A&A 647, A163 (2021) https://doi.org/10.1051/0004-6361/202039687 c© ESO 2021 Astronomy &Astrophysics VHE gamma-ray detection of FSRQ QSO B1420+326 and modeling of its enhanced broadband state in 2020 MAGIC Collaboration: V. A. Acciari1, S. Ansoldi2, L. A. Antonelli3, A. Arbet Engels4, M. Artero5, K. Asano6, D. Baack7, A. Babic´8, A. Baquero9, U. Barres de Almeida10, J. A. Barrio9, J. Becerra González1, W. Bednarek11, L. Bellizzi12, E. Bernardini13, M. Bernardos14, A. Berti15, J. Besenrieder16, W. Bhattacharyya13, C. Bigongiari3, A. Biland4, O. Blanch5, G. Bonnoli12, Ž. Bošnjak8, G. Busetto14, R. Carosi17, G. Ceribella16, M. Cerruti18, Y. Chai16, A. Chilingarian19, S. Cikota8, S. M. Colak5, E. Colombo1, J. L. Contreras9, J. Cortina20, S. Covino3, G. D’Amico16, V. D’Elia3, P. Da Vela17,37, F. Dazzi3, A. De Angelis14, B. De Lotto2, M. Delfino5,38, J. Delgado5,38, C. Delgado Mendez20, D. Depaoli15, F. Di Pierro15, L. Di Venere21, E. Do Souto Espiñeira5, D. Dominis Prester22, A. Donini2, D. Dorner23, M. Doro14, D. Elsaesser7, V. Fallah Ramazani24, A. Fattorini7, G. Ferrara3, L. Foffano14, M. V. Fonseca9, L. Font25, C. Fruck16, S. Fukami6, R. J. García López1, M. Garczarczyk13, S. Gasparyan26, M. Gaug25, N. Giglietto21, F. Giordano21, P. Gliwny11, N. Godinovic´27, J. G. Green3, D. Green16, D. Hadasch6, A. Hahn16, L. Heckmann16, J. Herrera1, J. Hoang9, D. Hrupec28, M. Hütten16, T. Inada6, S. Inoue6, K. Ishio16, Y. Iwamura6, J. Jormanainen24, L. Jouvin5, Y. Kajiwara29, M. Karjalainen1, D. Kerszberg5, Y. Kobayashi6, H. Kubo29, J. Kushida30, A. Lamastra3, D. Lelas27, F. Leone3, E. Lindfors24, S. Lombardi3, F. Longo2,39, R. López-Coto14, M. López-Moya9, A. López-Oramas1, S. Loporchio21, B. Machado de Oliveira Fraga10, C. Maggio25, P. Majumdar31, M. Makariev32, M. Mallamaci14, G. Maneva32, M. Manganaro22, K. Mannheim23, L. Maraschi3, M. Mariotti14, M. Martínez5, D. Mazin6,40, S. Mender7, S. Mic´anovic´22, D. Miceli2, T. Miener9, M. Minev32, J. M. Miranda12, R. Mirzoyan16, E. Molina18, A. Moralejo5, D. Morcuende9, V. Moreno25, E. Moretti5, V. Neustroev33, C. Nigro5, K. Nilsson24, D. Ninci5, K. Nishijima30, K. Noda6, S. Nozaki29,?, Y. Ohtani6, T. Oka29, J. Otero-Santos1, S. Paiano3, M. Palatiello2, D. Paneque16, R. Paoletti12, J. M. Paredes18, L. Pavletic´22, P. Peñil9, C. Perennes14, M. Persic2,41, P. G. Prada Moroni17, E. Prandini14, C. Priyadarshi5, I. Puljak27, W. Rhode7, M. Ribó18, J. Rico5, C. Righi3, A. Rugliancich17, L. Saha9, N. Sahakyan26, T. Saito6, S. Sakurai6, K. Satalecka13, F. G. Saturni3, B. Schleicher23, K. Schmidt7, T. Schweizer16, J. Sitarek11,?, I. Šnidaric´34, D. Sobczynska11, A. Spolon14, A. Stamerra3, D. Strom16, M. Strzys6, Y. Suda16, T. Suric´34, M. Takahashi6, F. Tavecchio3, P. Temnikov32, T. Terzic´22, M. Teshima16,42, N. Torres-Albà18, L. Tosti35, S. Truzzi12, A. Tutone3, J. van Scherpenberg16, G. Vanzo1, M. Vazquez Acosta1, S. Ventura12, V. Verguilov32, C. F. Vigorito15, V. Vitale36, I. Vovk6, M. Will16, and D. Zaric´27, R. Angioni43,44,?, F. D’Ammando45,?, S. Ciprini43,46, C. C. Cheung47, M. Orienti45, L. Pacciani48, P. Prajapati49, P. Kumar49, S. Ganesh49, M. Minev50,51, A. Kurtenkov50, A. Marchini52, L. Carrasco53, G. Escobedo53, A. Porras53, E. Recillas53, A. Lähteenmäki54,55, M. Tornikoski54, M. Berton54,56, J. Tammi54, R. J. C. Vera54,55, S. G. Jorstad57,58, A. P. Marscher57, Z. R. Weaver57, M. Hart57, M. K. Hallum57, V. M. Larionov58,59,†, G. A. Borman60, T. S. Grishina58, E. N. Kopatskaya58, E. G. Larionova58, A. A. Nikiforova58,59, D. A. Morozova58, S. S. Savchenko58, Yu. V. Troitskaya58, I. S. Troitsky58, A. A. Vasilyev58, M. Hodges61, T. Hovatta56,54, S. Kiehlmann62,63, W. Max-Moerbeck64, A. C. S. Readhead61, R. Reeves65, and T. J. Pearson61 (Affiliations can be found after the references) Received 15 October 2020 / Accepted 18 December 2020 ABSTRACT Context. QSO B1420+326 is a blazar classified as a flat-spectrum radio quasar (FSRQ). At the beginning of the year 2020, it was found to be in an enhanced flux state and an extensive multiwavelength campaign allowed us to trace the evolution of the flare. Aims. We search for very high-energy (VHE) gamma-ray emission from QSO B1420+326 during this flaring state. We aim to characterize and model the broadband emission of the source over different phases of the flare. Methods. The source was observed with a number of instruments in radio, near-infrared, optical (including polarimetry and spectroscopy), ultra- violet, X-ray, and gamma-ray bands. We use dedicated optical spectroscopy results to estimate the accretion disk and the dust torus luminosity. We performed spectral energy distribution modeling in the framework of combined synchrotron-self-Compton and external Compton scenario in which the electron energy distribution is partially determined from acceleration and cooling processes. Results. During the enhanced state, the flux of both SED components of QSO B1420+326 drastically increased and the peaks were shifted to higher energies. Follow-up observations with the MAGIC telescopes led to the detection of VHE gamma-ray emission from this source, making it one of only a handful of FSRQs known in this energy range. Modeling allows us to constrain the evolution of the magnetic field and electron energy distribution in the emission region. The gamma-ray flare was accompanied by a rotation of the optical polarization vector during a low -polarization state. Also, a new superluminal radio knot contemporaneously appeared in the radio image of the jet. The optical spectroscopy shows a prominent FeII bump with flux evolving together with the continuum emission and a MgII line with varying equivalent width. Key words. gamma rays: galaxies – galaxies: jets – radiation mechanisms: non-thermal – quasars: individual: QSO B1420+326 ? Corresponding authors; e-mail: contact.magic@mpp.mpg.de † Deceased. Article published by EDP Sciences A163, page 1 of 19 A&A 647, A163 (2021) 1. Introduction QSO B1420+326, also known as OQ 334, is a blazar located at redshift of 0.682 (Hewett & Wild 2010). Based on its radio spectrum it has been classified as a flat-spectrum radio quasar (FSRQ; Healey et al. 2007). BL Lac objects are divided between low-, intermediate-, and high-synchrotron peaked (LSP, ISP, and HSP), while FSRQs are usually only LSP objects. At high energies (HE; i.e., GeV), FSRQs populate the majority of the extragalactic gamma-ray sky. Among the associated blazars in the Fermi-LAT Fourth Source Catalog (4FGL; Abdollahi et al. 2020) there are 694 FSRQs compared to 1131 BL Lac objects. In the very-high-energy (VHE, &100 GeV) range, despite the detection of over 60 BL Lac objects by Imaging Atmospheric Cherenkov Telescopes (IACTs), only about 8 FSRQs are known to emit in this energy range1. There are a few probable reasons that contribute to the difference between the number of blazars detected at HE and VHE. The peak of the gamma-ray emission in the spectral energy distribution (SED) of FSRQs is usually shifted to lower energies compared to BL Lac objects (see, e.g., Ghisellini 2016). Also, some of the source may have enhanced internal absorption in the radiation field of the broad line region (BLR; see e.g., Liu & Bai 2006, but note also Costamante et al. 2018) via the e+e− pair production process. FSRQs are consid- ered to be more luminous sources than BL Lacs, which means that they can be detected at greater cosmological distances. How- ever, for sources located at high redshift (z ∼ 1), the VHE gamma-ray part of the spectrum is severely absorbed in the pair- production process on extragalactic background light (EBL; see, e.g., Domínguez et al. 2011), hampering the discovery potential in this energy range. Softer gamma-ray spectra resulting from both effects make detection of FSRQs with IACTs even more difficult. Finally, FSRQs are known to be extremely variable (see, e.g., Meyer et al. 2019), which is another complication in observing these sources with instruments with relatively small fields of view such as IACTs. The VHE gamma-ray flux has been seen to vary by as much as two orders of magnitude (see, e.g., D’Ammando et al. 2019; Zacharias et al. 2019). Emission variability has been observed down to a timescale of ∼10 min (Aleksic´ et al. 2011). Due to the strong variability, the cur- rent most successful approach for studying the VHE gamma-ray emission of FSRQs is a follow up of alerts of enhanced activity at lower frequencies. To date, all the cases of discovery of VHE gamma-ray emission from FSRQs have occurred either during short flaring activities or longer high states. A notable counterex- ample is PKS 1510–089, from which persistent VHE gamma-ray emission was observed during a low flux level at HE (MAGIC Collaboration 2018). As the number of known VHE FSRQs is still very small, it is important to observe those objects to look for common patterns and differences in their emission and to investigate whether or not the same processes are responsible. Moreover, the observations need to be multiwavelength, thus contemporaneously covering the broad energy range of the spec- trum, which is often difficult to achieve due to fast variability. QSO B1420+326 is known to be strongly variable, in par- ticular in the HE range2. A few periods of HE flux enhance- ment from the source have been observed by Fermi-LAT so far, the most recent one starting in 2019 December (Ciprini & Cheung 2020). The high state continued, and based on the HE flux enhancement, the MAGIC telescopes performed follow-up 1 See http://tevcat.uchicago.edu/, in case of some sources the classification as FSRQ or BL Lac is however uncertain. 2 https://fermi.gsfc.nasa.gov/ssc/data/access/lat/msl_ lc/source/OQ_334 observations and discovered VHE gamma-ray emission from QSO B1420+326 in 2020 January (Mirzoyan 2020). The source is the fourth most distant VHE gamma-ray source known so far. Thanks to the duration of the enhanced state we were able to alert other observatories and trace the development of the flare in a broad range of wavelengths, from radio up to VHE gamma rays (e.g., Minev et al. 2020; D’Ammando et al. 2020; Fallah Ramazani et al. 2020). As ejection of new knots in the jets of FSRQs have often been reported contemporaneously to gamma- ray flaring activity (see, e.g., Aleksic´ et al. 2014; Lindfors 2015; Rani et al. 2018), we also organized very-long-baseline radio interferometry (VLBI) observations of the source following the high state. In this paper we report the broadband observations of QSO B1420+326 triggered by the 2019/2020 high state and other contemporaneous multiwavelength (MWL) data and their interpretation. In Sect. 2 we describe the instruments involved in the campaign, the data taken, and the analysis methods used. The results of the observations are reported in Sect. 3. In Sect. 4 we model the broadband emission of the source in different phases of the high state. Our results are summarized in Sect. 5. We use cosmological parameters H0 = 67.4 km s−1 Mpc−1, ΩΛ = 0.6847, and ΩM = 0.315 (Planck Collaboration VI 2020). 2. Observations and data analysis QSO B1420+326 was observed in 2020 January and February in a broad energy range by a number of instruments that either monitor the source, or had responded to a target of opportunity (ToO) announcement for a high state of the blazar, includ- ing also publicly released results. We report on the observa- tions performed in radio (VLBA, OVRO, Metsähovi, Badary RT32), near-infrared (NIR; CANICA), optical polarization and photometry (Perkins, LX-200, AZT-8+ST7), optical photome- try (Siena and Rozhen Observatories, ASAS-SN monitoring, MIRO/MFOSC-P, REM), optical spectroscopy (LDT), optical and UV from space satellites (Swift-UVOT and XMM-OM), X-rays (Swift-XRT and XMM-Newton), HE gamma-rays (Fermi- LAT), and VHE gamma-rays (MAGIC). To put the results in the context of earlier measurements we also use archival data retrieved via the Space Science Data Center3 from the catalogs: GB6 (Gregory et al. 1996), FIRST (White et al. 1997), NVSS (Condon et al. 1998), CLASS (Myers et al. 2003), JVASPOL (Jackson et al. 2007), WISE (Wright et al. 2010), GALEX (Bianchi et al. 2011), Planck (Planck Collaboration XXVIII 2014), 2RXS (Boller et al. 2016), and SDSS (Albareti et al. 2017). In the archival data sample we also include the lowest and the highest Swift-XRT states of the source (MJD 57631 and 58831), and the low state observed by Fermi-LAT (integrated from the mission start, MJD 54683 until MJD 57754). 