<<

Draft version May 20, 2020 Typeset using LATEX twocolumn style in AASTeX62

Stellar Flares from the First Tess Data Release: Exploring a New Sample of M dwarfs

Maximilian N. Gunther¨ ,1, 2 Zhuchang Zhan,3 Sara Seager,1, 3 Paul B. Rimmer,4, 5, 6, 7 Sukrit Ranjan,3, 8 Keivan G. Stassun,9 Ryan J. Oelkers,9 Tansu Daylan,1, 10 Elisabeth Newton,1, 11, 12 Martti H. Kristiansen,13, 14 Katalin Olah,15 Edward Gillen,5, 16 Saul Rappaport,1 George R. Ricker,1 Roland K. Vanderspek,1 David W. Latham,17 Joshua N. Winn,18 Jon M. Jenkins,19 Ana Glidden,20, 1 Michael Fausnaugh,1 Alan M. Levine,1 Jason A. Dittmann,21, 3 Samuel N. Quinn,17 Akshata Krishnamurthy,22 and Eric B. Ting19

1Department of , and Kavli Institute for and Space Research, Massachusetts Institute of Technology, Cambridge, MA 02139, USA 2Juan Carlos Torres 3Department of Earth, Atmospheric, and Planetary Sciences, MIT, 77 Massachusetts Avenue, Cambridge, MA 02139, USA 4Department of Earth Sciences, University of Cambridge, Downing Street, Cambridge, CB2 3EQ, UK 5Astrophysics Group, Cavendish Laboratory, J.J. Thomson Avenue, Cambridge CB3 0HE, UK 6MRC Laboratory of Molecular Biology, Francis Crick Ave, Cambridge CB2 0QH 7SCOL Senior Fellow 8SCOL Postdoctoral Fellow 9Vanderbilt University, Department of Physics & , 6301 Stevenson Center Ln., Nashville, TN 37235, USA 10Kavli Fellow 11Dartmouth College, Hanover, NH 03755, USA 12NSF Astronomy and Astrophysics Postdoctoral Fellow 13DTU Space, National Space Institute, Technical University of Denmark, Elektrovej 327, DK-2800 Lyngby, Denmark 14Brorfelde Observatory, Observator Gyldenkernes Vej 7, DK-4340 Tølløse, Denmark 15Konkoly Observatory, CSFK, H-1121 Budapest, Konkoly Thege M. ’ut 15-17, Hungary 16Winton Fellow 17Center for Astrophysics — Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138 18Department of Astrophysical Sciences, Princeton University, 4 Ivy Lane, Princeton, NJ 08544, USA 19NASA Ames Research Center, Moffett Field, CA, 94035, USA 20Department of Earth and Planetary Sciences, MIT, 77 Massachusetts Avenue, Cambridge, MA 02139, USA 2151 Pegasi b Postdoctoral Fellow 22Department of Aeronautics and Astronautics, MIT, 77 Massachusetts Avenue, Cambridge, MA 02139, USA

ABSTRACT We perform a study of stellar flares for the 24,809 stars observed with 2 minute cadence during the first two months of the TESS mission. Flares may erode ’ atmospheres and impact their habitability, but might also trigger the genesis of life around small stars. TESS provides a new sample of bright dwarf stars in our galactic neighborhood, collecting data for thousands of M dwarfs that might host habitable exoplanets. Here, we use an automated search for flares accompanied by visual inspection. Then, our public allesfitter code robustly selects the appropriate model for potentially complex flares via Bayesian evidence. We identify 1228 flaring stars, 673 of which are M dwarfs. Among 8695 flares in total, the largest superflare increased the stellar brightness by a factor of 16.1. arXiv:1901.00443v2 [astro-ph.EP] 19 May 2020 Bolometric flare energies range from 1031.0 to 1036.9 erg, with a median of 1033.1 erg. Furthermore, we study the flare rate and energy as a function of stellar type and rotation period. We solidify past findings that fast rotating M dwarfs are the most likely to flare, and that their flare amplitude is independent of the rotation period. Finally, we link our results to criteria for prebiotic chemistry, atmospheric loss through coronal mass ejections, and sterilization. Four of our flaring M dwarfs host candidates alerted on by TESS, for which we discuss how these effects can impact life.

Corresponding author: Maximilian N. G¨unther [email protected] 2

With upcoming TESS data releases, our flare analysis can be expanded to almost all bright small stars, aiding in defining criteria for exoplanet habitability.

Keywords: stars: flare, planetary systems, planets and satellites: general

1. INTRODUCTION of planetary atmospheres, meaning their impact must be The most extreme solar flare ever recorded, the ‘Car- considered when interpreting atmospheric signals from rington event’, hit the Earth in 1859 (Carrington 1859; a planetary atmosphere in search of and Hodgson 1859). It released a flare energy of 1032 erg, and geosignatures (Grenfell et al. 2012; Venot et al. 2016; came accompanied by a coronal mass ejection (CME) Airapetian et al. 2017). In sum, it is critical to con- which interacted with the Earth’s magnetic field and strain the flare properties of exoplanet host stars to un- led to destructive consequences. Stellar flares like the derstand the evolution and habitability of their planets, Carrington event are explosive magnetic reconnection and to robustly characterize their atmospheres. events in a star’s magnetosphere, releasing bursts of M dwarfs are of primary interest in the search for hab- isotropic radiation (see e.g. Benz & G¨udel 2010; Shi- itable exoplanets for several reasons. First, they consti- bata & Takasao 2016; Doyle et al. 2018). Over short tute a large fraction (∼70%) of the stellar population time scales of minutes to a few hours, they emit energy (Dole 1964; Henry et al. 1994; Reid et al. 2004; Covey ranging form 1023 erg (nanoflares; e.g. Parnell & Jupp et al. 2008). Second, their small radii and low temper- 2000) to 1033 − 1038 erg (superflares; e.g. Shibayama atures enable the detection and atmospheric character- et al. 2013). Most of the emission is in the X-ray spec- ization of habitable planets on short orbits (Kalteneg- trum, but in extreme cases (like the Carrington event) ger & Traub 2009). For example, we know three exo- 1 flares are also visible in white light. Large flares can Earths in the ‘liquid- habitable zone’ of the red thus be detected with optical photometric surveys such dwarf TRAPPIST-1 (Gillon et al. 2017), one around as those dedicated for exoplanet detection. CMEs, on LHS 1140b (Dittmann et al. 2017), and one around the other hand, are clouds of charged particles that get Proxima Centauri (Anglada-Escud´eet al. 2016). Never- ejected into space with a constrained direction. Large theless, other factors affecting the habitability of these flares are often accompanied by CMEs, but both events systems remain largely unknown. More information on can also appear independently. the frequency and energy of the host stars’ flares can Flares and CMEs on stars hosting exoplanets can be help better characterize the habitability of these plan- even stronger and more frequent than those on the Sun, ets. and can play a major role in planetary evolution and Seminal work on flares from selected stars includes the habitability. Flares may contribute to atmospheric ero- study by Lacy et al.(1976), and was later expanded with sion, destroy ozone layers on oxic planets, and act as surveys like the Sloan Digital Sky Survey (York et al. stressors for surface life (e.g. Segura et al. 2003; Lam- 2000; Kowalski et al. 2009). The largest number of flare mer et al. 2007; Scalo et al. 2007; Segura et al. 2010; Sea- discoveries to date come from distant stars observed by ger et al. 2013; Lingam & Loeb 2017; Atri 2017; Tilley the Kepler mission (Borucki et al. 2010). Extensive cat- et al. 2019; O’Malley-James & Kaltenegger 2018). In the alogs of stellar flares from Kepler were provided by, e.g., most extreme scenario, intense flare activity could ren- Walkowicz et al.(2011), Hawley et al.(2014), Daven- der the immediate planet surface uninhabitable, though port(2016) and Van Doorsselaere et al.(2017). The life could survive in the ocean, in rocks, or under shal- adjusted observing strategy of K2 allowed the study of low layers of soil or dust (Diaz & Schulze-Makuch 2006; bright M dwarfs in more detail (Stelzer et al. 2016). Kiang et al. 2007; Bryce et al. 2015; Tilley et al. 2019; Ground-based wide-field surveys like the Next Gener- Estrela & Valio 2018). More optimistically, flares may ation Survey (NGTS; Wheatley et al. 2018), also power prebiotic chemistry, produce surface biosig- EVRYSCOPE (Law et al. 2015), and others have also natures, or serve as a source for otherwise scarce visible- contributed to flare catalogs for bright nearby objects light photosynthesis on planets orbiting M dwarfs (Bj¨orn & others 2015; Airapetian & others 2016; Mullan & Bais 1 Note that definitions of the liquid-water habitable zone differ 2018). Indeed, recent work suggests that flares might and depend on the planet mass, atmospheric composition, for- be the only means for delivering the UV photons which mation and migration history, tidal locking of the orbits, space weather, and other factors (see e.g. Huang 1959; Kasting et al. have been proposed to be required to initiate life on exo- 1993; Pierrehumbert & Gaidos 2011; Kopparapu et al. 2013; Cul- planets around M dwarf stars (Ranjan et al. 2017; Rim- lum et al. 2014; Kasting et al. 2014; Kopparapu et al. 2014; mer et al. 2018). Finally, flares can alter the chemistry Ramirez & Kaltenegger 2014; Cullum & Stevens 2016). Stellar Flares from the First Tess Data Release 3

