law model. The functional form of this model is given by:

(A1) \begin{equation} P(f)=A \frac{f^{-\alpha_{{low}}}}{1+\left(\frac{f}{f_{{bend}}}\right)^{\alpha_{{high}} - \alpha_{{low}}}} + C\end{equation}

In this model, the parameters A, $\alpha_{{low}}$ , $\alpha_{{high}}$ , $f_{{bend}}$ , and C represent the normalisation, low-frequency slope, high-frequency slope, bend frequency, and Poisson noise level, respectively (Emmanoulopoulos et al. 2013). To model the power spectral densities (PSDs), we employed the curve $\_$ fit function from the Scipy package.Footnote j The best-fit parameters for the optical PSD are: $A = 22.07$ , $\alpha_{{low}} = 0.27$ , $\alpha_{{high}} = 5.23$ , $f_{{bend}} = 0.00044\;\mathrm{day^{-1}}$ , and $C = 0.73$ . For the gamma-ray PSD, the corresponding values are: $A = 128931.9$ , $\alpha_{{low}} = -0.96$ , $\alpha_{{high}} = 3.74$ , $f_{{bend}} = 0.0017\;\mathrm{day^{-1}}$ , and $C = 7.04$ . To assess the local significance of the dominant peaks in the original LSPs, we followed the procedure outlined in Section 5. The dominant gamma-ray LSP peak was found to have a significance level of approximately $3\sigma$ , while the prominent peak in the optical LSP exceeded the $3.48\sigma$ significance threshold, see Figure A1.

Figure A1. The top panels, from left to right, display the power spectral densities (PSDs) of the optical and gamma-ray light curves (in black), each fitted with a smooth bending power-law (BPL) model shown in red. The bottom panels present the corresponding local significance levels derived using the BPL modelling. In the bottom left panel (optical LSP), the dominant peak at $\sim0.00132 \ \mathrm{day^{-1}}$ reaches a significance level of $3.48\sigma$ . In the bottom right panel (gamma-ray LSP), the dominant peak at $\sim0.00134 \ \mathrm{day^{-1}}$ shows a significance level of $3\sigma$ . The blue dashed vertical lines in both bottom panels indicate the locations of the dominant peaks.