2.1. MAGIC MAGIC is a system of two IACTs, each with a mirror dish diam- eter of 17 m (Aleksic´ et al. 2016a). The telescopes are located on La Palma in the Canary Islands (28.7◦ N, 17.9◦W), at a height of 2200 m above sea level. QSO B1420+326 has been observed by MAGIC on a few occasions following enhanced states at lower energies. We report 3 https://www.ssdc.asi.it/ A163, page 2 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 on observations between 2019 December 31 (MJD = 58848) and 2020 February 6 (MJD = 58885). The observations con- sisted of several triggers of the MAGIC Target of Opportu- nity program and are therefore not continuous, also due to bad weather periods. The data selection was based on the atmo- spheric transmission and rates of background events. The total amount of good-quality data is 14.0 h, out of which 2.9 h, taken between 2019 December 31 (MJD = 58848) and 2020 January 4 (MJD = 58852), was taken with a special low-energy analog trigger: SUM-Trigger-II (García et al. 2014). The data are ana- lyzed using MARS, the standard analysis package of MAGIC (Zanin et al. 2013; Aleksic´ et al. 2016b). The data selection is based mainly on the atmospheric transmission measured with LIDAR (Fruck & Gaug 2015) and on hadronic background rates. The effect of atmospheric absorption is corrected using LIDAR information. We analyzed the Sum-Trigger-II part of the dataset with dedicated low-energy analysis procedures including a spe- cial image cleaning (the so-called MaTaJu cleaning) with the cleaning thresholds tuned to the extragalactic field of view of QSO B1420+326 (Shayduk 2013; Ceribella et al. 2019). For the part of the dataset during which no signal is detected, we com- puted upper limits on the flux following Rolke et al. (2005) at 95% confidence level. 2.2. Fermi-LAT The Large Area Telescope (LAT) is a pair-conversion telescope launched on 2008 June 11 (MJD = 54628) as one of the two sci- entific instruments on board the Fermi Gamma-ray Space Tele- scope (Atwood et al. 2009). Its energy range extends down to ∼30 MeV and up to ∼300 GeV, with peak sensitivity at ∼1 GeV. In the 4FGL (Abdollahi et al. 2020), QSO B1420+326 is associ- ated to the gamma-ray source 4FGL J1422.3+3223, which has a >100 MeV flux of (9.1 ± 1.3) × 10−9 cm−2 s−1 and a power-law spectrum with photon index 2.38 ± 0.07, which was obtained from data between 2008 August 4 (MJD = 54682.65) and 2016 August 2 (MJD = 57602.24). QSO B1420+326 does not appear in any of the Fermi-LAT hard-spectrum source catalogs (see, e.g., the 3FHL catalog, Ajello et al. 2017), which is consistent with its relatively steep spectrum. We use the Python package Fermipy (Wood et al. 2017) to analyze the Fermi-LAT data. We use Pass8 event data (Atwood et al. 2013) and select photons of the SOURCE class in a square region of interest (ROI) of 10◦ × 10◦, centered at the position of the target source. We perform a binned analysis with ten bins per decade in energy and 0.1◦ binning in RA and Dec in the energy range 0.1−300 GeV, adopting the instrument response functions P8R3_SOURCE_V2. A correction for energy dispersion is included for all sources in the model except for the isotropic diffuse components. We apply a cut to include only the gamma- rays with zenith angle <90◦ to limit contamination from the Earth’s limb. We include in the model of the region all sources listed in the 4FGL within 15◦ from the ROI center, along with the Galactic (Acero et al. 2016) and isotropic diffuse emission models4. We first performed a likelihood fit using the full Fermi- LAT data set available at the time of the analysis, including events in the time range 2008 August 4 to 2020 February 11 4 gll_iem_v06.fits and iso_P8R3_SOURCE_V2_v1.txt, respec- tively. (MJD 54682.66–58890.00). We fit the full spectrum of the tar- get source, the diffuse models, and the normalization of catalog sources within 5◦ as free parameters. We also optimized the tar- get source localization, taking advantage of the ∼3.5 extra years of data with respect to 4FGL. The detection significance was estimated with the Test Statistic (TS, Mattox et al. 1996). We searched for new sources by performing a TS map of the ROI. We find no significant (TS> 25) new gamma-ray source in this analysis. Although mild excesses with TS∼ 10 are seen in the residuals, such fluctuations are to be expected when periods over long time ranges such as this are considered, and therefore we choose not to add these excesses as point sources in the ROI model. The model resulting from this initial fit was used as an input for the computation of the HE gamma-ray light curves. The light curves were calculated by performing a likelihood fit in each time bin. The fitting strategy is designed to adjust the number of free parameters to the photon statistics available in each bin. In the Fermi-LAT source catalogs, a source is consid- ered detected if TS is at least 25. In each light curve bin, we fit the full spectrum of the target source and the normalizations of the sources within 3◦ from the ROI center. If the target source has TS< 25, we progressively restrict the free parameters in the fit, reloading the average model at each step. First, we reduce the sources with free normalization to a radius of 1◦, then we freeze all sources except the target, and finally, if the target is not significantly detected, we only fit its normalization, leaving the spectral parameters fixed to the average value from the ini- tial model. We consider the target source to be detected in a given time bin if TS> 9, and the signal-to-noise ratio (i.e., flux divided by its error, or F/∆F) in that bin is larger than two. If this is not the case, we report a 95% confidence upper limit on the flux. We report light curves with fixed binning of 1 day and 30 days, and one with adaptive binning (Lott et al. 2012), setting a constant relative flux uncertainty of 15%. The latter method provides an estimate of the shortest timescale over which it is possible to obtain a statistically significant detection and a robust determination of the spectral parameters of the target. We calculated 0.1−300 GeV SEDs in the time intervals listed in Sect. 3 by performing a likelihood fit in several energy bins. The number of bins is optimized as a trade-off between energy resolution and photon statistics. We also performed an analy- sis including data from 2008 August 4 (MJD = 54682) to 2017 January 1 (MJD = 57754), in order to compute a quiescent- state Fermi-LAT SED to which the high-state ones could be compared. This time range was chosen based on the monthly light curve, which shows some early signs of flaring activity between the second half of 2017 and the first half of 2018 (see Appendix A.3). This time range is quite similar to the one corre- sponding to the 4FGL catalog, but provides a small increase in photon statistics due to the later end time. Finally, we calculated the probability that each single gamma ray recorded by the Fermi-LAT is associated with QSO B1420+326 using the gtsrcprob tool in order to investi- gate the highest energy photons in the 0.1−300 GeV energy range. 2.3. X-ray The Neil Gehrels Swift Observatory (Gehrels et al. 2004) car- ried out 26 observations of the source between 2018 January 20 (MJD = 58138) and 2020 February 10 (MJD = 58889), in particular on 14 individual days between 2020 January 2 and February 10 (MJD = 58850–58889). The pointed observations A163, page 3 of 19 A&A 647, A163 (2021) were performed with both the X-ray Telescope (XRT; Burrows et al. 2005, 0.2−10.0 keV) and the Ultraviolet/Optical Telescope (UVOT; Roming et al. 2005, 170−600 nm). The hard X-ray flux of this source is below the sensitivity of the BAT instrument for the short exposures of these observations; therefore, the data from this instrument are not used. Moreover, the source is not present in the Swift BAT 105-month catalog (Oh et al. 2018). All XRT observations were performed in photon counting (PC) mode. The XRT spectra are generated with the Swift-XRT data products generator tool at the UK Swift Science Data Cen- tre5 (for details see Evans et al. 2009). Ancillary response files are generated with xrtmkarf, and account for different extrac- tion regions, vignetting, and point-spread function corrections. We use the spectral redistribution matrices v014 in the Cali- bration data base maintained by HEASARC. Some of the spectra have very few photons, and so we are not able to use χ2 statis- tics. To maintain the homogeneity in our analysis, we grouped the obtained spectra using the task grppha to have at least one count per bin and we performed the fit with the Cash statis- tics (Cash 1979). The data collected during 2019 June 27 and 29 (MJD = 58661 and 58663) were summed in order to have enough statistics to obtain a good spectral fit. We fit the spectra with an absorbed power-law using the photoelectric absorption model tbabs (Wilms et al. 2000), with a neutral hydrogen col- umn density fixed to its Galactic value (1.08 × 1020 cm−2; Ben Bekhti et al. 2016). We also applied a log-parabola model to the XRT data, testing if this model is preferred over a single power- law model on a statistical basis by applying an F-test. The log- parabola model is preferred over a simple power-law model only on 2020 January 28 and at 95% confidence level. However, this could be due to the low statistics of the single XRT observations. XMM-Newton (Jansen et al. 2001) observed the source on 2020 January 24 between 04:44:07 and 11:49:07 (MJD 58872.3– 58872.5) for a total duration of 25 ks (observation ID 0850180201). The three EPIC cameras were operated in the large-window mode with medium filter. The data are reduced using the XMM-Newton Science Analysis System (SAS v16.0.0), applying standard event selection and filtering. Inspection of the background light curves shows that no strong flares were present during the observation, with good exposure times of 20, 24, and 24 ks for the pn, MOS1, and MOS2, respectively. For each of the detectors, the source spectrum is extracted from a circular region of radius 30 arcsec centered on the source, and the background spectrum from a nearby region of radius 30 arcsec on the same chip. All the spectra are binned to contain at least 20 counts per bin to allow for χ2-based spectral fitting. All spectral fits are per- formed over the 0.3−10 keV energy range using XSPEC v.12.10.1. The energies of spectral features are quoted in the source rest frame. All errors are given at the 90% confidence level. The data from the three EPIC cameras were initially fitted separately, but because good agreement was found (<5%) we proceeded to fit them together. Galactic absorption was included in all fits using the tbabs model. Three different models were applied: a simple power-law, a broken power-law, and a log-parabola model. The results of the fits are presented in Table 1. The F-test shows an improvement of the fit using both a broken power-law and a log-parabola model with respect to the simple power-law, with the log-parabola model providing the best fit. In order to check for the presence of intrinsic absorption, a neutral absorber at the redshift of the source was added to this 5 http://www.swift.ac.uk/user_objects Table 1. Summary of fits to the 0.3−10 keV XMM-Newton spectrum of the source. Model Parameter Value Power-law Γ 1.95 ± 0.01 Flux (0.3–10 keV) (5.2 ± 0.1) × 10−12 χ2/d.o.f. 1627.79/379 Broken power-law Γ1 2.38 ± 0.05 Ebreak 1.22+0.10−0.08 Γ2 1.58 ± 0.04 Flux (0.3–10 keV) (6.0 ± 0.1) × 10−12 χ2/d.o.f. 472.21/377 Log parabola α 2.12 ± 0.01 β −0.60 ± 0.03 Flux (0.3–10 keV) (6.2 ± 0.1) × 10−12 χ2/d.o.f. 432.42/378 Notes. Fits also include absorption fixed at the Galactic value. Flux and Ebreak (in the source rest frame) are given in units of erg cm−2 s−1 and keV, respectively. model, but it does not improve the fit quality and thus is not required. Moreover, no Fe line was detected in the spectrum. The 90% upper limit on the equivalent width (EW) of a narrow emis- sion line at 6.4 keV is EW< 10 eV. 2.4. Optical and UV from space-based telescopes During the Swift pointings, the UVOT instrument observed QSO B1420+326 in all its optical (v, b, and u) and UV (w1, m2, and w2) photometric bands (Poole et al. 2008; Breeveld et al. 2010). For each epoch, possible multiple exposures in the same filter are first summed with the task uvotimsum and then analyzed using the uvotsource task included in the HEAsoft package (v6.28) with the 20201026 release of the Swift/UVOTA CALDB. We checked if the observations are affected by small- scale sensitivity problems6. Source counts were extracted from a circular region of 5 arcsec radius centered on the source, while background counts were derived from a circular region of 20 arcsec radius in a nearby source-free region. The Optical Monitor (OM) on board the XMM-Newton satel- lite observed the source in the u, w1, m2, and w2 filters in imaging mode together with a fast readout window. The total exposure times of the imaging observations are: 3500 s (u), 3500 s (w1), 4400 s (m2) and 4400 s (w2). The data were pro- cessed using the SAS tasks omichain and omfchain. The UVOT and OM flux densities were corrected for extinction using the E(B−V) value of 0.010 from Schlafly & Finkbeiner (2011) and the extinction laws from Cardelli et al. (1989). 