(e.g. Jackman et al. 2018, 2019; Howard et al. 2018). To search for potential flaring stars, we analyze the Dedicated M dwarf surveys like MEarth (Nutzman & 24,809 targets with TESS light curves from sectors 1 and Charbonneau 2008) provide additional possibilities for 2. We start from the pre-search data conditioned simple flare studies for the smallest stars. For example, Mon- aperture (PDC-SAP) lightcurves, which are detrended drik et al.(2019) reported flares on 32 mid-to-late M for instrument systematics. Additionally, we detrend dwarfs. Complementary, detailed spectroscopic moni- each orbit using a spline fit to remove any remaining toring campaigns of selected M dwarfs give insight into long-term variations in the light curves. We then re- their flares over multiple wavebands (e.g. MUSCLES; move strong periodicity from the lightcurves, which is France et al. 2016). However, the number of bright generally caused by stellar variability or rotation. For M dwarfs observed for high-precision and high-cadence this, we compute a Lomb-Scargle periodagram, and re- flare studies is still limited, and the limited precision of move a periodic signal using a sine wave fit if two crite- ground-based photometric observations only allows for ria are fulfilled: first, the signal must have a false alarm the detection of the most energetic flares. probability below 0.01; second, the standard deviation The TESS mission (Ricker et al. 2014), launched in of the residuals must decrease. We repeat this process April 2018, provides the opportunity to study flares on at maximum three times. early to late M dwarfs. TESS is specifically designed to Next, we compute a running median using a global observe bright, small stars in the solar neighborhood. and a local view. For the global view, the local me- It will photometrically monitor tens of thousands of dian flux and standard deviation, σ, is calculated using M dwarfs, which are bright enough to allow the study a 1024 data point bin (i.e. 1.4 d). This is meant to catch of flares with a high signal-to-noise ratio. the largest and longest flares. For the local view, we use Here, we present findings derived from the first two a 128 data bin (i.e. 4.3 h) to catch smaller and shorter months of TESS data (i.e., sectors 1 and 2). Section2 flares. We first run the global view iteratively and mask describes the TESS observations. Section3 outlines our out all 3 σ outliers until no more are detected. After- methodology to find flare candidates, vet against false wards, we run the local view with the same criterion and alarms, and model flare data to estimate the flare energy. collect a list of all outliers. The results are presented in Section4, and a discussion We identify outliers as flare candidates if at least six and outlook are provided in Sections5 and6. minutes of flux data lie above a 3 σ threshold. These criteria are empirically selected to separate noise and 2. OBSERVATIONS actual flaring objects. Each target can (and often does) have multiple ‘flare candidate peaks’. Fig.1 illustrates The primary goal of TESS is to search for transit- the methodology using TIC 25118964 as an example. ing Earth-sized planets around nearby and bright stars. The flare candidate peaks are detected despite having Its four 10 cm optical cameras simultaneously observe small signal-to-noise and the presence of strong stellar a total field of 24◦ × 96◦. In its two year primary mis- rotation modulation. Finally we visually inspect all can- sion, TESS will measure light curves of over 200,000 pre- didates peaks to remove false positives (e.g. asteroids) selected stars with a 2 minute cadence and of millions and false alarms due to noise features. of stars with a 30 minute cadence. This study is based on the short (2 minute) cadence data collected by TESS in sectors 1 and 2, which were 3.2. Defining outburst epochs made publicly available with the first data release in De- Flares are often not isolated events. When the star cember 2018. These data contain 24,809 unique targets; is active, multiple consecutive flares may occur within a 8,815 in sector 1 only, 8,920 in sector 2 only, and 7,074 short time period. We denote this collection of flares as in both sectors. an ‘outburst’. The effects of multiple flares during an The 2 minute cadence data were extracted using the outburst can overlap, resulting in compound features in Science Processing Operations Center (SPOC) pipeline, the observed light curve. a descendant of the Kepler mission pipeline (Jenkins We split each light curve into sections, one for each 2002; Jenkins et al. 2010, 2016; Jenkins 2017; Stumpe outburst epoch. To define which flares are part of which et al. 2014; Smith et al. 2012). The SPOC pipeline is outburst epoch, we iterate through all flare events sorted operated at the NASA Ames Research Center. in time. For a given flare, if there is no other candidate peak one hour before or three hours after the event, 3. METHODS the outburst is labeled as containing only this single flare. When there are other flares following, the outburst 3.1. Searching for flares epoch gets expanded accordingly. As a result, outburst 4

1.08

0.97 0 5 10 15 20 25

1.05 Flux

0.97 3 4 5 6 7 8 9 10 1.02 1.02

1.02 1.05

1.00 0.99 0.98 1.00 4.2 4.4 5.4 5.6 6.0 6.2 6.7 6.9 Time (BJD - 2458325 d)

Figure 1. Demonstration of the detection pipeline on the example of the lightcurve of TIC 25118964. The x-axes show the time in Barycentric Julian Days (BJD) and y-axes show the normalized TESS PDC-SAP flux. Orange star symbols highlight the detected flare candidate peaks. The lower panels show zoomed views onto the regions marked with grey boxes. The flare candidate peaks are detected despite having small signal-to-noise and the presence of stellar rotation modulation. epochs span from 1 h before their earliest flare peak to outbursts sample, we consider an outburst recovered if 3 h after their latest flare peak. any injected flare is recovered. We find a clear lower limit for the flare amplitude, 3.3. Completeness of the flare detection pipeline below which flares cannot be recovered (Fig.2). The recovery rate for F, G, and K-dwarfs allows the detec- We evaluate the completeness of our flare detection tion of the smallest injected flares, with amplitudes at pipeline using two separate injection-recovery tests. For the milli-mag level. For early M dwarfs, we find the re- each Sector, we randomly select lightcurves of 200 F, covery rates decrease for flares with amplitudes of less G and K-dwarfs, 200 early M dwarfs and 200 late than a few percent. Flares on late M dwarfs must gen- M dwarfs. In the first test sample (individual flares), erally be least a few percent to be detectable. This de- we inject 10 flares into each lightcurve at random times, pendency on stellar types is a direct consequence of the with amplitudes randomly drawn from a log-normal stars’ brightness, as for fainter stars (such as typical late distribution between 0.01 and 1 in relative flux, and M dwarfs) the photometric scatter (white noise) in the full-widths at half maximum (FWHM) randomly drawn lightcurve increases, decreasing the signal-to-noise ra- from a uniform distribution between 2 minutes and 2 tio of any flares. The detection efficiency shows only a hours. In the second test sample (outbursts), we choose slight dependency on the injected flare FWHM, increas- the same criteria, but force all flares to occur subse- ing slightly for longer flares. Finally, as intrinsic stellar quently, mimicking extreme outburst regions. For this, variability and rotation modulation are detrended by we draw the first flare peak time randomly, and all sub- our flare detection pipeline, the minimum flare ampli- sequent peak times from a uniform distribution between tude seems not to be affected. 10 and 30 minutes after each other. While this is a much more extreme outburst than ever occurring in reality, it serves to test the robustness of our pipeline. 3.4. allesfitter We then run our flare detection pipeline (section 3.1) allesfitter (G¨unther & Daylan 2020, 2019) is a on all injected lightcurves, and record which flares are publicly available, user-friendly software package for recovered. In the individual flares sample, we consider modeling photometric and RV data. It is based on a a flare recovered if a detection is registered at less than generative model that can accommodate multiple ex- 10 minutes difference from the true peak time. In the oplanets, multi-star systems, star spots, and stellar Stellar Flares from the First Tess Data Release 5

100 100 a FGK b M0-M4 c M4-M10

10 1 80

10 2

Inj. flare ampl. (rel. flux) 60 10 3 25 50 75 100 25 50 75 100 25 50 75 100 Inj. flare FWHM (min.) Inj. flare FWHM (min.) Inj. flare FWHM (min.) 100 d e f 40 Recovery rate (%)

10 1

20 10 2 Inj. flare ampl.(rel. flux)

10 3 0 3.0 2.5 2.0 1.5 1.0 8 10 12 14 167000 6000 5000 4000 3000 RMS (rel. flux) TESS mag. Teff (K)

Figure 2. Recovery rates of flares from our injection tests. Upper panels show the recovery rate as function of the injected flares’ amplitudes and full-widths at half maximum (FWHM) for (a) F, G, and K-dwarfs, (b) early M dwarfs and (c) mid/late M dwarfs. Lower panels show the recovery rate as a function of the (d) residual root-mean squared error (RMS) after detrending, (e) TESS magnitude, and (f) stellar effective temperature, all averaged over all stellar types. The different detection efficiency between stellar types is a direct consequence of their magnitudes, as the photometric scatter in the lightcurve increases for fainter stars. flares. For this, it constructs an inference framework idence for it as quantified by Kass & Raftery(1995). that unites the versatile packages ellc (light curve and Hence, we define RV models; Maxted 2016), aflare (flare model; Daven- port et al. 2014), dynesty (static and dynamic nested ∆ log Z := log ZM1 − log ZM0 , (1) sampling Speagle 2020), emcee (MCMC sampling; p 2 2 σ(∆ log Z) := [σ(log ZM )] − [σ(log ZM )] (2) Foreman-Mackey et al. 2013) and celerite (GP mod- 1 0 els; Foreman-Mackey et al. 2017). allesfitter is pub- and demand licly available at github.com/MNGuenther/allesfitter, feedback and contributions are welcome. ∆ log Z > 5, (3) ∆ log Z > σ(∆ log Z). (4) 3.5. Modeling flares Using allesfitter, we perform nested sampling to Second, the extra flare must have a signal-to-noise ra- infer flare models with up to 0, 1, ..., N + 2 flares for tio (SNR) larger than 5. For lightcurves with signif- each outburst epoch, where N is the number of flare icant red noise, estimating the Bayesian evidence can candidates. We sequentially add extra flares to ensure be dominated by the volume of the prior, and be bi- overlapping flares are distinguished even if they were ased if flares are misused to account for noise structures missed by the detection pipeline and visual inspection. (i.e. the nested sampler could find numerous solutions We start with fitting models of ‘only noise’ and one flare. for placing small flares). This risk can be mitigated by If the model with one flare passes, we test a model with introducing an SNR criterion. two flares, and so on, up to a maximum of N + 2. We apply all fits to the detrended PDC-SAP flux. To The model selection is performed using two comple- additionally detrend any systematic noise or stellar vari- mentary criteria, which both must be fulfilled. First, ability features without affecting flares, we fit a Gaus- adding an extra flare must increase the logarithm of the sian Process with a Matern 3/2 kernel in parallel with Bayesian evidence, log Z, by at least 5. Given a null all flare models. For each model, we first run two short model M0, the alternative (more complex) model M1 is Markov Chain Monte Carlo (MCMC) chains to explore only selected if there is sufficient relative Bayesian ev- the likelihood space and constrain the search space for 6 the nested sampler. This also allows us to label the se- bles1 and2. For these, it was not possible to determine quence of flares, mitigating the risk of label-swapping the true origin of the flaring, as the stars fall within during the nested sampling run. the same pixel and potentially are part of a binary sys- Fig.3 illustrates our methodology on four example tem. We also noted a single instance for TIC 229147922 targets. For TIC 139804406, we detected two flare can- at 2458363.726719 BJDTDB, in which one of the flares didates in the shown outburst epoch. Comparing all originated from a neighboring star. plausible models, we can confirm this scenario. For TIC 129646813, the candidate list initially consisted 3.7. Measuring the flare energy of a single flare. However, our model fit and com- We calculate the flare energy from the stellar luminos- parison show that there are, in fact, two flares. For ity and the best fitting flare profile, following Shibayama TIC 144217628, there is no evidence that the short fea- et al.(2013). The quiescent stellar luminosity L? is re- ture towards the end of the outburst is a flare; it is best trieved from the TESS Input Catalog (TIC) version 8 if explained as a noise feature. Finally, for TIC 52875048, available. If the TIC lists only the effective temperature the candidate peak was identified as a noise feature. Teff and stellar radius R?, we compute L? from these Note that the Bayesian evidence helps determine the parameters. If only the Teff is available, we estimate degree to which a model is supported by the data and R? using the updated values from Pecaut & Mamajek does not simply indicate its likelihood. For example, for 2 (2013) to then calculate L?. TIC 129646813 (Fig.3B) we have ∆ log Z1→2 flares = 25. We model the flare luminosity as black body radia- This is not to be confused with a statement such as “two tion with an effective temperature of 9000 ± 500 K, as a flares are 25 times more likely than one flare”. Instead, it conservative lower limit consistent with previous stud- implies that there is very strong evidence for the model ies (e.g. Hawley & Pettersen 1991; Kowalski et al. 2013; with two flares. Shibayama et al. 2013; Davenport 2016; Jackman et al. 2018, 2019; Howard et al. 2018; Chang et al. 2018). As 3.6. Vetting for asteroids and blends discussed e.g. by Shibayama et al.(2013), the luminosi- We inspected a sample of around 100 strongly flaring ties of the star and the flare in the observing bandpass 0 0 stars for misclassified flares caused by Solar System Ob- (L? and Lflare, respectively) are given as jects (SSO). This effect was discussed by P´alet al.(2018) Z for the TESS mission. Additional catalog inputs within 0 2 L? = πR? RλBλ(Teff )dλ (5) the same sector as the high amplitude peaks were in- Z cluded in our sample, making it a total of circa 600 flare 0 Lflare(t) = Aflare(t) RλBλ(Tflare)dλ. (6) events. We used SkyBot (Berthier et al. 2006, 2016) to search for known SSOs crossing the target stars at the time of flare peaks. In only a few cases we found SSOs Here, Rλ is the TESS response function, which is the product of the filter transmission and the detector quan- within a range of 100”. In these instances, we carried 3 out additional searches by slightly changing the input- tum efficiency . Note that the normalization of Rλ is time, in order to see whether the same objects would be irrelevant here, as it cancels out in the calculations be- re-detected. This was not the case in any of the com- low. Bλ(Teff ) and Bλ(Tflare) are the Planck functions putations and we therefore discards these SSOs as po- evaluated for the star’s effective temperature and the tential contaminants. In a single occurrence, associated flare temperature. Last, Aflare(t) is the area of the flare. Note that this assumes that the flare temperature is con- with TIC 160074646 at 2458379.648 BJDTDB, we notice a nearby presence (21.6”) of the outer main belt SSO stant throughout. 2010 KM11 (519036). However, the light curve profile Since the normalized lightcurve gives us the relative flare amplitude (∆F/F )(t) = L0 (t)/L0 , we can solve of this event match that of a flare profile with a fast rise flare ? and an exponential decay. Additionally, we created a these equations for Aflare: series of pixel-level animations which we then visually R R B (T )dλ inspected similar to Szab´oet al.(2015), to prevent the A (t) = (∆F/F )(t)πR2 λ λ eff . (7) flare ? R R B (T )dλ appearance of currently unknown SSOs in the sample. λ λ flare No SSO-crossings were detected, including 2010 KM11 most likely due to it visual magnitude of 20.7. 2 http://www.pas.rochester.edu/∼emamajek/EEM dwarf In the process, we also checked for contamination by UBVIJHK colors Teff.txt, online 28 Dec 2018 blended flaring stars. A low number of potential du- 3 heasarc.gsfc..gov/docs/tess/data/ plicate stars is marked in the column duplicates in Ta- tess-response-function-v1.0.csv, online 28 Dec 2018 Stellar Flares from the First Tess Data Release 7