References

Acharyya, A., et al. 2023, ApJ, 950, 152.Google Scholar
Ackermann, M., et al. 2015, ApJL, 813, L41.Google Scholar
Adhikari, S., Penil, P., Westernacher-Schneider, J. R., Dominguez, A., Ajello, M., Buson, S., Rico, A., & Zrake, J. 2024, ApJ, 965, 124.Google Scholar
Aleksic, J., et al. 2011, ApJL, 730, L8.Google Scholar
Alston, W. N., Parker, M. L., Markevičiute, J., Fabian, A. C., Middleton, M., Lohfink, A., Kara, E., & Pinto, C. 2015, MNRAS, 449, 467.Google Scholar
Atwood, W. B., et al. 2009, ApJ, 697, 1071.Google Scholar
Aydin, M. E. 2017, Eaydin/wwz: first release. V. v1.0.0, March. https://doi.org/10.5281/zenodo.375648.Google Scholar
Bade, N., Fink, H. H., & Engels, D. 1994, A&A 286, 381.Google Scholar
Banerjee, A., Sharma, A., Mandal, A., Kumar Das, A., Bhatta, G., & Bose, D. 2023, MNRAS, 523, L52.Google Scholar
Begelman, M. C., Blandford, R. D., & Rees, M. J. 1980, Natur, 287, 307.Google Scholar
Benbow, W., et al. 2017, in AIP Conference Proceedings, vol. 1792. 1 (AIP Publishing).Google Scholar
Benkhali, F. A., Hofmann, W., Rieger, F. M., & Chakraborty, N. 2020, A&A, 634, A120.Google Scholar
Bhatta, G., et al. 2016, ApJ, 832, 47.Google Scholar
Bhatta, G. 2021, ApJ, 923, 7.Google Scholar
Bhatta, G., & Dhital, N. 2020, ApJ, 891, 120.Google Scholar
Blandford, R., Meier, D., & Readhead, A. 2019, ARA&A, 57, 467.Google Scholar
Blandford, R. D., & McKee, C. F. 1982, ApJ, 255, 419.Google Scholar
Blandford, R. D., & Payne, D. G. 1982, MNRAS, 199, 883.Google Scholar
Blandford, R. D., & Znajek, R. L. 1977, MNRAS, 179, 433.Google Scholar
Burke, C. J., et al. 2021, Sci, 373, 789.Google Scholar
Carnerero, M. I., et al. 2017, MNRAS, 472, 3789.Google Scholar
Cash, W. 1979, ApJ, 228, 939.Google Scholar
Castignani, G., Haardt, F., Lapi, A., De Zotti, G., Celotti, A., & Danese, L. 2013, A&A, 560, A28. https://doi.org/10.1051/0004-6361/201321424. arXiv: https://arxiv.org/abs/1309.4108 1309.4108 [astro-ph.CO].Google Scholar
Chen, B.-Y., Bower, G. C., Dexter, J., Markoff, S., Ridenour, A., Gurwell, M. A., Rao, R., & Wallström, S. H. J. 2023, ApJ, 951, 93.Google Scholar
Cheng, K. S., Fan, J. H., & Zhang, L. 1999, arXiv preprint astro-ph/9910543.Google Scholar
Cohen, D. P., Romani, R. W., Filippenko, A. V., Bradley Cenko, S., Lott, B., Zheng, W. K., & Li, W. 2014, ApJ, 797, 137.Google Scholar
Das, A. K., Prince, R., Gupta, A. C., & Kushwaha, P. 2023, ApJ, 950, 173.Google Scholar
Dixon, R. S., & Kraus, J. D. 1968, AJ, 73, 381.Google Scholar
Emmanoulopoulos, D., McHardy, I. M., & Papadakis, I. E. 2013 MNRAS, 433, 907.Google Scholar
Fan, J. H., et al. 2016, ApJS, 226, 20.Google Scholar
Fan, J. H., Xie, G. Z., & Bacon, R. 1999, A&AS, 136, 13.Google Scholar
Fan, J.-H. 2005, A&A, 436, 799.Google Scholar
Fan, J.-H., et al. 2009, RAA, 9, 538.Google Scholar
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125, 306.Google Scholar
Fossati, G., Maraschi, L., Celotti, A., Comastri, A., & Ghisellini, G. 1998, MNRAS, 299, 433.Google Scholar