2.5. Optical from ground-based telescopes The REM telescope (Zerbi et al. 2001; Covino et al. 2004), a robotic telescope located at the ESO Cerro La Silla obser- vatory (Chile), performed optical photometric observations of QSO B1420+326 in the period 2020 January 24 to February 6 (MJD = 58872–58885). Observations were carried out with the Optical Slitless Spectrograph (ROSS2) obtaining three 240 s inte- gration images in the optical g′, r′, i′ bands. The REM data pre- sented here were obtained as ToO observations triggered by the 6 https://swift.gsfc.nasa.gov/analysis/uvot_digest/ sss_check.html A163, page 4 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 high gamma-ray flux observed by Fermi-LAT after the MAGIC detection. Instrumental magnitudes were obtained via aperture photometry and absolute calibration was performed by means of secondary standard stars in the field reported by the AAVSO Photometric All-Sky Survey (APASS) catalog7. Transformation between theu′ g′ r′ i′ z′ and UBVRI photometric systems was per- formed using the equations reported in Jester et al. (2005)8. Optical photometric (BVRI) and polarimetric (R band) observations were carried out at the 1.83 m Perkins telescope (Flagstaff, AZ, USA), 40 cm LX-200 telescope (St. Petersburg, Russia), and 70 cm AZT-8 telescope (Crimea) from 2020 Jan- uary 23 to April 21 (MJD 58871–58960). The photometric data were reduced using differential aperture photometry with respect to comparison stars in the quasar field9. The polarimetric obser- vations obtained at the Perkins telescope were performed and reduced in the same manner as described in Jorstad et al. (2010). The details of polarimetric observations carried out at the LX- 200 and AZT-8 telescopes can be found in Larionov et al. (2008). Photometric optical observations of QSO B1420+326 were carried out with the MFOSC-P instrument (Srivastava et al. 2018) mounted on the 1.2 m telescope of Mount Abu IR Obser- vatory (MIRO10) and used in imaging mode. The source was observed in B, V , R, and I bands (Johnson-Cousins filters) on 2020 February 2 and 6 (MJD = 58880 and 58885). MIRO is located at Gurushikhar peak in Mount Abu, India, at an altitude of 1680 m and is operated by the Physical Research Laboratory (PRL), Ahmedabad, India. The optical data from the National Astronomical Observa- tory (NAO) Rozhen, Bulgaria, were obtained between 2020 Jan- uary 24 and 26 (MJD = 58872–58874). We used the 2 m RCC telescope with Andor iKon-L CCD camera (2048× 2048 px, 13.5µm pixel−1) and the 50/70 cm Schmidt telescope with FLI PL-16803 CCD camera (4096× 4096 px, 9µm pixel−1). Addi- tional observations were carried out on 2020 January 31 to February 2 (MJD = 58879–58881) at the Student Astronomi- cal Observatory (SAO) Plana (Ovcharov et al. 2014) with the 35 cm Newton telescope and SBIG STL-11000M CCD Camera (4008× 2672 px, 9µm pixel−1). All cameras are equipped with standard photometric UBVRI Johnson-Cousins filters. The data were reduced (including bias subtraction, flat- fielding, and cosmic-ray correction) and analyzed using stan- dard photometry packages from IRAF11. For each image the PSF value was measured and aperture photometry was applied. Standard stars from the SDSS DR12 and VizieR catalogs were used for photometric calibration after applying transformation equations12. The Astronomical Observatory of the University of Siena observed QSO B1420+326 in its program devoted to optical photometry of blazars in support of MAGIC. The observatory runs a remotely operated 30 cm Maksutov-Cassegrain telescope installed on a Comec 10 micron GM2000-QCI equatorial mount. The focal plane hosts a Sbig STL-6303 camera equipped with a 3072 × 2048 pixels KAF-6303E sensor; Johnson-Cousins BVRI filters are available. Multiple 300 s images of QSO B1420+326 were acquired at each visit. After standard dark current subtrac- 7 https://www.aavso.org/apass 8 https://www.sdss.org/dr16/algorithms/ sdssUBVRITransform/ 9 See https://vo.astro.spbu.ru/vlar/opt_thumbs/b21420_ 1.png 10 https://www.prl.res.in/~miro/ 11 http://iraf.noao.edu/ 12 http://www.sdss3.org/dr8/algorithms/ sdssUBVRITransform.php#Lupton2005 tion and flat-fielding, images were averaged and aperture pho- tometry was performed on the average frame by means of the MaximDL software package. Reference and check stars in the field of view were selected from the APASS9 (Henden et al. 2016) catalog. The reference R magnitudes were derived from those reported in the same APASS9 catalog after conversion between the two different photometric systems, following a for- mula taken from Munari (2012). Additionally we use publicly available data in V-band and g-band of ASAS-SN (Shappee et al. 2014; Kochanek et al. 2017). Conversion of magnitudes to energy fluxes was done using zero points of Bessell et al. (1998). Correction for the Galac- tic extinction was applied using the E(B−V) value of 0.011 from Schlafly & Finkbeiner (2011) and the extinction laws from Cardelli et al. (1989). We performed observations of optical spectra of the quasar QSO B1420+326 using the 4.3 m Lowell Discovery Telescope (LDT; Lowell Obs., Flagstaff, AZ) equipped with the DeVeny spectrograph and the Large Monolithic Imager (LMI), in response to the detection of the source at VHE by MAGIC on 2020 January 21 (MJD = 58869). We employed a grating setting of 300 grooves per millimeter, which provides spec- tra from 3300 Å to 7500 Å with a dispersion of 2.17 Å per pixel, a blaze wavelength of 5000 Å, and a slit width of 2.5′′. The spectroscopic observations were performed on 2020 Jan- uary 28 (MJD 58876.578), and February 8 (MJD 58887.497) and 25 (MJD 58904.368). During this month the brightness of the quasar fell from 14.6 mag to 16.25 mag in R band. Each observa- tion of the quasar consisted of three exposures of 600 s (900 s on February 25). Observations of a comparison star HD 126944 of A type, located ∼86′ from the quasar, were performed before and after target observations, with two 30 s exposures for each obser- vation. Bias and flat-field images were obtained regularly. The LDT allows a switch between different instruments in 2−3 min. Therefore, photometry of the quasar using the LMI in V and R filters was performed just before or after spectral observa- tions to calibrate the spectra. The observations are reduced using programs written in IDL (v.8.6.1) that implement the technique described in Vacca et al. (2002) developed for reduction of spec- tra obtained with SpeX at the NASA Infrared Telescope Facility on Mauna Kea. 2.6. Near-infrared The NIR observations were carried out with the camera CANICA (Carrasco et al. 2017) and the Guillermo Haro 2.1 m telescope (OAGH) located at Cananea Sonora, Mexico. The camera is based on a Hawaii detector of 1024 by 1024 pixels with a plate scale of 0.32 arcsec per pixel. The data are part of the monitoring program “NIR photometry of AGNs with Gamma Ray emission detected by Fermi-LAT”. Relative pho- tometry is obtained with respect to the 2MASS point sources included in the field of view (5.5 arcmin). Absolute fluxes are obtained adopting the zero point values of 2MASS derived by Cohen et al. (2003). The host galaxy is not detected in IR by the 2MASS survey in any of the three bands, resulting in an upper limit of H ∼ 17.7 mag, much weaker than 13−11 mag observed by CANICA during the investigated period. Therefore, the effect of the host galaxy is negligible. 2.7. Radio The 37 GHz observations were made with the 13.7 m diameter Metsähovi radio telescope. The detection limit of the telescope at 37 GHz is on the order of 0.2 Jy under optimal conditions. A163, page 5 of 19 A&A 647, A163 (2021) Data points with S/N < 4 are handled as nondetections. The flux density scale was set by observations of DR 21. Sources NGC 7027, 3C 274, and 3C 84 were used as secondary calibra- tors. A detailed description of the data reduction and analysis is given in Teräsranta et al. (1998). The error estimate in the flux density includes the contribution from the measurement rms and the uncertainty of the absolute calibration. We also use pub- licly available 15 GHz OVRO data (Richards et al. 2011) and 8.63 GHz data from Badary RT32 reported in Kharinov (2020). We requested Director’s Discretionary Time (DDT) with the Very Long Baseline Array (VLBA) following the gamma- ray activity and VHE detection of the quasar QSO B1420+326 and were granted six epochs of observations of the source sep- arated by approximately 1 month intervals (ID BD227), with 8 h per epoch. Thus far, we have obtained three epochs of observations under the project, on 2020 March 8, May 10, and June 6 (MJD = 58916, 58979, 59006). The observations were performed with all ten antennas in continuum, dual cir- cular polarization mode at 43 GHz using four intermediate frequency bands (IFs), each of 64 MHz width. The data were cor- related at the National Radio Astronomy Observatory (NRAO, Soccoro, NM) using the VLBA DiFX software correlator. Five sources were observed at each epoch (QSO B1420+326 3C 279, 3C 345, PKS 1055+18, and B2 1308+326), with 60% of the time devoted to the main target, QSO B1420+326. The sources 3C 279, 3C 345, and PKS 1055+018 are used for fringe finding during the correlation. The data were reduced using the Astro- nomical Image Processing System software (AIPS, van Moorsel et al. 1996) and Difmap package (Shepherd 1997) in the same manner as described in Jorstad et al. (2017), except without aver- aging the final calibrated data over IFs. We used the sources observed along with QSO B1420+326 to perform absolute cal- ibration of the electric vector position angle (EVPA), as these sources are monitored in the VLBA-BU-BLAZAR program13 meaning that their polarization properties at 43 GHz are known. 3. Results In Fig. 1 we present the MWL light curve summarizing the evolution of the flare. Based on the VHE state of the source we define three periods selected for the further spectral analy- sis: A: 2019 December 29 to 2020 January 5 (MJD = 58846.5– 58853.5): without VHE gamma-ray detection; C: 2020 January 20 to 22 (MJD = 58868.3–58870.3): VHE gamma-ray flare; and D: 2020 January 25 to February 1 (MJD = 58873.5–58880.5): detection over longer timescale. In addition we define the fourth period: B: 2020 January 19 to 20 (MJD = 58867–58868), which does not have simultaneous MAGIC data, but contains the peak of the optical and IR flare as well as one of the local peaks of HE emission. The four periods (referred to throughout the paper as periods A–D) are summarized in Table 2. In each period we construct a broadband SED (see Fig. 2). In the case of gamma-ray data, all observations performed within a given time window are summed. X-ray data available from different observations are stacked for the period D, while in the other periods a single Swift-XRT observation is available and used. On the other hand, the low-energy data (radio up to UV) have mostly lower uncertainties and hence are more sensitive to variability. Therefore, if more than one measurement was taken at a given time period we average all the measurements and take the standard deviation of the measurements as the measure of 13 www.bu.edu/blazars/VLBAproject.html its uncertainty. A similar approach for constructing a broadband SED was applied by for example MAGIC Collaboration (2018). 