1.150 Best model:

Best model: Z

Z 50 2 flares g

2 flares g 25 o l o 1.20 1.125 l 0 0 0 2 100 0 2 1.100 20

1.15 SNR SNR 1.075 0 0 0 1 2 3 0 1 2 3 Nr. of flares Nr. of flares 1.10 1.050 Relative Flux Relative Flux 1.025 1.05

1.000

1.00 0.975

0.025 0.02

0.000 0.00 Residuals Residuals 0.025 0.02

0.000 0.025 0.050 0.075 0.100 0.125 0.150 0.175 1.850 1.875 1.900 1.925 1.950 1.975 2.000 2.025 Time (BJD) +2.458328e6 Time (BJD) +2.45834e6 (a) TIC 139804406 (b) TIC 129646813

1.20 Best model: 1.0

Best model: 50 Z Z

1.030 pure noise g

1 flare g

o 0.5 l o

l 1.15 1.025 0.0 0 0 1 0 2 2 10 1.10 1.020 SNR SNR 0 0 0 1 2 0 1 1.015 Nr. of flares 1.05 Nr. of flares

1.010 Relative Flux Relative Flux 1.00 1.005

1.000 0.95

0.995 0.90 0.05 0.005 0.00 0.000 Residuals Residuals 0.005 0.05

0.650 0.675 0.700 0.725 0.750 0.775 0.800 2.950 2.975 3.000 3.025 3.050 3.075 3.100 Time (BJD) +2.458325e6 Time (BJD) +2.45835e6 (c) TIC 144217628 (d) TIC 152875048 Figure 3. Model fits of candidate outbursts of four example stars. In the main figures, the x-axis shows the time in Barycentric Julian Date (BJD) and the y-axis shows the TESS PDC-SAP flux. Red curves show 20 posterior samples generated by the best model, while orange curves show their corresponding baselines drawn from a Gaussian Process with a Matern 3/2 kernel. The curves for the less suitable models are not shown. The inlets show the gain in Bayesian evidence, ∆ log Z, by adding an additional flare (upper inlay), and the SNR of the additional flare (see Section 3.5). Initially, N flare candidates were detected. Consequently, scenarios from 0 to at most N + 2 flares are fitted as long as ∆ log Z > 5 and SNR> 5 for an additional flare. This allows for a quantitative model comparison, confirming the suitable number of flares (see Section 3.5). (a) TIC 139804406: two candidate peaks were initially detected, and then confirmed using our model fit; introducing a third flare does not lead to any gain in Bayesian evidence. (b) TIC 129646813: only one candidate peak was detected, but our model comparison confirmed two flares; adding a third flare is not favored. (c) TIC 144217628: the Bayesian evidence rejects the scenario where the smaller peak is a flare; instead, it favours the peak being within the limits of a noise feature. (d) TIC 152875048: one candidate peak was detected, but the fit favors the pure noise model and rejects a flare scenario. 8

This leads to the bolometric flare luminosity Lflare: length of one TESS sector, and half have periods longer than 90 days. 4 Lflare = σSBTflareAflare. (8) KELT has performed high-cadence (10–30 min), time- series photometric observations for more than four mil- Finally, we arrive at the expression for the bolometric lion sources since 2007. KELT observations have sur- energy of the flare, given as: veyed more than 70% of the celestial sphere, reaching a Z limiting magnitude of about V = 13, and with a base- E = L (t)dt. (9) flare flare line of nine years using KELT North and five years using KELT South. Oelkers et al.(2018) provide a catalog of We estimate the error on the flare energy by propagating 52, 741 objects showing significant photometric fluctua- all uncertainties on the stellar properties from the TESS tions likely caused by stellar variability, as determined input catalog and the estimated uncertainty on the flare via the Welch-Stetson J and L statistics (Stetson 1996). temperature. Additionally, this catalog includes 62, 229 objects iden- 3.8. Identifying stellar rotation periods tified with likely stellar rotation periods as determined by a Lomb-Scargle periodogram analysis (Lomb 1976; The rate and energy of stellar flares depend on the sur- Scargle 1982). The detected variability ranges in root face magnetic activity of the star, thus flares are thought mean square (RMS) amplitude from 3 mmag to 2.3 mag, to result from strong dynamo activity (e.g. Parker 1979). and the detected periods range from 0.1 days to over Accordingly, the flaring activity should depend on the 2000 days. Oelkers et al.(2018) also provides variability effective temperature and rotation periods (elaborated upper limits for all other four million sources observed in Section 4.3). by KELT. These upper limits typically have 1σ sensi- We derive the stellar rotation period of all flaring tivity on 30 min timescales down to 5 mmag at V = 8, and non-flaring stars from the TESS lightcurves them- and 43 mmag at V = 13. selves, and from a study conducted with the Kilodegree Extremely Little Telescope (KELT) by Oelkers et al. 4. RESULTS (2018). For TESS data, we measure the rotation pe- 4.1. TESS catalog of stellar flares riod (Prot) using Fast Fourier Transform (FFT) compu- tations (see Zhan et al. 2019 for details). We choose All detected, vetted and modeled flare events are sum- the FFT method due to its robustness and computa- marized in Table1 per flare event, and in Table2 per tional speed. With the data being uniformly sampled, star. In Table1, each star can be listed multiple times, there was no advantage in using Lomb-Scargle trans- with one row per flare. For each flare event, the columns foms (Lomb 1976; Scargle 1982) or other methods. We show the following: the TIC ID; the TESS sector; the detect and analyze the FFT frequency peaks to find the outburst number; the flare number; the posterior me- primary rotation period. This is sensitive to the Nyquist dians for the peak time (tpeak), amplitude (Amp.) and limit at 4 minutes and a conservative upper limit of 5 full-width at half-maximum (FWHM ); the bolometric days (to minimize false alarms due to momentum dumps energy of the flare (Ebol.); and the possible mass of a and the 13-day space craft orbits). The reported periods coronal mass ejection following the flare (MCME; see Sec- and uncertainties are then determined using a Stellingw- tion 5.2). The machine-readable version is available in erf transform (Stellingwerf 1978). Finally, the raw light the online journal, and additionally contains lower and curve is phase folded onto the rotation period and ex- upper limits. For ease-of-use, it also includes a copy of amined by eye, ensuring that the signals are not caused the per-star columns shown in Table2 (see below). by instrumental systematics or astrophysical false posi- In Table2, each star only has one row; columns show tives. the maximum/mean values of the flaring and the stel- The 28 day baseline per TESS sector does not fa- lar parameters as follows: the TIC ID; the number of vor the identification of longer rotation periods, which TESS sectors (Nsec.); the number of identified outbursts are common among M dwarf stars. Studies measuring (Nout.); the number of identified flares (Nfla.); the max- the rotation periods of nearby low-mass stars (Newton imum and mean flare amplitude (Amp. max. and Amp. et al. 2016, 2018) have found a population of fast ro- mean); the maximum and mean flare full-width at half- tators (Prot < 10 d) and a population of slow rotators maximum (FWHM max. and FWHM mean); the max- (Prot > 70 d), and a dearth of objects in between. Con- imum bolometric energy of the star’s flares (Ebol. max. sidering only their highest quality lightcurves, Newton and Ebol. mean); the maximum and mean mass of coro- et al.(2018) found that two-thirds of mid-to-late M nal mass ejections (MCME max. and MCME mean); the dwarfs have rotation periods longer than 28 days, the TESS magnitude (TESS mag); the stellar effective tem- Stellar Flares from the First Tess Data Release 9 perature (Teff ); the stellar radius (R?); the logarithm white light. Yet, favorable for these stars is their lower of the surface gravity (log g); the rotation period ex- photometric noise in the TESS band. tracted from TESS and KELT data (TESS rot. and KELT rot.; see Section 3.8). an indication of which flaring, 1228 all targets, 24816 star flares enough to potentially trigger prebiotic chem- Spectral type istry on its exoplanets (preb. chem.; see Section 5.1); F0 F5 G0 K0 K5 M0 M5 M8 and an indication of whether a star’s flaring could lead 3 10 to ozone depletion on its exoplanets for a conservative (ozone depl. cons.) or permissive (ozone depl. perm.) 1 threshold (see Section 5.1); the TIC IDs of any potential 10 blends (duplicates). For readability, several columns are Nr. of stars hidden in the printed version of Table2. The machine- readable version is available in the online journal, and 40 contains all described columns, among all lower and up- per limits values. 20