Foster, G. 1996, AJ, 112, 1709.Google Scholar
Genzel, R., Eckart, A., Ott, Th, & Eisenhauer, F. 1997, MNRAS, 291, 219.Google Scholar
Ghisellini, G., & Madau, P. 1996, MNRAS, 280, 67.Google Scholar
Ghisellini, G., Padovani, P., Celotti, A., & Maraschi, L. 1993, ApJ, 407, 65.Google Scholar
Ghisellini, G., & Tavecchio, F. 2008, MNRAS, 387, 1669. https://doi.org/10.1111/j.1365-2966.2008.13360.x. arXiv: https://arxiv.org/abs/0802.1918 0802.1918 [astro-ph].Google Scholar
Gierlinski, M., Middleton, M., Ward, M., & Done, C. 2008, Natur, 455, 369.Google Scholar
Giommi, P., et al. 1999, arXiv preprint astro-ph/9909241.Google Scholar
Giommi, P., Padovani, P., Polenta, G., Turriziani, S., D’Elia, V., & Piranomonte, S. 2012, MNRAS, 420, 2899.Google Scholar
Gong, Y., et al. 2024, ApJ, 976, 51.Google Scholar
Gong, Y., Tian, S., Zhou, L., Yi, T., & Fang, J. 2023, ApJ, 949, 39.Google Scholar
Gong, Y., Zhou, L., Yuan, M., Zhang, H., Yi, T., & Fang, J. 2022a, ApJ, 931, 168. https://doi.org/10.3847/1538-4357/ac6c8c.Google Scholar
Gong, Y., Zhou, L., Yuan, M., Zhang, H., Yi, T., & Fang, J. 2022b, ApJ, https://doi.org/10.3847/1538-4357/ac6c8c931, 168.Google Scholar
Graham, M. J., et al. 2015, Natur, 518, 74.Google Scholar
Grossmann, A., & Morlet, J. 1984, SIAM JMA, 15, 723.Google Scholar
Gupta, A. C., et al. 2008, AJ, 136, 2359.Google Scholar
Gupta, A. C. 2017, Galaxies, 6, 1.Google Scholar
Gupta, A. C., et al. 2019, AJ, 157, 95.Google Scholar
Gupta, A. C., Srivastava, A. K., & Wiita, P. J. 2008, ApJ, 690, 216.Google Scholar
Gupta, A. C., Tripathi, A., Wiita, P. J., Kushwaha, P., Zhang, Z., & Bambi, C. 2019, MNRAS, 484, 5785.Google Scholar
Haiyan, Y., Xiefei, S., Xiaopan, L., Na, J., Haitao, Y., Yuhui, L., Li, Z., & Yan, C. 2023, Ap&SS, 368, 88.Google Scholar
Hart, K., et al. 2023, arXiv preprint arXiv:2304.03791.Google Scholar
Hartman, R. C. 1996, in Blazar Continuum Variability Astronomical Society of the Pacific Conference Series 110, Proceedings of an International Workshop held at Florida International University, Miami, FL, USA, 4–7 February 1996, San Francisco: Astronomical Society Pacific, ed. Richard Miller, H., Webb, J. R., & Noble, J. C., 333.Google Scholar
Huang, C.-Y., Wang, D.-X., Wang, J.-Z., & Wang, Z.-Y. 2013, RAA, 13, 705.Google Scholar
Jorstad, S. G., et al. 2022, Natur, 609, 265.Google Scholar
Kaspi, S., Smith, P. S., Netzer, H., Maoz, D., Jannuzi, B. T., & Giveon, U. 2000, ApJ, 533, 631.Google Scholar
Katarzynski, K., Sol, H., & Kus, A. 2001, A&A, 367, 809.Google Scholar
Kelly, B. C., Bechtold, J., & Siemiginowska, A. 2009, ApJ, 698, 895.Google Scholar
Kelly, B. C., Becker, A. C., Sobolewska, M., Siemiginowska, A., & Uttley, P. 2014, ApJ, 788, 33.Google Scholar
Kharb, P., Gabuzda, D., & Shastri, P. 2008, MNRAS, 384, 230.Google Scholar
King, O. G., et al. 2013, MNRAS, 436, L114.Google Scholar
Kochanek, C. S., et al. 2017, PASP, 129, 104502.Google Scholar
Kozłowski, S., et al. 2009, ApJ, 708, 927.Google Scholar