3.1. Very high-energy gamma-ray emission The first detection of the VHE gamma-ray emission from QSO B1420+326 was achieved on 2020 January 20 (MJD = 58868). A highly significant detection of 14.3σ was obtained in 1.6 h of effective time (see Fig. 3). In the subsequent period of 2020 January 26 to February 1 (MJD = 58874–58880) further hints of signal were obtained, with the highest signif- icance of the excess (6.6σ) on the night of 2020 January 31 (MJD = 58879), which has also the longest exposure of 2.5 h. During the flare (period C), the flux observed by MAGIC above 100 GeV reached (7.8 ± 1.3stat) × 10−11 cm−2 s−1. The observed spectrum in this period can be described by a power law: dN/dE = (2.49 ± 0.31stat) × 10−9(E/100 GeV)−4.22±0.24stat [TeV−1 cm−2 s−1]. Correcting for the EBL absorption according to Domínguez et al. (2011), the unattenuated spectrum can be described as dN/dE eτEBL(E) = (4.04±0.54stat)×10−9(E/100 GeV)−3.57±0.29stat [TeV−1 cm−2 s−1]. After the flare (period D), significant gamma-ray emis- sion was detected again with 5.6σ, but at about half the flare level: FD(>100 GeV) = (3.9 ± 0.6stat) × 10−11 cm−2 s−1. The observed spectrum in this period can be described as: dN/dE = (0.91±0.13stat)×10−9(E/100 GeV)−3.90±0.25stat [TeV−1 cm−2 s−1]. Correcting for the EBL absorption according to Domínguez et al. (2011), the spectrum can be described as dN/dE eτEBL(E) = (1.64±0.22stat)×10−9(E/100 GeV)−2.87±0.36stat [TeV−1 cm−2 s−1]. Despite enhancement of the VHE gamma-ray flux by a factor of two, the spectral indices during and after the flare are consistent within 1.5σ. However, it should be noted that in particular in period D, the uncertainties of the photon index are large. Before the flare (period A), possibly due to shorter observa- tions under less favorable zenith angle, no significant emission was detected and only a 95% C.L. limit of <4.1× 10−11 cm−2 s−1 can be placed for the flux above 100 GeV. The limit is compa- rable to the emission detected from the source in period D. The SED of QSO B1420+326 observed by MAGIC in different peri- ods is shown in Fig. 4. 3.2. High-energy gamma rays The first detection of a HE outburst from B2 1420+326 was reported in 2018 December (Ciprini 2018), with a flux increase by more than two orders of magnitude with respect to the average 4FGL value and significant spectral hardening. A similar spec- tral hardening was reported in 2019 July (Angioni 2019), where the first evidence of >10 GeV photons was provided, and again in 2020 January (Ciprini & Cheung 2020). The daily Fermi-LAT light curve is shown in Fig. 1, includ- ing the flux (second panel) and photon index (third panel). Both the flux and the photon index are significantly variable in this time interval, based on a simple χ2 test. The Fermi-LAT recorded a peak daily flux (E > 100 MeV) of (2.6 ± 0.2) × 10−6 cm−2 s−1 on 2020 January 19 (MJD = 58867), corresponding to about 300 times the average value reported in the 4FGL catalog. The highest-energy photon observed by the Fermi-LAT was recorded two days prior (2020 January 17, MJD = 58865), with an energy of ∼150 GeV, providing the first indication of VHE emission from QSO B1420+32614. Accordingly, the Fermi-LAT recorded the hardest daily spectrum on the same date, with a photon index 14 The probability of this photon being associated with the target source is >99.9%, as obtained from the gtsrcprob tool. A163, page 6 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0.5− 0 0.5 1 1.5 ] - 1 s - 2 cm - 10 F [ 10 MAGIC >100 GeV 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 10 20 30]-1 s - 2 cm - 7 F [ 10 Fermi-LAT F >0.1GeV 58845 58850 58855 58860 58865 58870 58875 58880 58885 588901.5 2 2.5 Ph ot on in de x Fermi-LAT index 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 2 4 6 8 ] - 1 s - 2 er g c m - 12 F[1 0 0.3-10 keV Swift-XRT XMM 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 1.5 2in de x 0.3-10 keV index Swift-XRT XMM 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 2 4 6 8 F [ mJ y] M2 & B-bands UVOT M2 XMM-OM M2 MIRO B UVOT B Perkins B 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 5 10 15 F [ mJ y] V-band ASAS-SN V&g Rozhen V MIRO VREM V UVOT V Perkins V 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 5 10 F [ mJ y] R-band Siena R Rozhen R MIRO R REM R Perkins++ 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 5 10 15 20 Po lar iza tio n [ %] R-band polarization 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 150− 100− 50− 0 EV PA [d eg] R-band EVPA 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 5 10 F [ mJ y] NIR (I-band) MIRO I REM I Perkins I 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 0 20 40 60 F [ mJ y] NIR (Ks-band) 58845 58850 58855 58860 58865 58870 58875 58880 58885 58890 Time [MJD] 0 1 2F [ Jy] Radio Metsahovi 37GHz OVRO 15 GHz RT32 8.63GHz Fig. 1. MWL light curve of QSO B1420+326 between 2019 December 28 (MJD = 58845) and 2020 February 11 (MJD = 58890); see titles and legends of individual panels. Optical and UV observations are corrected for the Galactic attenuation. The points in red are contemporaneous (±12 h) with MAGIC observations. The shaded regions show the time ranges of the four considered flare evolution periods (see Table 2). Flux upper limits in the first two panels are shown with downward triangles. A163, page 7 of 19 A&A 647, A163 (2021) Table 2. Periods selected for the SED modeling (the colors of the period tag correspond to colors used in the figures). Period MJD Comment A 58846.5–58853.5 Pre-flare B 58867–58868 Optical flare C 58868.3–58870.3 VHE flare D 58873.5–58880.5 Post-flare 1010 1210 1410 1610 1810 2010 2210 2410 2610 [Hz]ν 14−10 13−10 12−10 11−10 10−10 9−10 ] - 1 s - 2 dN /d E [er g c m 2 E Archival MJD 58846.5-53.5 (A) MJD 58867-8 (B) MJD 58868.3-70.3 (C) MJD 58872 MJD 58873.5-80.5 (D) Fig. 2. MWL SED of QSO B1420+326 in 2020 January and February. Points follow the colors of the shaded regions in Fig. 1. In addition we show XMM points taken on 2020 January 24 (MJD = 58872) in black. MAGIC points are not corrected for the EBL absorption. Gray points show archival data (most from the ASI Space Science Data Center, but also including low-state Fermi-LAT spectrum and lowest and highest X-ray spectrum from Swift-XRT). of 1.72± 0.08, a flux of (1.2± 0.2)× 10−6 cm−2 s−1 , and a TS of 496. We also note that the photon index was consistently harder than the 4FGL catalog value (2.38 ± 0.07) during most of the time range shown in Fig. 1. The adaptively binned light curve is shown in Fig. 5. The shortest adaptive bin is centered on 2020 January 19, 00:17:45 (MJD = 58867.012), and has a width of ∼6 h. The highest flux is recorded in the same bin, reaching a value of (3.6 ± 0.5) × 10−6 cm−2 s−1, i.e., 400 times higher than the 4FGL value. We note that a test for spectral curvature was performed in all time bins of all light curves, and a power-law spectrum was found to be the best representation in all time intervals. Additionally, we performed a likelihood fit over the full time interval included in Fig. 1 to characterize the average source properties in the HE band in this flaring state. For this time interval, the LogParabola model is preferred (TScurv = 3615) with respect to a simple power law to describe the gamma-ray spectrum of the source. It yields a photon flux (1.09 ± 0.02) × 10−6 cm−2 s−1, and spectral parameters α = 1.97 ± 0.02, and β = 0.07 ± 0.0116. We also verified that there are no new signif- 15 TScurv = 2(ln LLP − ln LPL) is used to check whether or not a statisti- cally significant curvature is detected using a LogParabola model com- pared with the PowerLaw model, where ln L indicates the logarithm of the likelihood for each model. A source is considered to have a statisti- cally significant curvature if TScurv > 25. 16 The functional form of the LogParabola spectral model is dN/dE = N0(E/Eb)−[α+β log(E/Eb)], where N0 is the normalization, Eb is the refer- ence energy at which N0 is measured, α is the slope, and β is the curva- ture parameter. ]2 [ deg2θ 0 0.1 0.2 0.3 0.4 e ve n ts N 0 100 200 300 400 500 Time = 1.63 h 8.3± = 358.3 off = 692; NonN 27.6± = 333.7 exN σSignificance (Li&Ma) = 14.0 Fig. 3. Distribution of the squared angular distance between the source nominal and event reconstructed directions (points) and corresponding background estimation (shaded region) for the night of 2020 January 20 (MJD = 58868) of MAGIC observations. Vertical line shows the cut value below which the event statistics are given in the inset. E [GeV] 40 50 60 70 80 90 100 200 300 ] - 1 s - 2 dN /d E [Te V cm 2 E 12−10 11−10 10−10 A (obs.) C (obs.) C (deabs.) D (obs.) D (deabs.) Fig. 4. SED of QSO B1420+326 observed by MAGIC in periods A, C, and D (see legend): observed (filled circles) and corrected for EBL absorption (empty circles). icant point sources in addition to the initial 4FGL model during this period. Finally, as mentioned in Sect. 2, we performed separate like- lihood fits corresponding to the periods used to build time- resolved SEDs. The results of these fits are summarized in Table 3, together with the fits for the quiescent state. 3.3. X-ray There is no strong variability of the X-ray flux in the inves- tigated period with an increasing trend from 2020 January 05 (MJD = 58853) to 2020 January 25 (MJD = 58873), where a fac- tor of two higher peak flux is observed. However, there is clear variability of the X-ray spectral index from hard values before the flare to much softer values at the time around the optical and gamma-ray flares, and then back to hard values. This indicates a shift of the synchrotron peak of the source, and connected with it the shift of the crossing point between the synchrotron and the IC component. This is clearly visible in the X-ray spectra during the flare and during XMM-Newton observations (see Fig. 2). Results A163, page 8 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 58850 58855 58860 58865 58870 58875 58880 ] - 1 s - 2 cm - 7 F [10 10 20 30 40 F >0.1GeV 58850 58855 58860 58865 58870 58875 58880 Ph ot on in de x 1.5 2 2.5 Photon index Time [MJD]58850 58855 58860 58865 58870 58875 58880 En er gy [G eV ] 10 100 Energies Fig. 5. Fermi-LAT adaptively binned light curve of QSO B1420+326 in January 2020. The vertical shaded areas follow the color coding of Fig. 1. Top panel: 0.1−300 GeV photon flux. Middle panel: power-law photon index (the horizontal shaded area represents the average value from the 4FGL with its uncertainty). Bottom panel: E > 10 GeV photons associated with QSO B1420+326 with probability >80%. Table 3. Results of Fermi-LAT analysis on the different activity states of QSO B1420+326. State Flux (a) Photon index TS Quiescent 0.86 ± 0.13 2.40 ± 0.08 218 A 115 ± 6 2.04 ± 0.04 1942 B 260 ± 20 1.88 ± 0.07 688 C 160 ± 20 1.88 ± 0.06 927 D 121 ± 7 1.93 ± 0.04 2280 Notes. The time periods are defined in Table 2. (a)Total flux in the energy range 0.1−300 GeV in units of 10−8 cm−2 s−1. of the spectral fits to individual days of Swift-XRT observations are given in Appendix A.2. 3.4. Optical Compared to the historical measurements, the optical emission is ∼1.5 orders of magnitude higher throughout the investigated period. Moreover, during that period a strong optical flare is observed on 2020 January 19 (MJD = 58867) with a variability timescale of the order of a few days. V-band observations per- formed during one of the Fermi-LAT peaks show an increase by nearly an order of magnitude with respect to observations at the beginning of Period A. Similar variability pattern is seen also in IR and UV ranges. However, the spectral shape in the IR-UV range varies during the investigated period, with the spectrum becoming bluer (harder) during the optical flare (see Fig. 2). This is consistent with the X-ray behavior of the source that also sug- gests a shift of the synchrotron peak position. Recently other occurrences of comparable flaring activity in the optical were observed (e.g., in July 2019, Marchini et al. 2019), reporting a R = 13.7 magnitude even slightly brighter than the brightest R point in Fig. 1. Fig. 6. Optical spectra of QSO B1420+326 obtained with the LDT; the spectra are in the frame of the observer. Also, optical polarization shows interesting behavior with a dip of the polarization percentage at a few percent level and a concurrent rotation by ∼150◦. Similar EVPA rotation dur- ing low-level polarization was also seen contemporaneous with VHE emission in, for example, PKS 1510–089 (Aleksic´ et al. 2014). High variability of EVPA (down to a timescale of 3 h) was also seen in PKS 0736+017, another FSRQ, during the period of VHE gamma-ray detection (H.E.S.S. Collaboration 2020). 