4.2. TESS explores a large sample of flares on bright 0 early and late M dwarfs Fraction of stars (%) 8000 7000 6000 5000 4000 3000 Teff (K) We identify 1228 flaring stars among a total of 24,809 targets with short-cadence TESS observations in sectors Figure 4. Histograms of the number (upper panel) and 1 and 2 (Fig.4). This includes 531 flaring early M dwarfs fraction (lower panel) of flaring stars (blue) compared with (M0–M4; 3905K-3200K) and 142 flaring late M dwarfs the total number of stars (grey) in the TESS short-cadence observations of sector 1 and 2; shown as a function of the (M4–M10; 3200-2285K). Fig.5 shows the subsample of stellar effective temperature Teff . The top axis indicates stel- M dwarfs categorized by stellar type. Flares are most lar types following the classification by Pecaut & Mamajek commonly detected on mid M dwarfs, which is partly (2013). M dwarfs dominate the sample of flaring stars, while influenced by the TESS target selection and signal-to- F, G and K stars rarely have detectable flares. We note that noise constraints for the flare detection. M dwarfs of for later M dwarfs, the sample size is smaller (or zero) due to type M4–M6 show the highest fraction of flaring stars. the TESS target selection. Additionally, the flare detection Mid to late M dwarfs are the most common flare stars, is limited by signal-to-noise constraints (see section 3.3). with more than 40% of these showing observable flares (Fig.4B). This occurrence rate is significantly lower for TESS is designed to survey bright dwarf stars, ex- early M dwarfs, as only ∼ 10% have observable flares. panding the number of M dwarf flares detected in the Note that this might directly relate to the convection solar neighborhood (Fig.6). Our sample includes 189 limit, where full convection of M dwarfs is supposed to early M dwarfs and 6 late M dwarfs brighter than 12th start around spectral type M4 (e.g. Stassun et al. 2011). TESS magnitude. These stars can be prime targets for Hotter stars of type K and G seem to rarely host flaring atmospheric characterization and radial velocity follow- events large enough to be detected (only ∼ 5% of these up if they are found to host transiting planets. Flares stars in our sample). We find flares only on 85 F, 114 can provide constraints on the existence of an atmo- G, and 184 K stars, all of which have low amplitudes. sphere or habitability constraints (see Section5). This confirms previous observations that M dwarfs flare With 24 more sectors to be observed in the next two more frequently and stronger than F, G and K stars. years, and flare detection in the full-frame images, the However, we note the interplay of two detection biases. expected TESS yield of flaring M dwarfs is on the or- Flares on M dwarfs have a strong white light contribu- der of 104. In contrast, the Kepler mission focused on a tion and a stronger contrast against their red spectra different sample, namely F, G, and K stars (see Fig.6). and low quiescent luminosity. This enhances M dwarfs The Kepler flare catalog (Davenport 2016) collects a to- observable amplitudes in the TESS band, favoring the tal of 4041 objects. Of these, 4036 can be crossmatched detection of less energetic flares. On the other hand, with the TIC. The crossmatched list contains 110 flaring M dwarfs also have higher photometric noise, which in early M dwarfs, and 4 flaring late M dwarf observed with turn decreases the SNR of any flares (see Section 3.3). Kepler. Additional studies with Kepler (e.g. Van Doors- For F, G, and K dwarfs, on the other hand, only the selaere et al. 2017), K2 (Stelzer et al. 2016), MEarth most energetic flares cause high enough amplitudes to (Mondrik et al. 2019) and others can provided an ex- be detected in contrast to their quiescent brightness in panded view into the M dwarf regime, yet still for lim- 10

TESS Kepler 1000 flaring all targets 500 M4

Nr. of stars M3 M5 M0 M1 M2 M6 M7 M8 0 F0 F5 G0 K0 K5 M0 M5 M8 60 6 M5 M6 40 M4 no samples 8 20 M0 M2 M3 M1 0 10

Fraction of stars (%) 4000 3750 3500 3250 3000 2750 2500 Teff (K) TESS-mag Figure 5. Variation of Figure4 focused on M dwarfs, with 12 bins matching the stellar types following the classification by Pecaut & Mamajek(2013). Mid M dwarfs of type M4– M6 constitute both the highest number (upper panel) and 14 highest fraction (lower panel) of flaring stars, with up to 30% of these having flares. Earlier M dwarfs seem to flare 16 less, showing a significantly lower fraction of 5–10%. Later 8000 7000 6000 5000 4000 3000 M dwarfs were not observed in a high enough sample size. We Teff (K) note that especially for M dwarfs later than M4 the sample size is lower due to the TESS target selection (favoring bright Figure 6. TESS explores bright early to late M dwarfs, stars), and the flare detection is limited by signal-to-noise expanding the sample size for flare studies in this parameter constraints (see section 3.3). space. Shown are the effective temperature Teff versus TESS magnitude of flaring stars in the TESS sample (blue) and the Kepler flare catalog (grey; Davenport 2016). The top axis ited sample sizes. Fig.6 highlights how TESS enables to indicates stellar types following the classification by Pecaut explore this parameter space for flare studies in an un- & Mamajek(2013). The first two sectors already include precedented manner. Moreover, TESS also detects small 531 early M dwarfs and 142 late M dwarfs. Out of these, 189 flares on F, G, and K dwarfs brighter than 12th TESS and 6 are brighter than 12th TESS magnitude. In addition, magnitude, filling in another parameter space which was TESS detects small flares on F, G, and K dwarfs that are less explored by the Kepler mission. brighter than there average Kepler targets.

4.3. Flares on fast rotators type stars to M dwarfs, and across the fully-convective transition at the lowest stellar masses (Newton et al. Stellar rotation is suggested to be linked to the flaring 2017; Wright et al. 2018). of a star according to the dynamo theory of magnetic We here define two categories: ‘photometric rotator’ field generation (e.g. Moffatt 1978; Parker 1979). The and ‘other’. We use the phrase ‘photometric rotator’ interaction between rotation and convection can cause synonymous to ‘a star with a photometrically measured high magnetic activity (e.g. Browning 2008), whose en- rotation period by TESS or KELT’. Consequently, rotat- ergy is then released through flares. Two groups of ing stars without detected modulation are not included rotators are observed among low-mass stars: fast ro- in this definition. Our rotation analysis based on TESS tators (P < 10 d) and slow rotators (P > 70 d) rot rot data extends to fast rotators with P < 5 d. The KELT (e.g. Newton et al. 2016, 2018). Fast rotating M dwarfs rot data covers a limited sample of stars with longer rotation (P < 10 d) exhibit a saturated level of activity, while rot periods. Hence, the ‘other’ category might still contain for slower rotators, the ratio of X-ray, Hα, and Ca H&K many fast rotators with P > 5 d, for which we do not flux to bolometric luminosity declines rapidly (West rot have measured rotation periods. et al. 2015; Astudillo-Defru et al. 2017; Newton et al. We find that about 60% of fast rotating early and 2017; Wright et al. 2018). The relationship between ac- late M dwarfs in the TESS sample show flaring that is tivity and rotation is typically parameterized in terms detectable (Fig.7). In contrast, only 10% of all ‘other’ of the Rossby number, which seeks to remove the mass M dwarfs have flares. Among F, G and K stars, less than dependence by dividing the rotation period by the con- 5% of ‘photometric rotators’ and almost no ‘others’ have vective overturn timescale (Noyes et al. 1984). This be- flares. This solidifies past findings from, for example, the havior is consistent across the main sequence from solar- Stellar Flares from the First Tess Data Release 11

Kepler (Davenport 2016; Van Doorsselaere et al. 2017) riod (e.g. Maehara et al. 2012). For our sample, however, and MEarth (Mondrik et al. 2019) surveys. the available rotation period information is too sparse to Without fast rotation, there might not be enough en- study this link for rotation periods P > 1 d. ergy stored in the magnetic field lines to trigger frequent strong flaring. However, in fully-convective stars the dy- namo mechanism is not well-understood. For stars with flaring not flaring a radiative zone and a convective zone, the interface 3 10 between these two zones is what is believed to power

the dynamo (Durney 1993). Lacking this interface, fully Nr. of stars convective M dwarfs (typically with spectral types M4 100 and later) might generate their magnetic fields through other means (see e.g. Stassun et al. 2011). Neverthe- less, the similar behavior in activity and rotation period 50 (measured via Hα Newton et al. 2017 and X-ray Wright et al. 2018 data) suggests that the manifestations of the 0 magnetic fields in partially and fully convective objects Fraction of stars (%) are similar. FGK M0-M4 FGK other other M4-M10 For our sample, it is difficult to disentangle whether M0-M4 M4-M10 other the slow rotation impacts both the flare rate and flare phot. rotator phot. rotator phot. rotator amplitude, or just one of them. First, if the flare rate Figure 7. Histograms of the number (upper panel) and of slow rotators is indeed lower, the 28 days observation fraction (lower panel) of flaring stars (blue) compared to ones span with TESS is not sufficient to detect their flaring. for which we do not detect flares (grey). Bins separate the Second, our detection algorithm is not focused on the sample into F, G and K stars, early M dwarfs of type M0– detection of the smallest flares; slow rotators could still M4, and late M dwarfs of type M4–M10; each of these are flare frequently, yet at low amplitudes that are not de- further grouped by whether or not a rotation period could be tected. photometrically measured. Flares are detected for ∼ 60% of Moreover, we find that the maximum flare amplitude fast rotating early and late M dwarfs in the TESS sample. In and FWHM increase for stars with more frequent flaring contrast, only ∼ 10% of all other M dwarfs show detectable flaring. (Fig.8). This seems to be independent of whether a star is a photometric rotator or not. Furthermore, stars with higher flare rates show similar mean values of their flares amplitudes and durations. 4.4. 99 stars show flares with flux increases of multiple We do not find a correlation between stars’ rotation magnitudes periods and their flare amplitudes (Fig.9). Large flares Chang et al.(2018) conducted a study of flaring seem to be as likely to occur on photometric fast ro- M dwarf stars in the Kepler sample. They found eight tators as on the other stars in our sample. This is in flares that increase the stellar brightness at peak by a agreement with previous findings (e.g. Maehara et al. factor of two or more. The authors define these events as 2012), and can be explained if the magnetic field energy ‘hyperflares’ (we do not follow this nomenclature here). is stored near a star spot. It was shown for different The strongest flare they detected has a peak luminosity star types, that in this scenario the rotation period does 5 times the quiescent flux. not influence the maximum energy (e.g. Rodono et al. The TESS sample profits from the mission’s observing 1986). strategy, allowing us to identify 99 flaring stars with Our results suggest a link between the flare frequency 140 individual flares that fulfill the Chang et al.(2018) and the rotation period for the fastest photometric ro- criterion of at least doubling the stellar brightness. A list tators (Fig.9). Stars with P < 0.3 d reach only half the of these targets can be created from Table1 by sorting maximum flare frequency as those with P > 0.3 d. How- by the amplitude column. ever, this could also be sampling issue; there are fewer photometric rotators with P < 0.3 d and the distribu- 4.5. The largest superflare and the most energetic tion of the flare frequency, peaking at ∼ 0.1 flares per superflare in the TESS sample day, shows a long tail until ∼ 0.5 flares per day. 260506296 (2MASS J06270005-5622041), an M4.5V For P > 1 d, the frequency of superflares (flares with dwarf star (3189 K) with TESS-mag 13.7, shows the energies > 1033 erg) was reported to decrease with pe- largest flare amplitude in our sample (Fig. 10). Over 12

phot. rotators other the course of 1 h, the star increases its optical bright- ness by a factor of 16.1, releasing a bolometric energy of 1034.7 erg. The superflare is preceded by a series of smaller flares. We inspect the individual target pixel

0 files for this outburst using the lightkurve (Vin´ıcius 10 et al. 2018) module. This confirms that the flare is in- deed on TIC 260506296.