Lachowicz, P., Gupta, A. C., Gaur, H., & Wiita, P. J. 2009, A&A, 506, L17.Google Scholar
Li, H. Z., Chen, L. E., Yi, T. F., Jiang, Y. G., Chen, X., , L. Z., & Li, K. Y. 2015, PASP, 127, 1. https://doi.org/10.1086/679753.Google Scholar
Li, H. Z., Jiang, Y. G., Guo, D. F., Chen, X., & Yi, T. F. 2016, PASP, 128, 074101.Google Scholar
Li, H. Z., Xie, G. Z., Chen, L. E., Dai, H., Lei, B. Y., Yi, T. F., & Ren, J. Y. 2009, PASP, 121, 1172. https://doi.org/10.1086/648433.Google Scholar
Li, H.-Z., Qin, L.-H., Gong, Y.-L., Liu, F., Guo, D.-F., Gao, Q.-G., Yi, T.-F., & Liu, H.-T. 2024, MNRAS, 534, 2986.Google Scholar
Li, S., Xia, Z.-Q., Liang, Y.-F., Liao, N.-H., & Fan, Y.-Z. 2018, ApJ, 853, 159.Google Scholar
Li, X.-P., et al. 2023, MNRAS, 519, 4893.Google Scholar
Li, X.-P., Cai, Y., Yang, H.-T., Luo, Y.-H., Yan, Y., He, J.-Y., & Wang, L.-S. 2021, MNRAS, 506, 1540.Google Scholar
Liao, N. H., Bai, J. M., Liu, H. T., Weng, S. S., Chen, L., & Li, F. 2014, ApJ, 783, 83.Google Scholar
Liu, H. T., & Bai, J. M. 2015, AJ, 149, 191.Google Scholar
Lomb, N. R. 1976, Ap&SS, 39, 447.Google Scholar
Lu, L., Sun, B., Fang, Z.-X., Wan, M., & Gong, Y. 2024, ApJ, 961, 180.Google Scholar
MacLeod, C. L., et al. 2012, ApJ, 753, 106.Google Scholar
Malmrose, M. P., Marscher, A. P., Jorstad, S. G., Nikutta, R., & Elitzur, M. 2011, ApJ, 732, 116.Google Scholar
Mangalam, A. V., & Wiita, P. J. 1993, ApJ, 406, 420.Google Scholar
Mao, L., & Zhang, X. 2024, MNRAS, 531, 3927.Google Scholar
Maraschi, L., Ghisellini, G., & Celotti, A. 1992, ApJ, 397, L5.Google Scholar
Mattox, J. R., et al. 1996, ApJ, 461, 396.Google Scholar
Megan, U. C., & Padovani, P. 1995, PASP, 107, 803.Google Scholar
Meier, D. L., Koide, S., & Uchida, Y. 2001, Sci, 291, 84.Google Scholar
Mohan, P., & Mangalam, A. 2015, ApJ, 805, 91.Google Scholar
Moreno, J., Vogeley, M. S., Richards, G. T., & Yu, W. 2019, PASP, 131, 063001.Google Scholar
Otero-Santos, J., et al. 2020, MNRAS, 492, 5524.Google Scholar
Otero-Santos, J., Peñil, P., Acosta-Pulido, J. A., Becerra González, J., Raiteri, C. M., Carnerero, M. I., & Villata, M. 2023, MNRAS, 518, 5788.Google Scholar
Padovani, P. 2017, FASS, 4, 35.Google Scholar
Paiano, S., Landoni, M., Falomo, R., Treves, A., Scarpa, R., & Righi, C. 2017, ApJ, 837, 144.Google Scholar
Paliya, V. S., Domnguez, A., Ajello, M., Olmo-Garca, A., & Hartmann, D. 2021, ApJS, 253, 46. https://doi.org/10.3847/1538-4365/abe135.Google Scholar
Pan, H.-W., Yuan, W., Yao, S., Zhou, X.-L., Liu, B., Zhou, H., & Zhang, S.-N. 2016, ApJL, 819, L19.Google Scholar
Pandey, A., & Stalin, C. S. 2022, A&A, 668, A152.Google Scholar
Pauliny-Toth, I. I. K., Kellermann, K. I., Davis, M. M., Fomalont, E. B., & Shaffer, D. B. 1972, AJ, 77, 265.Google Scholar
Pei, Z., Fan, J., Yang, J., Huang, D., & Li, Z. 2022, ApJ, 925, 97.Google Scholar
Peñil, P., et al. 2020, ApJ, 896, 134.Google Scholar
Peterson, B. M., et al. 2004, ApJ, 613, 682.Google Scholar
Peterson, B. M., Wanders, I., Horne, K., Collier, S., Alexander, T., Kaspi, S., & Maoz, D. 1998, PASP, 110, 660.Google Scholar