3.5. Optical spectroscopy Figure 6 presents the optical spectra of QSO B1420+326 obtained with LDT at three epochs of different activity states (taken mostly after the gamma-ray flaring activity). The spec- tra show the presence of MgII emission line at λ = 4706 Å (rest wavelength of 2798 Å) and a bump between 3800 Å and 4100 Å (rest frame ∼2260−2450 Å). The latter appears to be part of an FeII emission complex whose strongest UV lines fall in the 2200−2600 Å range (e.g., Baldwin et al. 2004). The spectra also include three prominent absorption lines at λ ∼ 4001Å, 4337 Å, and 4477 Å that intensify as the quasar brightens. The absorp- tion lines are most likely intervening MgII systems at redshifts of z ≈ 0.43, 0.55, and 0.60, respectively. We analyzed the characteristics of the MgII emission line and FeII bump as a function of continuum brightness. Figure 7 shows an approximate Gaussian fit to the MgII line, while Fig. 8 plots similar modeling of the FeII bump at all three epochs. The parameters of the Gaussian and velocities of clouds, as well as the flux of the lines, are given in Table 4, although we are unable to estimate the velocity of gas producing FeII lines, because the bump consists of more than 100 lines. There is a significant dif- ference between the MgII line and FeII bump behavior: (i) The flux of the MgII line remains constant within 1σ uncertainty independent of the continuum brightness, while the FeII bump increases in flux with the continuum level (see Fig. 9); (ii) the central wavelength of the MgII line fits does not show a shift with respect to the rest wavelength, while λ◦ of the FeII bump shifts toward the blue side as the time after the VHE event passes. Unfortunately, it is not possible to distinguish whether the shift is due to a relative change of the brightness of lines that form the FeII bump, or due to gas motion toward the observer; and (iii) the FWHM of the FeII bump is very stable despite the correla- tion of its flux with the continuum level, while the velocity of gas where the MgII line originates increases with time after the A163, page 9 of 19 A&A 647, A163 (2021) 2740 2760 2780 2800 2820 2840 Wavelength, Α 0.0 0.5 1.0 1.5 F λ , e rg s- 1 cm -2 Α -1 x 10 -1 5 MgII 58876 58887 58904 Fig. 7. Observed (solid) and modeled (dotted) MgII line of QSO B1420+326 in the rest frame at three epochs. The lines on 2020 January 28 (MJD = 58876), February 8 (MJD = 58887), and February 25 (MJD = 58904) are shifted by values of 0.8, 0.3, and 0.1, respec- tively, for clarity. VHE event. We also note a significant change of the EW of the MgII line with the continuum, with EW decreasing as the contin- uum rises. This questions the identification of QSO B1420+326 as a FSRQ, but Table 4 shows that at the lower levels of activity EW> 5 Å for the MgII line17. The increase in the flux of the FeII bump with the continuum and a possible motion of gas producing FeII lines toward the observer are quite interesting. The short time-lag between the continuum and line variability suggests that the FeII emission- line region is much smaller than the region producing the MgII line and/or lies close to the line of sight. It is possible that the nonthermal jet is interacting with an FeII-emitting cloud, while the MgII emission is excited by the underlying thermal accretion disk continuum, which varies on a much longer timescale. We adopt an approach suggested by Ghisellini et al. (2014) (see also references therein), who used an estimate of the accre- tion disk (AD) luminosity based on the luminosity of the BLR, LAD ≈ 10 LBLR. The known flux density of the MgII line, com- bined with the BLR template constructed by Vanden Berk et al. (2001) for a composite emission spectrum of a quasar using SDSS spectra, allow us to estimate the total luminosity of the BLR in QSO B1420+326 LBLR = (1.8± 0.2)× 1045 erg s−1. This, for a luminosity distance of 4256.4 Mpc, translates to LAD ≈ 2 × 1046 erg s−1, with an uncertainty of a factor of about two, as suggested by Ghisellini et al. (2014). Interestingly, the obtained luminosity of the accretion disk is rather high, in the high-end part of values shown for other sources (Ghisellini et al. 2014). 3.6. Radio Moderate variability is seen in radio observations (see also Appendix A.3). OVRO data during the investigated period show a gradual increase in the flux. No monotonous behavior is seen in flux at 37 GHz by Metsähovi, but a constant fit can be rejected at χ2/Nd.o.f. = 51/21. The amplitude of the variability is ∼10%. The total and polarized intensity VLBA maps of QSO B1420+326 are presented in Fig. 10. The images exhibit the bright VLBI core located at the northwestern end of the jet and a weak extended jet at a position angle of ∼130◦. 17 EW of 5 Å is the classical threshold between BL Lac and FSRQ objects, see, e.g., Sambruna et al. (1996). Fig. 8. Observed (black) and modeled (blue) FeII bump of QSO B1420+326 at three epochs in the rest frame; red lines represent Gaussian fits to absorption lines, and gray dashed lines indicate the con- tinuum level. The total and polarized intensity images are modeled by cir- cular components with Gaussian brightness distributions. Two bright features are apparent at each epoch: the core A0, and a knot K20. We assume that the core is a stationary feature of the jet and calculate parameters of K20 with respect to the core. Parameters of the modeling are given in Table 5. According to Fig. 10 and Table 5, knot K20 is the most polarized feature of the jet. Figure 11 shows the separa- tion of K20 from the core as a function of time. Accord- ing to a linear approximation, K20 moves with a proper motion µ= 0.30± 0.04 mas yr−1, which translates into superlu- minal motion down the jet with apparent speed βapp = 12.0± 1.7 in units of c. Such motion suggests that the ejection18 of K20 through the VLBI core occurred on MJD 58831± 21 (2019 December 13). Figure 11 also shows the light curve of K20, which reveals a very fast decrease of the knot flux density by a factor of about four in three months, which corresponds to the timescale of variability of 0.16± 0.03 yr, as defined according 18 Passage of the centroid of K20 through the centroid of A0. A163, page 10 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 Table 4. Parameters of lines. MJD Line λ◦ Amp FWHM v Flux S cont EW (1) (2) (3) (4) (5) (6) (7) (8) (9) 58876.578 MgII 2795± 2 0.37± 0.04 18.4± 1.3 1982± 140 13± 3 7.1± 0.2 2± 1 58887.497 MgII 2798± 3 0.36± 0.06 26.0± 2.6 2787± 278 18± 4 3.3± 0.2 6± 3 58904.368 MgII 2797± 2 0.31± 0.02 28.6± 0.7 3074± 75 16± 2 1.5± 0.2 11± 2 58876.578 FeII 2390± 5 0.49± 0.05 91± 3 – 84± 9 9.2± 0.5 – 58887.497 FeII 2357± 5 0.23± 0.06 91± 2 – 58± 5 4.5± 0.2 – 58904.368 FeII 2348± 5 0.15± 0.03 91± 1 – 27± 7 1.8± 0.3 – Notes. Columns as follows: (1) epoch of observations; (2) emission line; (3) central wavelength of Gaussian fit in Å; (4) amplitude of Gaussian fit in units of 10−15 erg cm−2 s−1 Å−1; (5) full width of Gaussian at half-maximum (FWHM) in Å; (6) velocity of gas in km s−1; (7) flux of line in 10−15 erg cm−2 s−1; (8) flux density of the continuum at the peak of line in 10−15 erg cm−2 s−1 Å−1; (9) equivalent width in Å. Fig. 9. Dependence between flux of the FeII bump and flux density of the continuum at different epochs (points show the measurements, and the line is a linear fit). to Burbidge et al. (1974). Applying the formalism proposed by Jorstad et al. (2005), this timescale of variability and the aver- age size of K20 (0.084± 0.015 mas according to Table 5) give a value of the Doppler factor of K20 δ= 33± 9. The latter, along with the apparent speed of the knot, allow us to estimate the Lorentz factor, Γb = 19± 9, and viewing angle, Θ◦ = 1.1◦ ± 0.6◦, of K20. Using the proper motion and size of K20, we can esti- mate that during the time of VHE emission the upstream edge of the knot was passing through the centroid of the core. The direc- tion of polarization in K20 is close to the jet direction, which suggests that K20 is most likely a moving shock whose sur- face is oriented transverse to the jet axis. This is supported by a higher degree of polarization of the knot compared with the core, which implies ordering of the magnetic field as expected by such a shock. 4. Spectral energy distribution modeling FSRQ gamma-ray emission is usually explained in the frame- work of an external Compton (EC) model. Moreover, the opti- cal spectroscopy measurements performed during the dimming phase of the emission (see Fig. 8) show a significant increase in flux of FeII + FeIII lines. As the source emits up to VHE range, the most natural target for EC process is the dust torus (DT) radiation field (see, e.g., Costamante et al. 2018; van den Berg et al. 2019). DT radiation field, contrary to BLR, would not strongly absorb the sub-TeV gamma-rays, and thus is the com- mon assumption for the dominating radiation field in the model- ing of FSRQs detected at VHE gamma-rays. On the other hand, the large increase in the optical flux during the gamma-ray flare can also provide a significant target for the synchrotron-self- Compton (SSC) process. We therefore investigated a scenario in which both SSC and EC processes are possible. Intriguingly, compared to the low-state spectrum, the inves- tigated synchrotron and gamma-ray peaks are shifted towards higher energies. Similar behavior has been observed in a few other FSRQs during enhanced states, e.g., PMN J2345– 1555 (Ghisellini et al. 2013), 4C +49.22 (Cutini et al. 2014), PKS 1441+25 (Ahnen et al. 2015; Abeysekara et al. 2015), PKS 1510–089 (D’Ammando et al. 2011), and PKS 0346–27 (Angioni et al. 2019). However, for most of these cases, the peak frequency did not reach beyond 1014 Hz. Therefore, the behavior observed in QSO B1420+326, in particular the SED peaking at a few times 1014 Hz in the optical range during period B, while not being unique, is still rarely observed in FSRQs. The peak position traces the electron energy distribution (EED), but it is also dependent on other physical parameters of the source (e.g., on the beaming). We model the source in a framework of a simple one-zone model in which a spherical emission region is homogeneously and isotropically filled with an electron distribution and mag- netic field. We consider a broken power-law energy distribu- tion of electrons, i.e., dNe/dγ ∝ γ−p1 for γmin < γ < γb and dNe/dγ ∝ γ−p2 for γb < γ < γmax. The electrons in the blob are also exposed to an additional, directional radiation field coming from the DT. The model assumes ring geometry of the DT, and thus depends on the distance of the emission region from the black hole. The SED model was generated with agnpy19 (Nigro et al. 2020), which implements the synchrotron and Compton processes following the prescriptions described in Dermer & Menon (2009), Finke (2016). We fix the Lorentz and Doppler factors of the blob to Γ = 40 and δ = 40, respectively. We note that those values are somewhat larger (in particular the Γ fac- tor of the blob) than the jet parameters estimated using VLBI observations (see Sect. 3.6). However, the VLBI measurements are performed a few months after the flaring event and trace the later evolution of the blob, and therefore some change of parameters of the jet might have happened in the meantime (in particular deceleration). The size of the blob is limited by the variability condition. The values that we use in the modeling, rb = 3−6 × 1016 cm, correspond to the light-crossing time of 12−24 h, of the order of the timescale of the observed variability. 19 https://github.com/cosimoNigro/agnpy/ A163, page 11 of 19 A&A 647, A163 (2021) Fig. 10. Total (contours) and polarized (color scale) intensity maps of QSO B1420+326 obtained with the VLBA at 43 GHz; the peak total intensity is 1.14 Jy beam−1; the beam is displayed in the bottom left cor- ner; the contours are 0.25, 0.5, 1, . . . 