2 In contrast, the most energetic superflare in our sam- 10 Ampl. (rel. flux) ple reaches 1036.9 erg. It is found on 332487879 (2MASS max. mean J23211550-2659121), a G-type giant star (5192 K, 9.3 R ) with TESS-mag 4.9. Such energetic flares

0 on giant stars are common (e.g. Balona et al. 2015; 10 Van Doorsselaere et al. 2017), which is surprising, as the evolution off the main sequence should decrease 1 10 their magnetic field (Simon & Drake 1989). Possible explanations could be that the giant’s surface magnetic FWHM (h) 2 field is conserved, or that the flare actually originates 10 max. mean from an unresolved dwarf star companion. Notably, 0.0 0.5 0.0 0.5 332487879’s enormous release of energy only leads to Flare rate (d 1) Flare rate (d 1) a flare amplitude of 1.01. This is a direct consequence of the stellar type, namely effective temperature and Figure 8. The maximum flare amplitude (upper left), radius, following Eqs.5–9. In comparison, the energy mean flare amplitude (upper right), maximum FWHM output for the M dwarf TIC 260506296 is significantly (lower left), and mean FWHM (lower right) of each star in lower, despite its much larger amplitude; this is due to dependency of the flare rate per day. The samples are further the M dwarf’s small radius and lower effective temper- separated into stars which have photometrically measured ature. rotation periods (blue) and ones that do not (grey). Stars with higher flare rates have a significantly higher maximum flare amplitude and maximum FWHM. There is no signifi- 4.6. Flare frequency distributions cant difference between photometric rotators and other stars. We study the flare energies and flare rates as a func- tion of the stellar type and rotation period (Fig. 11). This is commonly denoted as the flare frequency dis- tribution (FFD; e.g. Gershberg 1972; Lacy et al. 1976; Hawley et al. 2014). The FFD shows the cumulative )

1 rate of flares per day, i.e. how often a flare of a certain 0 0

10 d 10 ( energy or higher is detected.

e t

a Using the TIC information, we separate F, G, and K r

e 2 r 1 stars from early M dwarfs (M0–M4; 3905K–3200K) and a 10 l 10

F late M dwarfs (M4–M10; 3200–2285K). These are fur-

Max. Ampl. (rel. flux) ther divided into stars with photometrically measured 1 0 1 1 0 1 10 10 10 10 10 10 rotation modulation (‘phot. rotator’) and ones without Prot (d) Prot (d) (‘other’; see Section 4.3). Each star’s FFD is fit with a line in the log-log space to extrapolate the trend, follow- Figure 9. The maximum flare amplitude (left panel) and number of flares per day (right panel) are studied as ing log10(flare rate) = α log10(Ebol.) + β. The best fit a function of the rotation periods Prot measured by TESS parameters α and β are included in Table1. and KELT. There is no significant correlation between pho- We find no significant difference between stars with tometric fast rotators and the amplitude of their superflares, and without detected rotation modulation. This is albeit we note that our sample is mostly limited to Prot < 1. somewhat surprising given that fast rotating M dwarfs There are signs that the fastest rotators (Prot < 0.3 d) do not are suggested to flare more frequently. Moreover, we flare as frequently as the rotators with Prot > 0.3 d. How- find that the average flare energy per star depends on ever, this remains to be re-examined with a larger sample size. the effective temperature and radius of the star, which is a direct consequence of Eqs.5–9. For F, G and K stars, this effect in addition with their lower contrast (see Sec- Stellar Flares from the First Tess Data Release 13

15 planet distance, a, to keep the planet at a distance from its host star so that the flux is the same as 1 Earth flux (as defined by Kane et al. 2016). The flare frequency 10 needed to drive the prebiotic chemistry, ν, is a function of the flare’s U-band energy EU , the stellar radius, R∗, Flux and the stellar temperature, T∗. Then, ν can be derived 5 following: a b c ! !2 !4 1 1034 erg R T ν ≥ 25.5 day−1 ∗ ∗ . (10) 0.50 0.55 0.60 0.65 0.70 0.75 E R T Time (BJD) +2.458373e6 U To apply this equation to the bolometric energies we a b c 995 derived, we first calculate the flares’ U-band energy from the bolometric energy. This uses the U-band spectral response function and assumes a 9000 K blackbody for 990 the flare. We find that 7.6% of the flare’s bolometric

CCD y (pixel) energy falls into the U-band, leading to EU ≈ 7.6%Ebol. 1460 1465 1460 1465 1460 1465 Flare frequencies satisfying Eq. 10 could drive the syn- CCD x (pixel) CCD x (pixel) CCD x (pixel) thesis of ribonucleotides, the building blocks of RNA, from cyanide and bisulfite in liquid Figure 10. The largest flare in our sample is on 260506296, an M4.5V dwarf star (3189 K) with TESS-mag 13.7. Upper water. These flare rates establish a necessary condi- panel: TESS PDC-SAP lightcurve. Lower panel: TESS tar- tion for the origin of life’s building blocks according to get pixel images at the times before the flare (a), during the the scenario proposed by Xu et al.(2018). Having the flare peak (b), and after the flare (c), as indicated in the necessary UV light from stellar activity is not a suffi- upper panel. Our best fit model shows that during the peak cient condition for this scenario. Liquid water, hydrogen the star increases its brightness by a factor of 16.1. This su- cyanide, bisulfite, and other feedstock are also perflare is preceded by a series of smaller flares (inlay). We necessary. inspect the target pixel files for this outburst, and confirm that the flare originates from the target. Fig. 11 displays the flare rates of the stars in our sam- ple, compared to this inequality. We find that 14 stars have enough flaring (at their current age) to deliver the tion 4.2), is why they generally show higher flare energies UV energy at a rate that could trigger prebiotic chem- than early M dwarfs and late M dwarfs. istry on a potential exoplanet. These include 9 early 5. DISCUSSION M dwarfs and 2 late M dwarfs. Out of these, 0 early M dwarfs are brighter than 12th TESS magnitude. A 5.1. Which flares deliver necessary energy to trigger list of the targets whose flaring have this potential can prebiotic chemistry? be created from Table1 by sorting by the ‘prebiotic We combine the measured flare energies and rates with chemistry’ column. the laboratory study conducted by Rimmer et al.(2018), It is important to note that most of the stars in our who delineate ‘ zones’. These are zones out- sample are likely old (albeit we currently have limited side of which a specific prebiotic chemical scenario can- knowledge of their exact ages). The onset of prebiotic not succeed around main sequence stars. This consid- chemistry likely occurred during earlier stages of exo- ers ribonucleotides, required for ribonucleic acid (RNA) solar systems, when the stars were younger. Given that synthesis, as a starting point for prebiotic chemistry (e.g. young M dwarfs were more active (West et al. 2008) and Sutherland 2015). The authors consider the competi- thus assuming that their flaring was more frequent and tion between reactions that produce ribonucleotides in energetic in the past, our results provide a lower limit. the presence of UV light (200-280 nm), with other bi- If a star’s FFD passes this cutoff now, it will also have molecular reactions that produce inert adducts with no done so in a younger stage. prebiotic interest. They calculate the UV light from a flare multiplied by the flare frequency, and compare it 5.2. The impact of coronal mass ejections to the lifetime of prebiotic intermediates required for ri- A coronal mass ejection (CME) often follows a stellar bonucleotide formation (e.g. Xu et al. 2018). Rimmer flare, and can have a substantial impact on potential et al.(2018) develop equations for a planet at a fixed exoplanets. A CME is a large release of plasma and the distance from its host star. Here, we adjust the star- connected magnetic field from the stellar corona. 14

log10 Eflare; U (erg) log10 Eflare; U (erg) log10 Eflare; U (erg) 1 30 32 34 36 30 32 34 36 30 32 34 36 10 | | | | | | | | | | | | ) 1 d

( 0

. 10 l o b ; e r a l

f 1

E 10

e t a

r 2

e 10 r a l F

3 FGK, phot. rotator M0-M4, phot. rotator M4-M10, phot. rotator 10

1 10 | | | | | | | | | | | | ) 1 d

( 0

. 10 l o b ; e r a l

f 1

E 10

e t a

r 2

e 10 r a l F

3 FGK, other M0-M4, other M4-M10, other 10 32 34 36 38 32 34 36 38 32 34 36 38 log10 Eflare; bol. (erg) log10 Eflare; bol. (erg) log10 Eflare; bol. (erg)