Petry, D., et al. 2000, ApJ, 536, 742.Google Scholar
Prince, R., et al. 2023, A&A, 678, A100.Google Scholar
Raiteri, C. M., et al. 2021a, MNRAS, 501, 1100.Google Scholar
Raiteri, C. M., et al. 2021b, MNRAS, 504, 5629.Google Scholar
Ren, C., Sun, S., & Zhang, P. 2024, ApJ, 961, 38.Google Scholar
Ren, G.-W., Ding, N., Zhang, X., Xue, R., Zhang, H.-J., Xiong, D.-R., Li, F.-T., & Li, H. 2021a, MNRAS, 506, 3791.Google Scholar
Ren, G.-W., Zhang, H.-J., Zhang, X., Ding, N., Yang, X., Li, F.-T., Yan, P.-L., & Xu, X.-L. 2021b, RAA, 21, 075.Google Scholar
Ren, H. X., Cerruti, M., & Sahakyan, N. 2023, A&A, 672, A86.Google Scholar
Rieger, F. M. 2004, ApJ, 615, L5.Google Scholar
Rieger, F. M. 2005, ChJAA, 5, 305.Google Scholar
Roland, J., Britzen, S., Caproni, A., Fromm, C., Glück, C., & Zensus, A. 2013, A&A, 557, A85. https://doi.org/10.1051/0004-6361/201219165. arXiv: https://arxiv.org/abs/1307.3700 [astro-ph.HE].Google Scholar
Romero, G. E., Chajet, L. S., Abraham, Z., & Fan, J. H. 2000, A&A, 360.Google Scholar
Roy, A., et al. 2022, MNRAS, 513, 5238.Google Scholar
Roy, A., Sarkar, A., Chatterjee, A., Gupta, A. C., Chitnis, V., & Wiita, P. J. 2022, MNRAS, 510, 3641.Google Scholar
Ruan, J. J., et al. 2012, ApJ, 760, 51.Google Scholar
Ryan, J. L., Siemiginowska, A., Sobolewska, M. A., & Grindlay, J. 2019, ApJ, 885, 12.Google Scholar
Sandrinelli, A., Covino, S., Dotti, M., & Treves, A. 2016, AJ, 151, 54.Google Scholar
Sandrinelli, A., Covino, S., & Treves, A. 2014, A&A, 562, A79.Google Scholar
Sarkar, A., et al. 2019, ApJ, 887, 185.Google Scholar
Sarkar, A., Gupta, A. C., Chitnis, V. R., & Wiita, P. J. 2021, MNRAS, 501, 50.Google Scholar
Scargle, J. D. 1979, National Aeronautics/Space Administration, Vol. 81148, Ames Research Center.Google Scholar
Schulz, M., & Mudelsee, M. 2002, C&G, 28, 421.Google Scholar
Shappee, B. J., et al. 2014, ApJ, 788, 48.Google Scholar
Sharma, A., Banerjee, A., Das, A. K., Mandal, A., & Bose, D. 2024, ApJ, 975, 56.Google Scholar
Sharma, A., Kamaram, S. R., Prince, R., Khatoon, R., & Bose, D. 2024, MNRAS, 527, 2672.Google Scholar
Sharma, A., Prince, R., & Bose, D. 2024, arXiv preprint arXiv:2410.06653.Google Scholar
Sillanpaa, A., Haarala, S., Valtonen, M. J., Sundelius, B., & Byrd, G. G. 1988, ApJ, 325, 628.Google Scholar
Smith, E., Oramas, L., & Perlman, E. 2023, ApJ, 950, 174.Google Scholar
Sobacchi, E., Sormani, M. C., & Stamerra, A. 2017, MNRAS, 465, 161. https://doi.org/10.1093/mnras/stw2684. arXiv: https://arxiv.org/abs/1610.04709 [astro-ph.HE].Google Scholar
Sobolewska, M. A., Siemiginowska, A., Kelly, B. C., & Nalewajko, K. 2014, ApJ, 786, 143.Google Scholar
Sołtan, A. 1982, MNRAS, 200, 115.Google Scholar
Stella, L., & Vietri, M. 1997, ApJ, 492, L59.Google Scholar
Sun, M., Grier, C. J., & Peterson, B. M. 2018, Astrophysics Source Code Library, record ascl:1805.032, May.Google Scholar