64, and 90% of the peak total inten- sity. Linear segments within images show direction of polarization, the black vertical line indicates the position of the core, A0, and red circles mark positions of knot K20; 1 mas corresponds to 7.29 pc. We note that the polarized flux intensities on May 10 and June 6 are multiplied by factors of three and five, respectively, to reveal locations of peaks of polarized flux intensity, which are 15 mJy beam−1 and 9.5 mJy beam−1, respectively, for the May 10 (MJD = 58979) and June 6 (MJD = 59006) images. We assume that the emission region is located at a distance of d = 2.5×1018 cm, that is, of the order of ∼Γrb. We use the accre- tion disk luminosity Ldisk = 2 × 1046 erg s−1 (see Sect. 3.5) to estimate the size of the BLR and DT following the scaling rela- tions of Ghisellini & Tavecchio (2009). We note that while the optical spectra used in this estimation are not fully simultaneous with the broadband emission data used for the modeling, the size of the DT makes the emission quasi-stable at the timescales of years. The dust torus is simulated as a thin ring with a radius of RDT = 1.1 × 1019 cm (3.6 pc). As the estimated size of the BLR is RBLR = 4.4 × 1017 cm (0.14 pc), the emission region is not affected by the BLR but it is deep in the DT radiation field. We assume that 0.6 of the disk luminosity is reprocessed in the DT radiation. In order to obtain EED in a self-consistent way, we introduce the acceleration parameter ξ, defined such that acceleration gain of electrons is (dE/dt)acc = ξcE/RL, where c is the speed of light and RL is the Larmor radius. The maximum energy of the elec- trons is obtained from comparing the acceleration energy gain with energy losses due to IC (in Thomson regime) cooling: γmax = √ 3ξeB/4σTu′ph, (1) where B is the co-moving magnetic field in the blob, e is ele- mentary charge, σT is the Thomson cross-section, and u′ph is the co-moving energy density of the dominating radiation field. In the case of the parameter sets used in the modeling, the dominat- ing radiation field is originating from the DT. Nevertheless, we explicitly also check a possible limit from the SSC process. The maximum electron energies are also tested against the dynamical timescale by comparing the acceleration timescale with the bal- listic timescale20, Tbal ' rb/c, of the blob crossing its size, and against synchrotron energy losses. However, neither of the two processes is dominant in the cases investigated here. As during Tbal the blob crosses its size, it might escape the region in the jet (e.g., a stationary shock) where efficient acceleration occurs. We note that Tbal also determines the timescale of the adiabatic losses of the electron (following the assumption that the blob distance is d = rbΓ). We also introduce a canonical cooling break (steepening of EED by 1) at the energy where the dominating cooling timescale is equal to the ballistic timescale, Tbal: γb = 3mec2/4σTu′phrb, (2) where me is the electron mass. In order not to overshoot the X- ray flux with IC photons and still be able to explain the gamma- ray emission during the periods B, C, and D, it is necessary to assume that the EED starts at γmin > 1. We apply values of γmin = 10−15 in the modeling, while for period A we use γmin = 1. We then tune the magnetic field B and the energy den- sity of the electrons u′e to reproduce the levels of synchrotron and IC emission. The index of EED before the break is selected to roughly reproduce the spectral shape in IR-UV (except of period D) and gamma-ray bands. The resulting value p1 = 1.7−2, is close to the canonical nonrelativistic value of 2 (see, e.g., Drury 1983). The location of the valley between the peaks is most sen- sitive to the maximum γ factor of the electrons and the onset of the IC peak. The maximum energies of the electrons are tuned by the acceleration parameter ξ (see Eq. (1)), with an additional constraint from the VHE gamma-ray spectrum. The depth of the valley is additionally modified by the SSC component. By tuning the compactness of the blob (i.e., varying the radius of the blob within a factor of two while simultaneously keeping the same total power in electrons that fixes the synchrotron and EC emis- sion at roughly the same level), which affects the SSC emission, we fit the depth of the dip in periods B, C, and D. As in all the studied energy bands the emission during the whole investigated period was significantly larger than the historical measurements, we neglect the possible low-state emission of the source in the 20 Also called light crossing time scale. A163, page 12 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 Table 5. Parameters of core A0 and knot K20. MJD Knot S R Θ a P EVPA (1) (2) (3) (4) (5) (6) (7) (8) 58916 A0 0.90± 0.06 0.0 – 0.029± 0.007 3.5± 0.3 110± 6 58979 A0 0.52± 0.06 0.0 – 0.023± 0.007 1.5± 0.5 −20± 8 59006 A0 0.44± 0.05 0.0 – 0.066± 0.010 0.8± 0.3 78± 11 58916 K20 0.46± 0.04 0.073± 0.018 118± 9 0.088± 0.014 6.5± 0.4 113± 7 58979 K20 0.26± 0.03 0.118± 0.024 120± 8 0.096± 0.017 5.7± 0.3 114± 6 59006 K20 0.10± 0.02 0.151± 0.015 117± 8 0.067± 0.015 8.3± 0.5 91± 6 Notes. Columns of table are: (1) epoch of observation in MJD; (2) designation of knot; (3) flux density, S , in Jy; (4) distance from the core, R, in mas; (5) position angle with respect to the core, Θ, in degrees; (6) FWHM size of knot, a, in mas; (7) degree of polarization, P, in %; and (8) position angle of polarization, EVPA, in degrees. For model parameters, 1σ uncertainties are provided. Fig. 11. Separation of K20 from the core with time (red triangles) according to modeling; the red line represents a linear approximation of the motion. Blue circles and the dashed line correspond to the light curve of K20 at 43 GHz; the black vertical line indicates the time of the VHE event. modeling. The broadband spectra obtained from the modeling are compared with measured ones in Fig. 12. The corresponding evolution of EED is shown in Fig. 13. Parameters of the model- ing are shown in Table 6. It should be stressed that, mainly due to fixing the Doppler factor, these latter parameters are not uniquely determined (see, e.g., Ahnen et al. 2017a). For example, assuming δ and Γ moti- vated by VLBI measurements, a similar fit to the data can be achieved, albeit with the size of the emission region compressed by about an order of magnitude in periods A, C, and D. However, they allow us to trace the relative evolution of the parameters with the assumption that the beaming did not change during the flare. The gamma-ray spectrum in the modeling is reproduced by almost the EC process alone. The high spectral variability of the X-ray emission is naturally explained by three processes that contribute to it: mainly the EC and SSC emission, with a par- tial component from the synchrotron radiation from the highest energies during period C. Because of its simplicity, the modeling has some caveats. The radio emission is underestimated because of pronounced synchrotron self absorption in the model curve. Such emission is often attributed to a larger scale jet, rather than the blob. How- ever, it should be noted that during the 2020 high state, the radio emission was at a higher level than in historical measurements, and also the emission showed some variability, and therefore it should be also at least partially associated with the high state. It is plausible that, because of the evolution of the flare, low- energy electrons from the blob escape to the large-scale jet with- out being cooled completely. Such an enhancement of the EED in the large-scale jet during the high state would explain higher radio emission. In addition, such escape of high-energy particles along the jet would naturally explain the new component appear- ing in the VLBI follow-up of the flare. In period C (and partially also in period D) the X-ray data show a clear valley between the two peaks. The shape of the low-energy (i.e., falling) part of the valley in this period is not fully reproduced by the model. This part of the SED is strongly dependent on the shape of the high-energy tail of the EED which is also constrained by the highest energy gamma-rays. The sim- plifications in the modeling (homogeneous region, no nonsta- tionary processes modifying EED within the considered period, resulting sharp cut-off of the EED) do not allow us to realisti- cally model the full shape of the valley. The SED in this period might also be affected by fast variability. The slope of the spectrum is not accurately modeled in all the cases. In particular, in period D (and partially also in period C) the optical range would require softer electron distribution than the gamma-ray range. In period C, the NIR data are slightly over- shooting the model suggesting a softer spectrum. This might be connected with the fast variability of the source, or with addi- tional emission of the population of partially cooled particles from the previous phases of the flare. Despite those caveats, it is interesting to see that the obtained model parameters provide a self-consistent description of the main features of the emission during different phases of the flare. Comparing period B to period A, modeling suggests a compres- sion of the emission region coincident with the increase of the minimum Lorentz factor of the electrons and a mild increase of the magnetic field density and the total energy stored in electrons. The increased γmin factor needed in the modeling (see also Katarzyn´ski et al. 2006) might point to a two-stage accelera- tion process. First, injection of particles with such a minimum Lorentz factor has to occur, for example because of accelera- tion in a small potential drop due to reconnection of magnetic fields (see, e.g., Lazarian et al. 2015; Comisso & Sironi 2019). A second process (e.g., Fermi second-order acceleration) would then boost the particles to a power-law spectrum up to maxi- mal gamma factors of γmax. The acceleration coefficient ξ has a rather small value, of the order of 10−7. The value of ξ in the modeling increases by an order of magnitude in period C in order to explain the VHE gamma-ray emission. Such small values are needed to saturate the acceleration process with EC energy losses in order not to overly increase γmax and in turn not A163, page 13 of 19 A&A 647, A163 (2021) 108 1011 1014 1017 1020 1023 1026 [Hz] 10 14 10 13 10 12 10 11 10 10 10 9 10 8 F [e rg cm 2 s 1 ] Synch. SSC EC DT Total SSDC A A U.L. 108 1011 1014 1017 1020 1023 1026 [Hz] 10 14 10 13 10 12 10 11 10 10 10 9 10 8 F [e rg cm 2 s 1 ] Synch. SSC EC DT Total SSDC B B U.L. 108 1011 1014 1017 1020 1023 1026 [Hz] 10 14 10 13 10 12 10 11 10 10 10 9 10 8 F [e rg cm 2 s 1 ] Synch. SSC EC DT Total SSDC C C U.L. 108 1011 1014 1017 1020 1023 1026 [Hz] 10 14 10 13 10 12 10 11 10 10 10 9 10 8 F [e rg cm 2 s 1 ] Synch. SSC EC DT Total SSDC D D U.L. Fig. 12. Multiwavelength SED of QSO B1420+326 in the four periods: A – before the flare, B – optical flare, C – VHE gamma-ray flare, D – after the flare and archival data (gray). Different radiation processes are shown with different line styles: dotted lines – synchrotron, dashed – SSC, dot-dashed – EC, solid – sum of components. Model lines are corrected for EBL absorption according to Domínguez et al. (2011). 100 101 102 103 104 102 103 2 d n e /d [e l. cm 3 ] A B C D 100 101 102 103 104 1052 1053 1054 2 d N e/d [e l.] A B C D Fig. 13. Evolution of the EED in the frame of the blob: energy density (left) and total energy integrated over the blob (right). Different line colors represent different stages of the flare. to overshoot the X-ray emission by synchrotron component. As shown in the modeling, such acceleration would still be efficient enough to explain the observed optical and HE flare. A natu- ral explanation for such low values of ξ is a second-order Fermi process with nonrelativistic scattering centers accelerating elec- trons in the emission region. During period D, the VHE emission requires a further small increase of the ξ parameter coincident with lower magnetic field in order not to overshoot the soft X-ray flux with the synchrotron component. 5. Discussion and conclusions Observations of QSO B1420+326 with MAGIC during the enhanced state allowed us to add a new member to the sparse family of FSRQs emitting in the VHE range. The observations were performed during an impressive flare, with the flux of both SED peaks enhanced by about two orders of magnitude with respect to the low state. Monitoring observations of the source and a massive MWL campaign provided us a dataset that was used to trace the evolution of the EED during the flare. Inter- A163, page 14 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 Table 6. Parameters used for the modeling. Period δ rb [cm] ξ B [G] U′e [1048 erg] p1 γmin p2 γbreak γmax u′e [erg cm−3] u′e/u′B A 40 6.16 × 1016 0.3 × 10−7 0.70 1.18 1.7 1 2.7 63 6900 1.2 × 10−3 0.06 B 40 3.70 × 1016 0.3 × 10−7 0.95 1.76 1.8 15 2.8 104 8000 8.3 × 10−3 0.23 C 40 3.08 × 1016 3.0 × 10−7 0.83 2.12 2.0 10 3.0 125 23700 17.3 × 10−3 0.63 D 40 3.08 × 1016 6.0 × 10−7 0.55 2.35 2.0 10 3.0 125 27300 19.2 × 10−3 1.6 Notes. Columns are: Doppler factor δ (Γ = δ is assumed), co-moving size of the emission region rb, acceleration parameter ξ, magnetic field B, total energy of electron U′e, EED: slope before the break: p1, minimum Lorentz factor γmin, slope after the break p2, the Lorentz factor of the break γbreak, maximum Lorentz factor γmax, electron energy density u′e, energy equipartition u ′ e/u ′ B. Free parameters of the model and derived parameters are put on the left and right side of the double vertical line, respectively. estingly, the synchrotron spectrum in the optical range during the flare (in particular period B) is hard. Comparing to histori- cal data, both low- and high-energy peaks shifted by at least two orders of magnitude in frequency, such a large shift being rare for a FSRQ object. Shifts of the peaks towards higher energies dur- ing high states is a behavior commonly observed in a sister class of objects, the BL Lacs. The spectra are Compton dominated, which is typical for FSRQs. However, the dominance during the peak of the flare is just a factor of a few. Similarly to other VHE-detected FSRQs, we get a satisfac- tory description of the two broad peaks of the SED as the syn- chrotron and EC on the DT radiation field. The valley between the peaks is well constrained by the