Figure 11. Flare frequency distributions (FFD) in the context of prebiotic chemistry (green area) and ozone sterilization (red areas). The x-axis shows the flare energy, as bolometric energy Eflare;bol. on the lower ticks, and U-band energy Eflare;U on the upper ticks. The y-axis shows the cumulative rate of flares per day, i.e. how often a flare with at least a certain energy appears. Different panels show F, G and K stars (orange), early M dwarfs (red) and late M dwarfs (black), separated into photometric rotators (filled circle) and others (unfilled circle). Solid lines are linear fits to the double-logarithmic FFD of each star, extrapolating into regimes that could not be observed. The green area denotes the minimum flare rate and energy required to trigger prebiotic chemistry on a potential exoplanet (expanded from Rimmer et al. 2018, see Section 5.1). The different green shadings show each threshold for each star, which depends on the stellar radius and effective temperature (Eq. 10). In the red shaded region derived from Tilley et al.(2019), intense flares are frequent enough that ozone layers cannot survive, and planet surfaces may be sterile (see Section 5.5). We mark two ozone sterilization regions: a permissive threshold for flare rates ≥ 0.1 per day (lighter red area), and a conservative threshold for flare rates ≥ 0.4 per day (darker red area). 14 stars, including 9 early M dwarfs and 2 late M dwarfs, in the TESS sample fulfill the criteria of prebiotic chemistry. On the other hand, potential exoplanets around 100 stars, including only 15 M dwarfs, might suffer from ozone depletion. To estimate the impact of CMEs, we apply the em- bolometrically (Section 3.7). To apply it to our data, we pirical relationship between flare energy and CME mass estimate the X-ray flare energy as 1% of the bolometric found by Aarnio et al.(2012). This has been calibrated energy, which is based on findings for the strongest solar against observations of flares and CMEs from the Sun, flares. These have X-ray fluxes of about 10−4 W m−2 as well as a sample of well-studied, very active pre-main- (Aarnio et al. 2011), corresponding to X-ray energies of sequence stars. The callibration encompasses flare ener- 1029 erg (Aarnio et al. 2012) and to bolometric energies gies from ∼1028 erg up to ∼1038 erg—comparable to the of 1031 erg (Maehara et al. 2015). Hence, for the same strongest flares included in this study—and which cor- flare strength, the energy emitted bolometrically is typ- respond to CME masses of up to ∼1022 g (Aarnio et al. ically larger by a factor of 100 (see also Osten & Wolk 2011, 2012). 2015). This empirical relation, however, is applicable to the Thus we adapt the following equation from Aarnio flare energy that is emitted in X-rays, whereas in this et al.(2012) for the CME mass, MCME, in terms of the work we determined the flare energy that is emitted Stellar Flares from the First Tess Data Release 15 bolometric flare energy: planet surface. Tilley et al.(2019) estimate that, conser- vatively, 0.083 of M dwarf flares will strike a habitable- E (0.63±0.04) M = (2.7 ± 1.2) × flare;bol (11) zone exoplanet, meaning that terrestrial-type ozone lay- CME 100 ers should not persist on stars with flares of ≥ 1034 erg All values are in cgs units. at a frequency of ≥ 0.4 day−1. A more permissive limit From Eq. 11, we find a median CME mass of 1020.0 g, could be drawn at ≥ 0.1 day−1. These parameter spaces ranging from a minimum of 1018.5 g to 1022.3 g for the are marked in red in Figure 11. most massive CME. This could have accompanied the Few of the stars we consider in this study fall into this strongest flare in our sample, which was found on the parameter space. For the permissive limit of 0.1 day−1, giant star 332487879 (5192 K, 9.3 R ). All estimates potential exoplanets around 100 stars, including 15 are also given in Tables1 and2. M dwarfs, might suffer from ozone depletion. For the CME events mainly happen along a given direction conservative limit of 0.4 day−1, 22 stars could be af- and, as such, are non-isotrop events. Their impact de- fected. However, it is possible that more frequent but pends among other factors on the semi-major axis and smaller flares and/or less frequent but more energetic inclination of the exoplanet’s orbit (Kay et al. 2016). As flares may also remove ozone layers. Further modelling exoplanets’ habitable zones around M dwarfs are close work is required to rule on this possibility. We note that to their star, the probability of being hit by a CME is even in the absence of an ozone layer, life can survive higher than for their counterparts around Sun-like stars. in the ocean or the subsurface, meaning that flares may Even though the overall chance of a strong CME hitting not be a strict barrier to habitability. an exoplanet is relatively low, it is interesting to consider From a different point of view, ozone has been pro- the potential impact on habitability. CMEs impinging posed as a (Segura et al. 2003). Exoplanets on a -rich atmosphere can be useful for prebi- subject to sufficiently intense flares would not be able to otic chemistry via the efficient production of hydrogen develop ozone layers, leading to a potential ‘false neg- cyanide, and possibly for atmospheric warming, via the ative’ scenario for ozone. Other biosignatures may be production of (Airapetian & others 2016). similarly affected. The FFDs we derive in this paper On the other hand, CMEs can contribute significantly will enable detailed modelling of the atmospheric states to the potential loss or transformation of exoplanet at- of planets orbiting these stars, and will enable the as- mospheres through ion pick up (Lammer et al. 2007; sessment of false positive and false negative scenarios for Cohen et al. 2015). For M dwarfs, the exoplanet’s mag- biosignature search for these objects. netic field has to be up to hundreds of Gauss strong to shield from atmospheric loss through CMEs, making 5.4. Additional constrains for habitability them a more detrimental factor than stellar winds (Kay We discussed selected aspects of how flare events can et al. 2016). Very energetic particles from the CME influence an exoplanet’s habitability: how flares could may produce secondary particles that can reach the sur- trigger prebiotic chemistry (Section 5.1), the impact of face, changing mutation rates for life there (Smith et al. coronal mass ejections accompanied with flares (Sec- 2004). tion 5.2), and ozone depletion (Section 5.3). A more comprehensive view of habitability should 5.3. Ozone depletion also consider the following factors: the exoplanet’s or- The radiation from stellar proton events (SPEs) could bit (semi-major axis and eccentricity); the stellar type; deleteriously impact the habitability of exoplanets. stellar companions; the exoplanet’s atmospheric compo- However, Atri(2017) find that while energetic SPEs sition or absence of an atmosphere; the exoplanet’s sur- could trigger extinction events among complex life, face; potential sub-surface habitability; and many other SPEs are unlikely to lead surface sterilization. In partic- factors (see e.g. Seager et al. 2013; Barnes et al. 2015; ular, exoplanets with magnetic fields and/or substantial Shields et al. 2016; Meadows & Barnes 2018). atmospheres are sufficiently shielded from SPEs. The impact of protons and photons from M dwarf 5.5. On the habitability around TIC 260506296 stellar flares on the atmosphere of a modern-Earth ana- and other superflare M dwarfs log exoplanet has been modeled by Tilley et al.(2019) One of the most interesting targets for habitable ex- (see also e.g. Youngblood et al. 2017). They find that oplanet search will be M dwarfs whose FFDs fill the the impact of flares of energy 1034 ergs at a frequency lower right corner of the panels in Fig. 11. In these ar- of 1 month−1 or greater will result in the removal of eas, the flaring is energetic enough to trigger prebiotic 99.99% or more of the ozone layer, incidentally admit- chemistry (green area), but not with such a large fre- ting potentially-sterilizing doses of UV radiation to the quency that would lead to ozone depletion (red areas). 16

This is an particularly interesting parameter space for exoplanets lists provided by Stephen Kane4 and John M dwarfs, which can soon be explored further with con- Southworth5 for the TESS Guest Observer program. tinuous one-year observations in the TESS continuous We do, however, find four matches among the cur- viewing zone. Note, however, that we here assume the rent TESS alerts6, which could potentially be exoplanets FFDs can be extrapolated in this regime; findings from transiting flaring stars from our sample: Kepler suggest that this might not always be the case • TIC 32090583 is an M dwarf with a 4.84 R exo- for M dwarfs (e.g. Davenport 2016). ⊕ planet candidate on a 0.438 day orbit. What can we say about TIC 260506296, which shows the largest superflare in the sample? Its flux increases • TIC 70797900 is an early M dwarf hosting a single- by a factor of 16.1 in the TESS band over the course of transit candidate with unknown period. How- 34.7 1 h, releasing a bolometric energy of 10 erg (see Sec- ever, its large candidate radius of 17.06 R⊕ and tion 4.5). In contrast, the Carrington event on our Sun V-shaped transit could indicate an eclipsing bi- (G2V; 5772 K) released 1032 erg. Yet, the flare energy of nary scenario. these events on M dwarfs depend on the quiescent flux. • TIC 206609630 is potentially an M dwarf with a Taking the radius and temperature differences between 5.51 R exoplanet candidate on a 0.335 day orbit M dwarfs and the Sun into account (Eqs.5–9), it be- ⊕ (note that it has a Gaia source duplication flag). comes clear how much more extreme flares are on these prime targets for exoplanet missions. • TIC 272086159 is a mid M dwarf with a 9.77 R⊕ The superflare on 260506296 is comparable to, yet exoplanet candidate transiting every 16.156 days. smaller than, the largest superflare reported for Proxima Centauri by Howard et al.(2018). The Proxima Cen- These exo-solar systems, if confirmed, could allow in- tauri superflare temporarily increased the star’s bright- teresting case studies. However, the four candidates ness by a factor of at least 38, releasing 1033.5 erg. Inter- themselves likely will have thick gas envelopes, increas- estingly, 260506296 (M4.5V, 3189 K) and Proxima Cen- ing the temperature beneath the atmosphere to a level tauri (M5.5V, 3050 K) are comparable stars. Such large which is too high for liquid water and a solid surface. stellar flares have been suggested to significantly impact Nevertheless, potential planetary companions or exo- the habitability of putative planetary companions (e.g., moons (see e.g. Teachey & Kipping 2018) might provide Howard et al. 2018). the necessary conditions. We find that 260506296 indeed provides substantial 6. CONCLUSION flaring to trigger prebiotic chemistry on nearby worlds We detect and analyze stellar flares in the short ca- (see Tables1 and2). Its superflare was potentially ac- dence (2 minute) lightcurves from the first TESS data companied by a CME event of 1023 g, and the star’s FFD release. To do so, we develop a flare detection pipeline, does not fall into the regime of ozone depletion by stellar whose candidates we visually inspect to create a vetted proton events. All of this is opening interesting avenues flare candidate list. We then apply our newly developed for future studies on this potential exoplanet host, as allesfitter software to fit the profiles of each flare well as other comparable M dwarfs. with different models, ranging from pure noise to com- plex flare sequences. Using Nested Sampling, we com- pute the Bayesian evidence of each model. This allows 5.6. Four TESS exoplanet candidates orbiting flaring us to robustly select the favoured scenario. M dwarfs We find 1228 flaring stars in the first two TESS sec- Whether we can actually detect exoplanets around tors, with a total count of 8695 flares. The largest am- such strongly flaring stars is a separate matter. A large plitude flare appears on the M dwarf 260506296 and fraction of them are fast rotators with high flare frequen- increases the brightness by a factor of 16.1. The flare cies, which can inhibit the detection of small planets with the highest energy output is on the G-type giant (see e.g. Berta et al. 2012; Kipping et al. 2017). Addi- star 332487879, releasing 1036.9 erg and could be accom- tionally, planetary mass measurements can be hindered panied by a coronal mass ejection (CME) of 1022.1 g. if the stellar rotation and planetary orbital periods are similar or harmonics. This is particularly a problem for 4 heasarc.gsfc.nasa.gov/docs/tess/data/approved-programs/ habitable-zone planets around field-aged early M dwarfs G011183.txt, online 28 Dec 2018 (Newton et al. 2016; Vanderburg et al. 2016). 5 heasarc.gsfc.nasa.gov/docs/tess/data/approved-programs/ We do not find any known exoplanets among our G011112.txt, online 28 Dec 2018 flare stars when cross-matching Table2 with the known 6 archive.stsci.edu/prepds/tess-data-alerts/, online 28 Dec 2018 Stellar Flares from the First Tess Data Release 17