Tagliaferri, G., Ravasio, M., Ghisellini, G., Tavecchio, F., Giommi, P., Massaro, E., Nesci, R., & Tosti, G. 2003, A&A, 412, 711.Google Scholar
Tantry, J., Sharma, A., Shah, Z., Iqbal, N., & Bose, D. 2025, JHEA, 100372.Google Scholar
Tavani, M., Cavaliere, A., Munar-Adrover, P., & Argan, A. 2018, ApJ, 854, 11.Google Scholar
Tavecchio, F., Maraschi, L., & Ghisellini, G. 1998, ApJ, 509, 608.Google Scholar
Tripathi, A., Gupta, A. C., Aller, M. F., Wiita, P. J., Bambi, C., Aller, H., & Gu, M. 2021, MNRAS, 501, 5997.Google Scholar
Ulrich, M.-H., Maraschi, L., & Urry, C. M. 1997, ARA&A, 35, 445.Google Scholar
Urry, C. M., et al. 1993, ApJ, 411, 614.Google Scholar
Valtonen, M. J., et al. 2008, Natur, 452, 851.Google Scholar
Valtonen, M. J., Lehto, H. J., Takalo, L. O., & Sillanpää, A. 2011, ApJ, 729, 33.Google Scholar
VanderPlas, J. T. 2018, ApJS, 236, 16.Google Scholar
Vestergaard, M. 2002, ApJ, 571, 733.Google Scholar
Villata, M., & Raiteri, C. M. 1999, A&A, 347, 30.Google Scholar
Villforth, C., et al. 2010, MNRAS, 402, 2087.Google Scholar
Wagner, S. J., & Witzel, A. 1995, ARA&A, 33, 163 Google Scholar
Wang, G. G., Cai, J. T., & Fan, J. H. 2022, ApJ, 929, 130.Google Scholar
Wang, H., Yin, C., & Xiang, F. 2017, Ap&SS, 362, 19.Google Scholar
Wang, J.-Y., An, T., Baan, W. A., & Lu, X.-L. 2014, MNRAS, 443, 58. issn: 0035-8711. https://doi.org/10.1093/mnras/stu1135.Google Scholar
Wu, Z.-Z., Gu, M.-F., & Jiang, D.-R. 2009, RAA, 9, 168.Google Scholar
Xie, G. Z., Bai, J. M., Zhang, X., & Fan, J. H. 1998, A&A, 334, L29.Google Scholar
Xie, G. Z., Yi, T. F., Li, H. Z., Zhou, S. B., & Chen, L. E. 2008, AJ, 135, 2212.Google Scholar
Xiong, D. R., & Zhang, X. 2014, MNRAS, 441, 3375. issn: 0035-8711. https://doi.org/10.1093/mnras/stu755.Google Scholar
Yang, J., Cao, G., Zhou, B., & Qin, L. 2021, PASP, 133, 024101. https://doi.org/10.1088/1538-3873/abd152.Google Scholar
Yang, J. H., & Fan, J. H. 2010, SCPMA, 53, 1921.Google Scholar
Yang, S., Yan, D., Zhang, P., Dai, B., & Zhang, L. 2021, ApJ, 907, 105.Google Scholar
Zhang, H., Yan, D., & Zhang, L. 2022, ApJ, 930, 157.Google Scholar
Zhang, H., Yan, D., & Zhang, L. 2023, ApJ, 944, 103.Google Scholar
Zhang, H., Yang, S., & Dai, B. 2024, ApJL, 967, L18.Google Scholar
Zhang, P.-F., Yan, D.-H., Zhou, J.-N., Fan, Y.-Z., Wang, J.-C., & Zhang, L. 2017, ApJ, 845, 82.Google Scholar
Zhang, X., Xiong, D.-r., Gao, Q.-g., Yang, G.-q., Lu, F.-w., Na, W.-w., & Qin, L.-h. 2024, MNRAS, 529, 3699. issn: 0035-8711. https://doi.org/10.1093/mnras/stae765.Google Scholar
Zhang, X., & Bao, G. 1990, Technical report. International Centre for Theoretical Physics.Google Scholar
Zheng, Y. G., Zhang, L., Huang, B. R., & Kang, S. J. 2013, MNRAS, 431, 2356.Google Scholar
Zhou, J., Wang, Z., Chen, L., Wiita, P. J., Vadakkumthani, J., Morrell, N., Zhang, P., & Zhang, J. 2018, NatCo, 9, 4599.Google Scholar
Zhou, R. X., Zheng, Y. G., Zhu, K. R., & Kang, S. J. 2021, ApJ, 915, 59.Google Scholar
Zu, Y., Kochanek, C. S., Kozowski, S., & Udalski, A. 2013, ApJ, 765, 106.Google Scholar
Figure 0