X-ray data, and we use it to track the changes of the compactness of the blob during the evo- lution of the high state. The modeling scenario is self-consistent in the sense that the shape of the EED is determined by the bal- ance of acceleration and cooling processes. The variability of the emission between different phases of the enhanced state is explained mainly by a combination of variations of the com- pactness of the emission region, the minimal injection energy of electrons, and the increase in the acceleration parameters. In addition, to achieve a satisfactory fit, coincident small variations of the magnetic field, total energy stored in electrons, and an injection slope have been assumed. The optical spectroscopy observations revealed a prominent MgII line that does not show flux variability exceeding the uncer- tainties of the measurements. We explain this as the line being produced within a canonical BLR, which means that it has a much longer timescale of variability. Therefore, the MgII line is a good proxy for estimating the accretion disk luminosity at 2 × 1046 erg s−1. Additionally, a broad FeII bump has been observed, with the luminosity increasing with the increase of the optical continuum emission. The fast variability of the FeII bump suggests that it originates in a much smaller region (possi- bly located close to the jet axis) than the regular BLR. Moreover, as the flux of FeII bump correlates with the synchrotron contin- uum, the bump should be produced farther from the black hole in the vicinity of the jet. Additionally, the shift of the bump to the blue side could be explained if it is produced in a wind sur- rounding the jet. This suggests a possible interaction of the jet with a FeII-emitting cloud. Optical polarimetry that started a few days after optical peak shows a very low polarization and smooth EVPA rotation. This makes QSO B1420+326 another FSRQ in which VHE gamma- ray emission is detected contemporaneous to EVPA rotations. Intriguingly, the VLBI observations in the follow-up observa- tions of the flare show an emission of a superluminal radio knot contemporaneous with the high gamma-ray state. A similar asso- ciation of VHE gamma-ray emission, with EVPA rotation, and VLBI component ejection has previously been suggested for another FSRQ, PKS 1510−089 (Ahnen et al. 2017b). The detection of QSO B1420+326 in the VHE gamma-ray range and the extensive monitoring campaign during this event add another piece to the puzzle of the origin of the highest energy emission of FSRQ objects. It is interesting that some of the emission features associated with these observations of QSO B1420+326 are similar to observations of other FSRQs, in particular of PKS 1510−089, so far the most thoroughly studied FSRQ in the VHE range. QSO B1420+326 might be a cousin of PKS 1510−089, twice as distant but intrinsically more luminous. Acknowledgements. We would like to dedicate this paper to the memory of our friend and colleague, Dr. Valeri Larionov (1950−2020), who enthusias- tically contributed to this and many other projects aimed at understanding blazars. We would like to thank the Instituto de Astrofísica de Canarias for the excellent working conditions at the Observatorio del Roque de los Mucha- chos in La Palma. The financial support of the German BMBF and MPG; the Italian INFN and INAF; the Swiss National Fund SNF; the ERDF under the Spanish MINECO (FPA2017-87859-P, FPA2017-85668-P, FPA2017-82729- C6-2-R, FPA2017-82729-C6-6-R, FPA2017-82729-C6-5-R, AYA2015-71042- P, AYA2016-76012-C3-1-P, ESP2017-87055-C2-2-P, FPA2017-90566-REDC); the Indian Department of Atomic Energy; the Japanese ICRR, the University of Tokyo, JSPS, and MEXT; the Bulgarian Ministry of Education and Science, National RI Roadmap Project DO1-268/16.12.2019 and the Academy of Finland grant nr. 320045 is gratefully acknowledged. This work was also supported by the Spanish Centro de Excelencia “Severo Ochoa” SEV-2016-0588 and SEV- 2015-0548, the Unidad de Excelencia “María de Maeztu” MDM-2014-0369 and the “la Caixa” Foundation (fellowship LCF/BQ/PI18/11630012), by the Croa- tian Science Foundation (HrZZ) Project IP-2016-06-9782 and the University of Rijeka Project 13.12.1.3.02, by the DFG Collaborative Research Centers SFB823/C4 and SFB876/C3, the Polish National Research Centre grant UMO- 2016/22/M/ST9/00382 and by the Brazilian MCTIC, CNPq and FAPERJ. The Fermi-LAT Collaboration acknowledges generous ongoing support from a num- ber of agencies and institutes that have supported both the development and the operation of the LAT, as well as scientific data analysis. These include the National Aeronautics and Space Administration and the Department of Energy in the United States; the Commissariat à l’Energie Atomique and the Centre National de la Recherche Scientifique/Institut National de Physique Nucléaire et de Physique des Particules in France; the Agenzia Spaziale Italiana and the Isti- tuto Nazionale di Fisica Nucleare in Italy; the Ministry of Education, Culture, Sports, Science and Technology (MEXT), High Energy Accelerator Research Organization (KEK), and Japan Aerospace Exploration Agency (JAXA) in Japan; and the K. A. Wallenberg Foundation, the Swedish Research Council, and the Swedish National Space Board in Sweden. Additional support for science analysis during the operations phase is gratefully acknowledged from the Isti- tuto Nazionale di Astrofisica in Italy and the Centre National d’Etudes Spatiales in France. This work was performed in part under DOE Contract DE-AC02- 76SF00515. This publication makes use of data obtained at the Metsähovi Radio Observatory, operated by Aalto University in Finland. This research has made use of data from the OVRO 40 m monitoring program Richards et al. (2011) which is supported in part by NASA grants NNX08AW31G, NNX11A043G, and NNX14AQ89G and NSF grants AST-0808050 and AST-1109911. This study was based in part on observations conducted using the 1.8 m Perkins Telescope Observatory (PTO) in Arizona, which is owned and operated by Boston Univer- sity. The research at Boston University was supported in part by NASA Fermi GI program grants 80NSSC17K0649, 80NSSC19K1504, and 80NSSC19K1505. A163, page 15 of 19 A&A 647, A163 (2021) We thank the ASAS-SN team for making their data publicly available. The VLBA is an instrument of the National Radio Astronomy Observatory. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated by Associated Universities, Inc. This work made use of the Lowell Discovery Telescope (formerly Discovery Channel Telescope) at Lowell Observatory. Lowell is a private, nonprofit institution dedicated to astrophys- ical research and public appreciation of astronomy and operates the LDT in partnership with Boston University, the University of Maryland, the University of Toledo, Northern Arizona University, and Yale University. We acknowledge support by Bulgarian National Science Fund under grant DN18-10/2017 and National RI Roadmap Projects DO1-277/16.12.2019 and DO1-268/16.12.2019 of the Ministry of Education and Science of the Republic of Bulgaria. References Abdollahi, S., Acero, F., Ackermann, M., et al. 2020, ApJS, 247, 33 Abeysekara, A. U., Archambault, S., Archer, A., et al. 2015, ApJ, 815, L22 Acero, F., Ackermann, M., Ajello, M., et al. 2016, ApJS, 223, 26 Ahnen, M. L., Ansoldi, S., Antonelli, L. A., et al. 2015, ApJ, 815, L23 Ahnen, M. L., Ansoldi, S., Antonelli, L. A., et al. 2017a, A&A, 603, A31 Ahnen, M. L., Ansoldi, S., Antonelli, L. A., et al. 2017b, A&A, 603, A29 Ajello, M., Atwood, W. B., Baldini, L., et al. 2017, ApJS, 232, 18 Albareti, F. D., Allende Prieto, C., Almeida, A., et al. 2017, ApJS, 233, 25 Aleksic´, J., Antonelli, L. A., Antoranz, P., et al. 2011, ApJ, 730, L8 Aleksic´, J., Ansoldi, S., Antonelli, L. A., et al. 2014, A&A, 569, A46 Aleksic´, J., Ansoldi, S., Antonelli, L. A., et al. 2016a, Astropart. Phys., 72, 61 Aleksic´, J., Ansoldi, S., Antonelli, L. A., et al. 2016b, Astropart. Phys., 72, 76 Angioni, R. 2019, ATel, 12942, 1 Angioni, R., Nesci, R., Finke, J. D., et al. 2019, A&A, 627, A140 Atwood, W. B., Abdo, A. A., Ackermann, M., et al. 2009, ApJ, 697, 1071 Atwood, W., Albert, A., Baldini, L., et al. 2013, ArXiv e-prints [arXiv:1303.3514] Baldwin, J. A., Ferland, G. J., Korista, K. T., et al. 2004, ApJ, 615, 610 Ben Bekhti, N., Flöer, L., Keller, R., et al. 2016, A&A, 594, A116 Bessell, M. S., Castelli, F., & Plez, B. 1998, A&A, 333, 231 Bianchi, L., Efremova, B., Herald, J., et al. 2011, MNRAS, 411, 2770 Boller, T., Freyberg, M. J., Trümper, J., et al. 2016, A&A, 588, A103 Breeveld, A. A., Curran, P. A., Hoversten, E. A., et al. 2010, MNRAS, 406, 1687 Burbidge, G. R., Jones, T. W., & O’Dell, S. L. 1974, ApJ, 193, 43 Burrows, D. N., Hill, J. E., Nousek, J. A., et al. 2005, Space Sci. Rev., 120, 165 Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245 Carrasco, L., Hernandez-Utrera, O., Vazquez, S., et al. 2017, Rev. Mex. Astron. Astrofis., 53, 497 Cash, W. 1979, ApJ, 228, 939 Ceribella, G., D’Amico, G., Dazzi, F., et al. 2019, 36th International Cosmic Ray Conference (ICRC 2019), 645 Ciprini, S. 2018, ATel, 12277, 1 Ciprini, S., & Cheung, C. C. 2020, ATel, 13382, 1 Cohen, M., Weaton, W. A., & Megeath, S. F. 2003, AJ, 126, 1090 Comisso, L., & Sironi, L. 2019, ApJ, 886, 122 Condon, J. J., Cotton, W. D., Greisen, E. W., et al. 1998, AJ, 115, 1693 Costamante, L., Cutini, S., Tosti, G., et al. 2018, MNRAS, 477, 4749 Covino, S., Zerbi, F. M., Chincarini, G., et al. 2004, Proc. SPIE, 5492, 1613 Cutini, S., Ciprini, S., Orienti, M., et al. 2014, MNRAS, 445, 4316 Dermer, C. D., & Menon, G. 2009, in High Energy Radiation from Black Holes: Gamma Rays, Cosmic Rays, and Neutrinos, eds. C. D. Dermer, & G. Menon (Princeton: Princeton University Press) D’Ammando, F., Raiteri, C. M., Villata, M., et al. 2011, A&A, 529, A145 D’Ammando, F., Raiteri, C. M., Villata, M., et al. 2019, MNRAS, 490, 5300 D’Ammando, F., Fugazza, D., & Covino, S. 2020, ATel, 13428, 1 Domínguez, A., Primack, J. R., Rosario, D. J., et al. 2011, MNRAS, 410, 2556 Drury, L. O. 1983, Rep. Progr. Phys., 46, 973 Evans, P. A., Beardmore, A. P., Page, K. L., et al. 2009, MNRAS, 397, 1177 Fallah Ramazani, V., Bonnoli, G., D’Ammando, F., et al. 2020, ATel, 13417, 1 Finke, J. D. 2016, ApJ, 830, 94 Fruck, C., & Gaug, M. 2015, Eur. Phys. J. Web Conf., 89, 02003 García, J. R., Dazzi, F., Häfner, D., et al. 2014, Proceedings of 33th International Cosmic Ray Conference (ICRC 2014) Gehrels, N., Chincarini, G., Giommi, P., et al. 2004, ApJ, 611, 1005 Ghisellini, G. 2016, Galaxies, 4, 36 Ghisellini, G., & Tavecchio, F. 2009, MNRAS, 397, 985 Ghisellini, G., Tavecchio, F., Foschini, L., et al. 2013, MNRAS, 432, L66 Ghisellini, G., Tavecchio, F., Maraschi, L., et al. 2014, Nature, 515, 376 Gregory, P. C., Scott, W. K., Douglas, K., et al. 1996, ApJS, 103, 427 Healey, S. E., Romani, R. W., Taylor, G. B., et al. 2007, ApJS, 171, 61 Henden, A. A., Templeton, M., Terrell, D., et al. 2016, VizieR On-line Data Catalog: II/336 H.E.S.S. Collaboration (Abdalla, H., et al.) 2020, A&A, 633, A162 Hewett, P. C., & Wild, V. 2010, MNRAS, 405, 2302 Jackson, N., Battye, R. A., Browne, I. W. A., et al. 2007, MNRAS, 376, 371 Jansen, F., Lumb, D., Altieri, B., et al. 2001, A&A, 365, L1 Jester, S., Schneider, D. P., Richards, G. T., et al. 2005, AJ, 130, 873 Jorstad, S. G., Marscher, A. P., Lister, M. L., et al. 2005, AJ, 130, 1418 Jorstad, S. G., Marscher, A. P., Larionov, V. M., et al. 2010, ApJ, 715, 362 Jorstad, S. G., Marscher, A. P., Morozova, D. A., et al. 2017, ApJ, 846, 98 Katarzyn´ski, K., Ghisellini, G., Tavecchio, F., et al. 2006, MNRAS, 368, L52 Kharinov, M. A. 2020, ATel, 13479, 1 Kochanek, C. S., Shappee, B. J., Stanek, K. Z., et al. 2017, PASP, 129, 104502 Larionov, V. M., Jorstad, S. G., Marscher, A. P., et al. 2008, A&A, 492, 389 Lazarian, A., Eyink, G. L., Vishniac, E. T., et al. 2015, Magnetic Fields in Diffuse Media (Berlin: Springer), 311 Lindfors, E. 2015, IAU Symp., 313, 27 Liu, H. T., & Bai, J. M. 2006, ApJ, 653, 1089 Lott, B., Escande, L., Larsson, S., et al. 2012, A&A, 544, A6 MAGIC Collaboration (Acciari, V. A., et al.) 2018, A&A, 619, A159 Marchini, A., Bonnoli, G., Bellizzi, L., et al. 2019, ATel, 12914, 1 Mattox, J. R., Bertsch, D. L., Chiang, J., et al. 1996, ApJ, 461, 396 Meyer, M., Scargle, J. D., & Blandford, R. D. 2019, ApJ, 877, 39 Minev, M., Kurtenkv, A., & Ovcharov, E. 2020, ATel, 13421, 1 Mirzoyan, R. 2020, ATel, 13412, 1 Munari, U. 2012, J. Am. Assoc. Var. Star Obs., 40, 582 Myers, S. T., Jackson, N. J., Browne, I. W. A., et al. 2003, MNRAS, 341, 1 Nigro, C., Sitarek, J., Craig, M., & Gliwny, P. 2020, https://doi.org/10. 