Among all observed stars, flares appear on 30% of mid tems can open an interesting avenue for future studies to late M dwarfs, on 5% of early M dwarfs, and on less of habitability. than 1% of F, G and K stars. 673 of all flaring stars are It is important to note that these findings alone do not early to late M dwarfs, highlighting that TESS explores allow to rule on the possibility of life on potential exo- an important parameter space for flare studies. In total, planets. Our study provides novel methods and insights we find 531 early M dwarfs and 142 late M dwarfs. to derive an overview. Yet, when addressing complex Moreover, we investigate flaring as a function of pho- topics such as exoplanet habitability, any interesting in- tometrically measured stellar rotation periods (‘photo- dividual system deserves a detailed study to consider a metric rotator’). 60% of fast rotating M dwarfs are flare variety of interrelated factors. Fortunately, TESS will stars. Of the M dwarfs without detected rotation pe- continue delivering these prime targets. Those in the riods, only 10% flare. We further find that star’s with mission’s continuous viewing zone in particular will pro- higher flare rates also have an increased maximum flare vide reliable statistics on the largest, and potentially amplitude and FWHM. Photometric rotators and other rarest, superflares. stars show comparable flare rates, amplitudes, and en- ACKNOWLEDGMENTS ergies; however, we note that our rotation information We thank Ward Howard for engaging discussions mainly covers the regime of Prot < 5 d. Among photo- about stellar flares. Funding for the TESS mission metric rotators, there is a tentative decrease of the flare is provided by NASA’s Science Mission directorate. rate for Prot < 0.3 d. This paper includes data collected by the TESS mis- We analyze the flare frequency distributions (FFDs) sion, which are publicly available from the Mikulski in the context of prebiotic chemistry, CMEs and ozone Archive for Space Telescopes (MAST). MNG acknowl- depletion. On the one hand, flares have been suggested edges support from MIT’s Kavli Institute as a Torres to deliver the required ultraviolet energy to trigger bio- postdoctoral fellow. TD acknowledges support from genesis on exoplanets. We find 14 stars, including 11 MIT’s Kavli Institute as a Kavli postdoctoral fellow. M dwarfs, which could fulfill the criteria of a minimum PBR acknowledges support from the Simons Founda- flare rate and energy. Nevertheless, most stars do not tion [SCOL awards 599634]. This work was supported seem to be able to provide the necessary ultraviolet en- in part by grants from the Simons Foundation (SCOL ergy through their flares alone. On the other hand, grant # 495062 to S.R). MHK acknowledges Allan R. CMEs and stellar proton events (SPEs) associated with Schmitt and Troy Winarski for making their light curve flares could further impact existing life, with SPEs po- softwares freely available. This research has made use tentially causing ozone depletion for exoplanet’s atmo- of IMCCE’s SkyBoT VO tool. spheres. We find that potential exoplanets around up to Facilities: TESS (Ricker et al. 2014) 100 stars might suffer from this effect. Noteably, these Software: python (van Rossum 1995), numpy are mostly F, G, and K type stars, while only up to 100 (van der Walt et al. 2011), scipy (Jones et al. M dwarfs seem affected. This can negatively impact 2001), matplotlib (Hunter 2007), tqdm (doi:10.5281/ habitability and the search for biosignatures. zenodo.1468033), seaborn (seaborn.pydata.org/index. Four flaring M dwarfs host exoplanet candidates from html), allesfitter (G¨unther & Daylan 2020, 2019), recent TESS alerts: TIC 32090583, TIC 70797900, TIC ellc (Maxted 2016), aflare (Davenport et al. 2014), 206609630, and TIC 272086159. Together with other dynesty (Speagle 2020), emcee (Foreman-Mackey M dwarfs, such as the superflaring 260506296, these sys- et al. 2013), celerite (Foreman-Mackey et al. 2017), corner (Foreman-Mackey 2016). 18

Table 1. Catalog of all individual flares found in TESS sectors 1 and 2.

TIC ID sector outburst flare tpeak Amp. FWHM Ebol MCME ... (BJD) (rel. flux) (d) (erg) (g) ...... 167602025 2 2 1 2458373.524 0.008 0.006 8.15e+33 3.42e+20 ... 167695269 1 1 1 2458326.726 0.093 0.002 7.62e+32 8.22e+19 ... 167695269 1 1 1 2458336.636 0.157 0.004 2.18e+33 1.71e+20 ... 167695269 1 2 1 2458343.026 0.512 0.002 4.25e+33 2.74e+20 ... 167696018 1 1 1 2458325.491 0.010 0.010 5.43e+33 3.23e+20 ... 167696018 1 2 1 2458328.993 0.013 0.019 2.18e+34 5.24e+20 ...... With one row for each flare, the following values are listed: the TIC ID; the TESS sector; the outburst number; the flare number; the posterior median for the peak time (tpeak), amplitude (Amp.) and full-width at half-maximum (FWHM ); the bolometric energy of the flare (Ebol.); and the possible mass of a coronal mass ejection following the flare (MCME; see Section 5.2). The machine-readable version is available in the online journal. It additionally contains lower and upper limits where applicable. For ease-of-use, it also includes a copy of the per-star columns shown in Table2 (see below). Stellar Flares from the First Tess Data Release 19 ); the . sec N ); the stellar radius ); the maximum and mean); the maximum eff . T bol E Amp. mean preb. ozone depl. ozone depl. ; see Section 3.8 ). an indication of and max. and . FFD bol ; the number of TESS sectors ( β E KELT rot. and Amp. max. FFD TIC ID α rot ); the stellar effective temperature ( P TESS rot. ) threshold (see Section 5.1 ). For readability, several columns rot TESS mag ) (d) (d) ; see Section 5.1 ); and an indication of whether a star’s flaring could lead 2 s cm g P ozone depl. perm. log ) (log preb. chem.

...... ? R R ); the maximum and mean flare amplitude ( eff . T ) or permissive ( fla ); the maximum bolometric energy of the star’s flares ( N mean); the TESS magnitude ( TESS- CME M FWHM mean . Catalog of all flaring stars found in TESS sectors 1 and 2. ozone depl. cons. CME and M max. and ); the rotation period extracted from TESS and KELT data ( Table 2 g CME 1) ... (g) (K) ( M − FWHM max. ); the number of identified flares ( . Amp. ... max.(flux ... mean mag TESS KELT chem. cons. perm. out . N fla N . out N . sec N ); the logarithm of the surface gravity (log ? R ( which star flares enoughto to ozone potentially depletion trigger on prebioticare its chemistry hidden exoplanets on in for its thisvalues a exoplanets where view. conservative ( ( applicable. The machine-readable version is available in the online journal, and contains all described columns. It also includes all lower and upper limits number of identified outbursts ( and mean mass of coronal mass ejections ( With one row for each star, themean maximum/mean flare values full-width at of half-maximum the ( flaring and the stellar values are shown as follows: the TIC ID 167602025 2167695269 1167814740 5 2 3102032397 8 1 7 3 0.061 11 11 0.512 0.589 13 ... 2.18E+21 ... 0.093 11.277 ... 2.74E+20 5341 6.62E+20 13.001 0.944 13.243 ... 3381 - 3283 1.68E+21 0.374 0.524 10.593 4.851 4.717 5435 1.105 - 2.477 4.329 0.403 . 0.719 - 3.715 - -0.677 1.343 22.236 no -0.586 -1.045 - 18.594 35.482 no yes no - no yes no no yes no no no 20

REFERENCES

Aarnio, A. N., Matt, S. P., & Stassun, K. G. 2012, ApJ, Dole, S. H. 1964, Habitable planets for man (Blaisdell 760, 9 Publishing Company) Aarnio, A. N., Stassun, K. G., Hughes, W. J., & McGregor, Doyle, J. G., Shetye, J., Antonova, A. E., et al. 2018, S. L. 2011, Solar Physics, 268, 195 MNRAS, 475, doi:10.1093/mnras/sty032 Airapetian, V. S., Jackman, C. H., Mlynczak, M., Danchi, Durney, B. R. 1993, ApJ, 407, 367 W., & Hunt, L. 2017, Scientific Reports, 7, 14141 Estrela, R., & Valio, A. 2018, , 18, 1414 Airapetian, V. S., & others. 2016, Geosci., 9, 452 Foreman-Mackey, D. 2016, JOSS, 24, Anglada-Escud´e,G., Amado, P. J., Barnes, J., et al. 2016, doi:10.21105/joss.00024 Nature, 536, 437 Foreman-Mackey, D., Agol, E., Ambikasaran, S., & Angus, Astudillo-Defru, N., Delfosse, X., Bonfils, X., et al. 2017, R. 2017, celerite: Scalable 1D Gaussian Processes in A&A, 600, A13 C++, Python, and Julia, Astrophysics Source Code Atri, D. 2017, MNRAS, 465, L34 Library, , Balona, L. A., Broomhall, A.-M., Kosovichev, A., et al. Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, 2015, MNRAS, 450, 956 J. 2013, PASP, 125, 306 Barnes, R., Meadows, V. S., & Evans, N. 2015, ApJ, 814, 91 France, K., Parke Loyd, R. O., Youngblood, A., et al. 2016, Benz, A. O., & G¨udel,M. 2010, ARA&A, 48, 241 ApJ, 820, 89 Berta, Z. K., Irwin, J., Charbonneau, D., Burke, C. J., & Falco, E. E. 2012, AJ, 144, 145 Gershberg, R. E. 1972, APSS, 19, 75 Berthier, J., Carry, B., Vachier, F., Eggl, S., & Santerne, A. Gillon, M., Triaud, A. H. M. J., Demory, B.-O., et al. 2017, 2016, MNRAS, 458, 3394 Nature, 542, 456 Berthier, J., Vachier, F., Thuillot, W., et al. 2006, in Grenfell, J. L., Grießmeier, J.-M., von Paris, P., et al. 2012, Astronomical Society of the Pacific Conference Series, Astrobiology, 12, 1109 Vol. 351, Astronomical Data Analysis Software and G¨unther, M. N., & Daylan, T. 2019, Allesfitter: Flexible Systems XV, ed. C. Gabriel, C. Arviset, D. Ponz, & Star and Exoplanet Inference From Photometry and S. Enrique, 367 Radial Velocity, Astrophysics Source Code Library, , Bj¨orn,L. O., & others. 2015, in Photobiology: The Science —. 2020, arXiv e-prints, arXiv:2003.14371 of Light and Life, ed. L. O. Bj¨orn(Springer New York), Hawley, S. L., Davenport, J. R. A., Kowalski, A. F., et al. 415–420 2014, ApJ, 797, 121 Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, Hawley, S. L., & Pettersen, B. R. 1991, ApJ, 378, 725 327, 977 Henry, T. J., Kirkpatrick, J. D., & Simons, D. A. 1994, AJ, Browning, M. K. 2008, ApJ, 676, 1262 108, 1437 Bryce, C. C., Horneck, G., Rabbow, E., Edwards, H. G. M., Hodgson, R. 1859, MNRAS, 20, 15 & Cockell, C. S. 2015, IJAsB, 14, 115 Howard, W. S., Tilley, M. A., Corbett, H., et al. 2018, ApJ, Carrington, R. C. 1859, MNRAS, 20, 13 860, L30 Chang, H. Y., Lin, C. L., Ip, W. H., et al. 2018, ApJ, 867, Huang, S.-S. 1959, AmSci, 47, 397 78 Hunter, J. D. 2007, CSE, 9, 90 Cohen, O., Ma, Y., Drake, J. J., et al. 2015, ApJ, 806, 41 Jackman, J. A., Wheatley, P. J., Pugh, C. E., et al. 2019, Covey, K. R., Hawley, S. L., Bochanski, J. J., et al. 2008, MNRAS, 482, 5553 AJ, 136, 1778 Jackman, J. A. G., Wheatley, P. J., Pugh, C. E., et al. Cullum, J., Stevens, D., & Joshi, M. 2014, Astrobiology, 14, 2018, MNRAS, 477, 4655 645 Cullum, J., & Stevens, D. P. 2016, Proceedings of the Jenkins, J. M. 2002, ApJ, 575, 493 National Academy of Science, 113, 4278 —. 2017, Kepler Data Processing Handbook, Tech. rep. Davenport, J. R. A. 2016, ApJ, 829, 23 Jenkins, J. M., Chandrasekaran, H., McCauliff, S. D., et al. Davenport, J. R. A., Hawley, S. L., Hebb, L., et al. 2014, 2010, in \procspie, Vol. 7740, Software and ApJ, 797, 122 Cyberinfrastructure for Astronomy, 77400D Diaz, B., & Schulze-Makuch, D. 2006, Astrobiology, 6, 332 Jenkins, J. M., Twicken, J. D., McCauliff, S., et al. 2016, in Dittmann, J. A., Irwin, J. M., Charbonneau, D., et al. \procspie, Vol. 9913, Software and Cyberinfrastructure 2017, Nature, 544, 333 for Astronomy IV, 99133E Stellar Flares from the First Tess Data Release 21