Figure 1. The figure presents the $\gamma$-ray and optical light curves observed between MJD 55900 and 60150. The top panel shows the 10-day binned $\gamma$-ray flux (blue points), with the Bayesian block (BB) representation overlaid as a black curve for illustrative purposes only. The bottom panel shows the ASAS-SN optical light curve (green) with the corresponding BB representation (black curve). The grey horizontal lines in both panels indicate the mean $\gamma$-ray flux and optical magnitude, respectively.

Figure 1

Figure 2. The $\gamma$-ray light curve is analysed using the Lomb-Scargle Periodogram (LSP) and Weighted Wavelet Z-transform (WWZ) methods. The top panel shows the local significance of the detected peak at $\sim$0.00134 day$^{-1}$ in $\gamma$-ray LSP is exceeding 99.73$\%$. The bottom panels display the WWZ map (left) and average wavelet power (right). The observed local significance of the detected peak at $\sim$0.00132 day$^{-1}$ in avg. wavelet has a significance level of $99.73\%$.

Figure 2

Figure 3. The detected QPO signals in the optical emissions from ON 246. The top panel shows the LSP with a dominant peak at $\sim$0.00132 day$^{-1}$ has a local significance level exceeding 99.73$\%$. The bottom panels display the wavelet map (bottom left panel) and avg. wavelet power at frequency of $\sim0.00131$ day$^{-1}$ with a significance level greater than 99.73$\%$.

Figure 3

Figure 4. Analysis of the light curves, left panel represent the REDFIT curve of $\gamma$-ray emissions and right panel exhibit the REDFIT curve of optical emissions, using the AR(1) process with the REDFIT software. The red noise-corrected power spectrum (black) is presented alongside theoretical (blue) and average AR(1) (cyan) spectra. The significance levels of 99$\%$, 95$\%$, and 90$\%$ are indicated in red, green, and brown, respectively.

Figure 4

Figure 5. The figure displays the posterior probability distributions of the DRW model parameters, obtained from the $\gamma$-ray light curve (left panel) and the ASAS-SN light curve (right panel).

Figure 5

Figure 6. The celerite modelling with DRW model was performed using the 10-day binned $\gamma$-ray light curve of blazar ON 246 over 4 000 days from the time stamp MJD 55932. In this figure, the top panel depicts the $\gamma$-ray flux points with their uncertainties, along with the best-fitting profile of the DRW model in blue, including the 1$\sigma$ confidence interval. The bottom panels represent the autocorrelation functions (ACFs) of the standardised residuals (bottom left) and the squared of standardised residuals (bottom right), respectively, along with 95$\%$ confidence intervals of the white noise.