5281/zenodo.4055176 Oh, K., Koss, M., Markwardt, C. B., et al. 2018, ApJS, 235, 4 Ovcharov, E. P., Kurtenkov, A., Metodieva, Y., et al. 2014, Bulg. Astron. J., 21, 19 Poole, T. S., Breeveld, A. A., Page, M. J., et al. 2008, MNRAS, 383, 627 Planck Collaboration XXVIII. 2014, A&A, 571, A28 Planck Collaboration VI. 2020, A&A, 641, A6 Rani, B., Jorstad, S. G., Marscher, A. P., et al. 2018, ApJ, 858, 80 Richards, J. L., Max-Moerbeck, W., Pavlidou, V., et al. 2011, ApJS, 194, 29 Rolke, W. A., López, A. M., & Conrad, J. 2005, Nucl. Instrum. Meth. Phys. Res. Sect. A, 551, 493 Roming, P. W. A., Kennedy, T. E., Mason, K. O., et al. 2005, Space Sci. Rev., 120, 95 Sambruna, R. M., Maraschi, L., & Urry, C. M. 1996, ApJ, 463, 444 Schlafly, E. F., & Finkbeiner, D. P. 2011, ApJ, 737, 103 Shappee, B. J., Prieto, J. L., Grupe, D., et al. 2014, ApJ, 788, 48 Shayduk, M. 2013, Proceedings of the 33rd International Cosmic Ray Conference (ICRC 2013) Shepherd, M. C. 1997, in Astronomical Data Analysis Software and Systems VI, eds. G. Hunt, & H. E. Payne (San Francisco: ASP), ASP Conf. Ser., 125, 77 Srivastava, M. K., Jangra, M., Dixit, V., et al. 2018, Proc. SPIE, 10702, 107024I Stanek, K. Z., Kochanek, C. S., Thompson, T. A., et al. 2017, ATel, 11110, 1 Teräsranta, H., Tornikoski, M., Mujunen, A., et al. 1998, A&AS, 132, 305 Vacca, W. D., Cushing, M. C., & Rayner, J. T. 2002, PASP, 115, 389 Vanden Berk, D. E., Richards, G. T., & Bauer, A. 2001, AJ, 122, 549 van den Berg, J. P., Böttcher, M., Domínguez, A., et al. 2019, ApJ, 874, 47 van Moorsel, G., Kemball, A., & Greisen, E. 1996, in Astronomical Data Analysis Software and Systems V, eds. G. H. Jacoby, & J. Barnes, ASP Conf. Ser., 101, 37 Vaughan, S., Edelson, R., Warwick, R. S., et al. 2003, MNRAS, 345, 1271 White, R. L., Becker, R. H., Helfand, D. J., et al. 1997, ApJ, 475, 479 Wilms, J., Allen, A., & McCray, R. 2000, ApJ, 542, 914 Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, AJ, 140, 1868 Wood, M., Caputo, R., Charles, E., et al. 2017, 35th International Cosmic Ray Conference (ICRC 2017), 824 Zacharias, M., Dominis Prester, D., Jankowsky, F., et al. 2019, Galaxies, 7, 41 Zanin, R., Carmona, E., Sitarek, J., et al. 2013, Proc. 33rd ICRC, Rio de Janeiro, Brazil, 773 Zerbi, R. M., Chincarini, G., Ghisellini, G., et al. 2001, Astron. Nachr., 322, 275 1 Instituto de Astrofísica de Canarias and Dpto. de Astrofísica, Universidad de La Laguna, 38200 La Laguna, Tenerife, Spain 2 Università di Udine and INFN Trieste, 33100 Udine, Italy 3 National Institute for Astrophysics (INAF), 00136 Rome, Italy 4 ETH Zürich, 8093 Zürich, Switzerland 5 Institut de Física d’Altes Energies (IFAE), The Barcelona Institute of Science and Technology (BIST), 08193 Bellaterra, Barcelona, Spain A163, page 16 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 6 Japanese MAGIC Group: Institute for Cosmic Ray Research (ICRR), The University of Tokyo, Kashiwa 277-8582, Chiba, Japan 7 Technische Universität Dortmund, 44221 Dortmund, Germany 8 Croatian MAGIC Group: University of Zagreb, Faculty of Electrical Engineering and Computing (FER), 10000 Zagreb, Croatia 9 IPARCOS Institute and EMFTEL Department, Universidad Com- plutense de Madrid, 28040 Madrid, Spain 10 Centro Brasileiro de Pesquisas Físicas (CBPF), 22290-180 URCA, Rio de Janeiro, RJ, Brazil 11 University of Lodz, Faculty of Physics and Applied Informatics, Department of Astrophysics, 90-236 Lodz, Poland 12 Università di Siena and INFN Pisa, 53100 Siena, Italy 13 Deutsches Elektronen-Synchrotron (DESY), 15738 Zeuthen, Ger- many 14 Università di Padova and INFN, 35131 Padova, Italy 15 INFN MAGIC Group: INFN Sezione di Torino and Università degli Studi di Torino, 10125 Torino, Italy 16 Max-Planck-Institut für Physik, 80805 München, Germany 17 Università di Pisa and INFN Pisa, 56126 Pisa, Italy 18 Universitat de Barcelona, ICCUB, IEEC-UB, 08028 Barcelona, Spain 19 Armenian MAGIC Group: A. Alikhanyan National Science Labora- tory, Yerevan, Armenia 20 Centro de Investigaciones Energéticas, Medioambientales y Tec- nológicas, 28040 Madrid, Spain 21 INFN MAGIC Group: INFN Sezione di Bari and Dipartimento Interateneo di Fisica dell’Università e del Politecnico di Bari, 70125 Bari, Italy 22 Croatian MAGIC Group: University of Rijeka, Department of Physics, 51000 Rijeka, Croatia 23 Universität Würzburg, 97074 Würzburg, Germany 24 Finnish MAGIC Group: Finnish Centre for Astronomy with ESO, University of Turku, 20014 Turku, Finland 25 Departament de Física, and CERES-IEEC, Universitat Autònoma de Barcelona, 08193 Bellaterra, Spain 26 Armenian MAGIC Group: ICRANet-Armenia at NAS RA, Yerevan, Armenia 27 Croatian MAGIC Group: University of Split, Faculty of Electri- cal Engineering, Mechanical Engineering and Naval Architecture (FESB), 21000 Split, Croatia 28 Croatian MAGIC Group: Josip Juraj Strossmayer University of Osi- jek, Department of Physics, 31000 Osijek, Croatia 29 Japanese MAGIC Group: Department of Physics, Kyoto University, 606-8502 Kyoto, Japan 30 Japanese MAGIC Group: Department of Physics, Tokai University, Hiratsuka 259-1292, Kanagawa, Japan 31 Saha Institute of Nuclear Physics, HBNI, 1/AF Bidhannagar, Salt Lake, Sector-1, Kolkata 700064, India 32 Inst. for Nucl. Research and Nucl. Energy, Bulgarian Academy of Sciences, 1784 Sofia, Bulgaria 33 Finnish MAGIC Group: Astronomy Research Unit, University of Oulu, 90014 Oulu, Finland 34 Croatian MAGIC Group: Rud¯er Boškovic´ Institute, 10000 Zagreb, Croatia 35 INFN MAGIC Group: INFN Sezione di Perugia, 06123 Perugia, Italy 36 INFN MAGIC Group: INFN Roma Tor Vergata, 00133 Roma, Italy 37 Now at University of Innsbruck, Innsbruck, Austria 38 Also at Port d’Informació Científica (PIC), 08193 Bellaterra, Barcelona, Spain 39 Also at Dipartimento di Fisica, Università di Trieste, 34127 Trieste, Italy 40 Max-Planck-Institut für Physik, 80805 München, Germany 41 Also at INAF Trieste and Dept. of Physics and Astronomy, Univer- sity of Bologna, Bologna, Italy 42 Japanese MAGIC Group: Institute for Cosmic Ray Research (ICRR), The University of Tokyo, Kashiwa 277-8582, Chiba, Japan 43 ASI Space Science Data Center, Via del Politecnico, snc., 00133 Rome, Italy 44 INFN – Roma Tor Vergata, Via della Ricerca Scientifica, 1., 00133 Rome, Italy 45 INAF – IRA Bologna, Via P. Gobetti 101, 40129 Bologna, Italy 46 Istituto Nazionale di Fisica Nucleare, Sezione di Perugia, 06123 Perugia, Italy 47 Naval Research Laboratory, Washington, DC 20375, USA 48 INAF – Istituto di Astrofisica e Planetologia Spaziali, Via Fosso del Cavaliere, 100, 00133 Rome, Italy 49 Physical Research Laboratory, Ahmedabad, India 50 Institute of Astronomy and NAO, Bulgarian Academy of Sciences, 72 Tsarigradsko Shose Blvd., 1784 Sofia, Bulgaria 51 Department of Astronomy, Faculty of Physics, University of Sofia, 1164 Sofia, Bulgaria 52 Università di Siena, 53100 Siena, Italy 53 Instituto Nacional de Astrofisica, Optica & Electronics Tonantzintla, Puebla, Mexico 54 Aalto University Metsähovi Radio Observatory, Metsähovintie 114, 02540 Kylmälä, Finland 55 Aalto University Department of Electronics and Nanoengineering, PO Box 15500, 00076 Aalto, Finland 56 Finnish Centre for Astronomy with ESO (FINCA), University of Turku, Vesilinnantie 5, 20014 Turku, Finland 57 Institute for Astrophysical Research, Boston University, 725 Com- monwealth Avenue, Boston, MA 02215, USA 58 Astronomical Institute, St. Petersburg University, Universitetskij Pr. 28, Petrodvorets 198504, St. Petersburg, Russia 59 Pulkovo Observatory, St. Petersburg, Russia 60 Crimean Astrophysical Observatory, Crimea 61 Owens Valley Radio Observatory, California Institute of Technol- ogy, Pasadena, CA 91125, USA 62 Institute of Astrophysics, Foundation for Research and Technology- Hellas, 71110 Heraklion, Greece 63 Department of Physics, Univ. of Crete, 70013 Heraklion, Greece 64 Departamento de Astronomía, Universidad de Chile, Camino El Observatorio 1515, Las Condes, Santiago, Chile 65 CePIA, Departamento de Astronomía, Universidad de Concepción, Concepción, Chile A163, page 17 of 19 A&A 647, A163 (2021) Appendix A: X-ray short-term variations, spectral fits and long-term multiwavelength behaviour A.1. XMM-Newton and Swift-XRT short-term variations We investigated short-term variability of the source in UV and X-rays by analyzing Swift and XMM-Newton data. Correcting for instrumental artifacts and pile-up, no significant (>3-σ) increase of count rate was observed in the background-subtracted consec- utive XRT observation segments. We also searched for significant (>3-σ) changes of magni- tude between two consecutive UVOT exposures collected with the same filter at the same epoch. This resulted in two events observed in w1 filter on 2020 January 19 (MJD = 58867) and 31 (MJD = 58879), with a change of 0.16 mag in 27.6 ks and 0.41 mag in 34.4 ks, respectively. Moreover, we produced a light curve of the XMM-Newton pn count rate with a bin of 500 s (Fig. A.1). The light curve shows only moderate variability, with the count rate varying between 2.08 and 2.86 cps. The fractional variability (see Vaughan et al. 2003 for details) is 0.064± 0.005. A larger variability was observed in the second part of the observation, in particular an increase of ∼15% of the count rate at the time of the highest peak. A.2. Swift-XRT spectral fits In Table A.1 we report the results of analysis of all Swift-XRT observations, including both historical observations and the ones performed during the high state of the source shown in detail in Fig. 1. A.3. Long-term behaviour In Fig. A.2 we report the monitoring observations of QSO B1420+326 in order to put the flaring state of 2020 January and February in the context of the long-term behavior of the source. It is clear that the flaring period has been unique in the Fermi-LAT dataset of this source. Similarly, the radio flux during the flare is also unique compared to previous measurements. On the other hand, there is a prior report of a similar magnitude opti- cal flare on 2016 March 11 (MJD = 57458.5) in the ASAS-SN monitoring data (Stanek et al. 2017), albeit without a HE coun- terpart. Past X-ray data are unfortunately too sparse (and biased by Target of Opportunity observations) to make conclusions as to the typical behavior of the source. Fig. A.1. XMM-Newton EPIC pn light curve over 0.3−10 keV with 500 s bins of QSO B1420+326. Table A.1. Log and fitting results of Swift-XRT observations of QSO B2 1420+326 using a PL model. Date MJD Net exposure time Photon index Flux0.3−10 keV (UT) (s) (ΓX) (10−12 erg cm−2 s−1) 2018-02-22 58171.915197 2872 1.49± 0.22 1.99± 0.37 2018-12-12 58464.132329 1818 1.51± 0.25 2.76± 0.54 2018-12-14 58466.253218 2110 1.37± 0.18 4.15± 0.65 2018-12-16 58468.386537 1943 1.50± 0.21 3.22± 0.54 2019-06-25 58659.880024 1975 1.89± 0.21 3.18± 0.45 2019-06-27/29 58663 2008 1.25± 0.24 3.07± 0.64 2019-07-12 58676.328225 1593 2.33± 0.15 7.11± 0.69 2019-07-17 58681.609499 1983 2.17± 0.15 4.83± 0.48 2019-07-22 58686.722614 1885 1.45± 0.22 2.98± 0.55 2019-12-13 58830.935420 2023 1.44± 0.13 7.85± 0.87 2020-01-02 58850.318109 2495 1.33± 0.17 5.47± 0.75 2020-01-05 58853.943061 1666 1.38± 0.27 4.09± 0.87 2020-01-08 58856.665569 1880 1.69± 0.20 3.73± 0.55 2020-01-11 58859.830600 1983 1.81± 0.17 5.54± 0.70 2020-01-19 58867.533771 1626 1.87± 0.15 5.88± 0.65 2020-01-21 58870.011577 1546 1.94± 0.16 6.00± 0.65 2020-01-24 58872.175468 1783 1.84± 0.14 6.57± 0.65 2020-01-25 58873.171868 1798 2.10± 0.14 7.12± 0.69 2020-01-27 58875.267506 1806 1.73± 0.19 4.42± 0.61 2020-01-28 58876.193806 2035 1.47± 0.18 4.80± 0.67 2020-01-31 58879.685875 1231 1.68± 0.19 5.88± 0.83 2020-02-01 58880.114444 1681 1.58± 0.16 5.87± 0.77 2020-02-05 58884.201846 2025 1.49± 0.15 6.30± 0.75 2020-02-10 58889.473095 2417 1.57± 0.18 3.50± 0.48 Notes. Fluxes are corrected for the Galactic absorption. A163, page 18 of 19 MAGIC Collaboration: MWL campaign on QSO B1420+326 Fermi-LAT F >0.1GeV 55000 55500 56000 56500 57000 57500 58000 58500 59000 0 5 10 15 ] - 1 s - 2 cm - 7 F [10 Fermi-LAT index 55000 55500 56000 56500 57000 57500 58000 58500 59000 2 2.5 3 Ph ot on in de x Swift XRT 0.3-10 keV 55000 55500 56000 56500 57000 57500 58000 58500 59000 0 2 4 6 8 ] - 1 s - 2 er g cm - 12 F[ 10 0.3-10 keV index 55000 55500 56000 56500 57000 57500 58000 58500 590001 1.5 2 2.5 in de x W2-band 55000 55500 56000 56500 57000 57500 58000 58500 59000 0 1 2 3 F [m Jy ] V-band 55000 55500 56000 56500 57000 57500 58000 58500 59000 0 5 10 15 F [m Jy ] ASAS-SN V&g ASAS-SN V&g U.L. UVOT V OVRO 55000 55500 56000 56500 57000 57500 58000 58500 59000 Time [MJD] 0 0.5 1 1.5 F [Jy ] Fig. A.2. Long-term MWL light curve of QSO B1420+326 (see titles and legends of individual panels). The gray-shaded region shows the flaring period in the beginning of 2020. A163, page 19 of 19