Jones, E., Oliphant, T., Peterson, P., & Others. 2001, Newton, E. R., Irwin, J., Charbonneau, D., et al. 2017, SciPy: Open Source Scientific Tools for Python, 2001 ApJ, 834, 85 (http://www.scipy.org/), , Newton, E. R., Irwin, J., Charbonneau, D., Kaltenegger, L., & Traub, W. A. 2009, ApJ, 698, 519 Berta-Thompson, Z. K., & Dittmann, J. A. 2016, ApJ, Kane, S. R., Hill, M. L., Kasting, J. F., et al. 2016, ApJ, 821, L19 830, 1 Newton, E. R., Mondrik, N., Irwin, J., Winters, J. G., & Kass, R. E., & Raftery, A. E. 1995, J. Am. Stat. Assoc., 90, Charbonneau, D. 2018, AJ, 156, 217 773 Noyes, R. W., Hartmann, L. W., Baliunas, S. L., Duncan, Kasting, J. F., Kopparapu, R., Ramirez, R. M., & Harman, D. K., & Vaughan, A. H. 1984, ApJ, 279, 763 C. E. 2014, Proceedings of the National Academy of Nutzman, P., & Charbonneau, D. 2008, PASP, 120, 317 Science, 111, 12641 Oelkers, R. J., Rodriguez, J. E., Stassun, K. G., et al. 2018, Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, AJ, 155, 39 Icarus, 101, 108 O’Malley-James, J. T., & Kaltenegger, L. 2018, MNRAS, Kay, C., Opher, M., & Kornbleuth, M. 2016, ApJ, 826, 195 481, 2487 Kiang, N. Y., Segura, A., Tinetti, G., et al. 2007, Osten, R. A., & Wolk, S. J. 2015, ApJ, 809, 79 Astrobiology, 7, 252 P´al,A., Moln´ar,L., & Kiss, C. 2018, PASP, 130, 114503 Kipping, D. M., Cameron, C., Hartman, J. D., et al. 2017, Parker, E. N. 1979, Cosmical magnetic fields. Their origin AJ, 153, 93 and their activity Kopparapu, R. K., Ramirez, R. M., SchottelKotte, J., et al. Parnell, C. E., & Jupp, P. E. 2000, ApJ, 529, 554 2014, ApJ, 787, L29 Pecaut, M. J., & Mamajek, E. E. 2013, ApJS, 208, 9 Pierrehumbert, R., & Gaidos, E. 2011, ApJ, 734, L13 Kopparapu, R. K., Ramirez, R., Kasting, J. F., et al. 2013, Ramirez, R. M., & Kaltenegger, L. 2014, ApJ, 797, L25 ApJ, 765, 131 Ranjan, S., Wordsworth, R., & Sasselov, D. D. 2017, ApJ, Kowalski, A. F., Hawley, S. L., Hilton, E. J., et al. 2009, 843, 110 AJ, 138, 633 Reid, I. N., Cruz, K. L., Allen, P., et al. 2004, AJ, 128, 463 Kowalski, A. F., Hawley, S. L., Wisniewski, J. P., et al. Ricker, G. R., Winn, J. N., Vanderspek, R., et al. 2014, in 2013, ApJS, 207, 15 SPIE Conf. Series, Vol. 9143, 20 Lacy, C. H., Moffett, T. J., & Evans, D. S. 1976, APJS, 30, Rimmer, P. B., Xu, J., Thompson, S. J., et al. 2018, Science 85 Advances, 4, eaar3302 Lammer, H., Lichtenegger, H. I. M., Kulikov, Y. N., et al. Rodono, M., Cutispoto, G., Pazzani, V., et al. 1986, A&A, 2007, Astrobiology, 7, 185 165, 135 Law, N. M., Fors, O., Ratzloff, J., et al. 2015, PASP, 127, Scalo, J., Kaltenegger, L., Segura, A. G., et al. 2007, 234 Astrobiology, 7, 85 Lingam, M., & Loeb, A. 2017, ApJ, 848, 41 Scargle, J. D. 1982, ApJ, 263, 835 Lomb, N. R. 1976, Ap&SS, 39, 447 Seager, S., Bains, W., & Hu, R. 2013, ApJ, 777, 95 Maehara, H., Shibayama, T., Notsu, Y., et al. 2015, Earth, Segura, A., Krelove, K., Kasting, J. F., et al. 2003, Planets and Space, 67, 59 Astrobiology, 3, 689 Maehara, H., Shibayama, T., Notsu, S., et al. 2012, Nature, Segura, A., Walkowicz, L. M., Meadows, V., Kasting, J., & 485, 478 Hawley, S. 2010, Astrobiology, 10, 751 Maxted, P. F. L. 2016, A&A, 591, A111 Shibata, K., & Takasao, S. 2016, in Astrophysics and Space Meadows, V. S., & Barnes, R. K. 2018, in Handbook of Science Library, Vol. 427, Magnetic Reconnection: Exoplanets, ISBN 978-3-319-55332-0. Springer Concepts and Applications, ed. W. Gonzalez & International Publishing AG, part of Springer Nature, E. Parker, 373 2018, id.57, 57 Shibayama, T., Maehara, H., Notsu, S., et al. 2013, ApJS, Moffatt, H. K. 1978, Magnetic field generation in 209, 5 electrically conducting fluids (Cambridge University Shields, A. L., Ballard, S., & Johnson, J. A. 2016, PhysRep, Press) 663, 1 Mondrik, N., Newton, E., Charbonneau, D., & Irwin, J. Simon, T., & Drake, S. A. 1989, ApJ, 346, 303 2019, ApJ, 870, 10 Smith, D. S., Scalo, J., & Wheeler, J. C. 2004, Icarus, 171, Mullan, D. J., & Bais, H. P. 2018, ApJ, 865, 101 229 22

Smith, J. C., Stumpe, M. C., Van Cleve, J. E., et al. 2012, van Rossum, G. 1995, Python tutorial, Tech. Rep. PASP, 124, 1000 CS-R9526, Centrum voor Wiskunde en Informatica Speagle, J. S. 2020, MNRAS, doi:10.1093/mnras/staa278 (CWI), Amsterdam Stassun, K. G., Hebb, L., Covey, K., et al. 2011, in Vanderburg, A., Plavchan, P., Johnson, J. A., et al. 2016, Astronomical Society of the Pacific Conference Series, MNRAS, 459, 3565 Vol. 448, 16$ˆ{th}$ Cambridge Workshop on Cool Stars, Venot, O., Rocchetto, M., Carl, S., Roshni Hashim, A., & Stellar Systems, and the Sun. ASP Conference Series, Decin, L. 2016, ApJ, 830, 77 Vol. 448, proceedings of a conference held August 28- Vin´ıcius,Z., Barentsen, G., Hedges, C., Gully-Santiago, M., September 2, 2010 at the University of Washington, & Cody, A. M. 2018, KeplerGO/lightkurve, , , Seattle, Washington. Edited by Christopher M. Johns-K, doi:10.5281/zenodo.1181928 ed. C. Johns-Krull, M. K. Browning, & A. A. West, 505 Walkowicz, L. M., Basri, G., Batalha, N., et al. 2011, AJ, Stellingwerf, R. F. 1978, ApJ, 224, 953 141, 50 Stelzer, B., Damasso, M., Scholz, A., & Matt, S. P. 2016, West, A. A., Hawley, S. L., Bochanski, J. J., et al. 2008, MNRAS, 463, 1844 AJ, 135, 785 Stetson, P. B. 1996, PASP, 108, 851 West, A. A., Weisenburger, K. L., Irwin, J., et al. 2015, Stumpe, M. C., Smith, J. C., Catanzarite, J. H., et al. 2014, ApJ, 812, 3 PASP, 126, 100 Wheatley, P., West, R., Goad, M., et al. 2018, MNRAS, Sutherland, J. 2015, in European 475, doi:10.1093/mnras/stx2836 Congress, 2015–1 Wright, N. J., Newton, E. R., Williams, P. K. G., Drake, Szab´o,R., S´arneczky, K., Szab´o,G. M., et al. 2015, AJ, J. J., & Yadav, R. K. 2018, MNRAS, 479, 2351 149, 112 Xu, J., Ritson, D. J., Ranjan, S., et al. 2018, Chemical Teachey, A., & Kipping, D. M. 2018, Science Advances, 4, Communications, 54, 5566 eaav1784 York, D. G., Adelman, J., Anderson John E., J., et al. Tilley, M. A., Segura, A., Meadows, V., Hawley, S., & 2000, AJ, 120, 1579 Davenport, J. 2019, Astrobiology, 19, 64 Youngblood, A., France, K., Loyd, R. O. P., et al. 2017, van der Walt, S., Colbert, S. C., & Varoquaux, G. 2011, ApJ, 843, 31 CSE, 13, 22 Zhan, Z., G¨unther, M., Rappaport, S., et al. 2019, ApJ, Van Doorsselaere, T., Shariati, H., & Debosscher, J. 2017, 876, doi:10.3847/1538-4357/ab158c ApJSS, 232, 26