Figure 6

Figure 7. This figure demonstrates the modelling of the ASAS-SN light curve of ON 246 with the DRW model. The top panel shows the best-fitting profile of the DRW modelling along with 1$\sigma$ confidence interval. The bottom panels represent the autocorrelation functions as described in Figure 6.

Figure 7

Figure 8. This figure presents the DRW PSDs obtained from $\gamma$-ray and ASAS-SN observations, along with their 1$\sigma$ confidence intervals. The two shaded regions highlight biased regions due to observational limitations, such as finite length and the mean cadence of the light curve. The regions with orange star symbols represent the invalid areas in $\gamma$-ray PSD, while the blue hatch line regions in PSD indicate limitations imposed by the ASAS-SN light curve’s duration and mean cadence.

Figure 8

Figure 9. The Lomb-Scargle periodograms of the $\gamma$-ray (left panel) and ASAS-SN (right panel) observations of the blazar ON 246 are shown. The figure presents the LSPs of the original light curves (black) and spectral windows (pink). The significance levels (blue) of the dominant observed peaks in both periodograms are estimated using DRW synthetic light curves. The detected peaks at 0.00105 day$^{-1}$ ($\sim$950 days) and 0.00105 day$^{-1}$ ($\sim$950 days) in $\gamma$-ray and ASAS-SN observations exceed 3$\sigma$ and 4$\sigma$ significance levels, respectively. The shaded regions in both figures represent the invalid areas estimated using the mean cadence and baseline of the light curves. Logarithmic versions of the periodograms are provided in the sub-figures.

Figure 9

Table 1. Summary of the quasi-periodic signal detection using three different methodologies. Column (1) lists the 4FGL name of the blazar ON 246, while Column (2) specifies the observational waveband. Columns (3) and (4) present the QPO frequencies obtained from the Lomb-Scargle periodogram and Weighted Wavelet Z-transform methods, along with their uncertainties. The local significance of each detected QPO signal is provided in parentheses next to the corresponding frequency value. Column (5) presents the estimated local significance level of the QPO in LSP based on DRW-modeled mock light curves, and Column (6) provides the QPO frequency and significance level derived from the REDFIT analysis.

Figure 10

Figure 10. Cross-correlation analysis between $\gamma$-ray and optical flux variations. The left panel shows ICCF (solid black curve) with a significance level (dash red curve) of 4$\sigma$. The cross-correlation centroid distribution (CCCD) in orange and cross-correlation peak distribution (CCPD) in blue are given in the right panel, where vertical dashes in orange and blue represent the median of CCCD and CCPD, respectively.

Figure 11

Figure 11. The folded Fermi-LAT and ASAS-SN light curves of ON 246 with a period of 746 and 757 days are shown in the top and bottom panels, respectively. The dashed blue line represents the mean value, and the sine functions (red) with frequencies of 0.00134 and 0.00132 day$^{-1}$ were fitted to the folded $\gamma$-ray and optical light curves, respectively. Two full period cycles are shown for better clarity.

Figure 12

Figure A1. The top panels, from left to right, display the power spectral densities (PSDs) of the optical and gamma-ray light curves (in black), each fitted with a smooth bending power-law (BPL) model shown in red. The bottom panels present the corresponding local significance levels derived using the BPL modelling. In the bottom left panel (optical LSP), the dominant peak at $\sim0.00132 \ \mathrm{day^{-1}}$ reaches a significance level of $3.48\sigma$. In the bottom right panel (gamma-ray LSP), the dominant peak at $\sim0.00134 \ \mathrm{day^{-1}}$ shows a significance level of $3\sigma$. The blue dashed vertical lines in both bottom panels indicate the locations of the dominant peaks.