TOI-2374 b and TOI-3071 b: two metal-rich sub-Saturns well within the Neptunian desertthanks: This paper includes data gathered with the 6.5 meter Magellan Telescopes located at Las Campanas Observatory, Chile.

Alejandro Hacker1, Rodrigo F. Díaz1, David J. Armstrong2,3, Jorge Fernández Fernández2,3, Simon Müller4, Elisa Delgado-Mena5, Sérgio G. Sousa5, Vardan Adibekyan5,6, Keivan G. Stassun7, Karen A. Collins8, Samuel W. Yee9,8, Daniel Bayliss2, Allyson Bieryla8, François Bouchy10, R. Paul Butler11, Jeffrey D. Crane12, Xavier Dumusque10, Joel D. Hartman9, Ravit Helled4, Jon Jenkins13, Marcelo Aron F. Keniger2,3, Hannah Lewis14, Jorge Lillo-Box15, Michael B. Lund16, Louise, D. Nielsen17,18, Ares Osborn19,20, David Osip21, Martin Paegert8, Don J. Radford22, Nuno C. Santos5,6, Sara Seager23,24,25, Stephen A. Shectman12, Gregor Srdoc26, Paul A. Strøm2,3, Thiam-Guan Tan27, Johanna K. Teske11, Michael Vezie23,24, David Watanabe28, Cristilyn N. Watkins8, Peter J. Wheatley2,3 Joshua N. Winn9, Bill Wohler29,13 and Carl Ziegler30
The authors’ affiliations are shown in Appendix E
E-mail: ahacker@unsam.edu.ar
(Accepted 2024 June 01. Received 2024 May 29; in original form 2024 May 08)
Abstract

We report the discovery of two transiting planets detected by the Transiting Exoplanet Survey Satellite (TESS), TOI-2374 b and TOI-3071 b, orbiting a K5V and an F8V star, respectively, with periods of 4.31 and 1.27 days, respectively. We confirm and characterize these two planets with a variety of ground-based and follow-up observations, including photometry, precise radial velocity monitoring, and high-resolution imaging. The planetary and orbital parameters were derived from a joint analysis of the radial velocities and photometric data. We found that the two planets have masses of (57±4)plus-or-minus574(57\pm 4)( 57 ± 4 ) Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT or (0.18±0.01)plus-or-minus0.180.01(0.18\pm 0.01)( 0.18 ± 0.01 ) MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, and (68±4)plus-or-minus684(68\pm 4)( 68 ± 4 ) Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT or (0.21±0.01)plus-or-minus0.210.01(0.21\pm 0.01)( 0.21 ± 0.01 ) MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, respectively, and they have radii of (6.8±0.3)plus-or-minus6.80.3(6.8\pm 0.3)( 6.8 ± 0.3 ) Rsubscript𝑅direct-sumR_{\oplus}italic_R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT or (0.61±0.03)plus-or-minus0.610.03(0.61\pm 0.03)( 0.61 ± 0.03 ) RJsubscript𝑅𝐽R_{J}italic_R start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT and (7.2±0.5)plus-or-minus7.20.5(7.2\pm 0.5)( 7.2 ± 0.5 ) Rsubscript𝑅direct-sumR_{\oplus}italic_R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT or (0.64±0.05)plus-or-minus0.640.05(0.64\pm 0.05)( 0.64 ± 0.05 ) RJsubscript𝑅𝐽R_{J}italic_R start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, respectively. These parameters correspond to sub-Saturns within the Neptunian desert, both planets being hot and highly irradiated, with Teq745subscript𝑇eq745T_{\rm eq}\approx 745italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT ≈ 745 K𝐾Kitalic_K and Teq1812subscript𝑇eq1812T_{\rm eq}\approx 1812italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT ≈ 1812 K𝐾Kitalic_K, respectively, assuming a Bond albedo of 0.5. TOI-3071 b has the hottest equilibrium temperature of all known planets with masses between 10101010 and 300300300300 Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT and radii less than 1.51.51.51.5 RJsubscript𝑅𝐽R_{J}italic_R start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT. By applying gas giant evolution models we found that both planets, especially TOI-3071 b, are very metal-rich. This challenges standard formation models which generally predict lower heavy-element masses for planets with similar characteristics. We studied the evolution of the planets’ atmospheres under photoevaporation and concluded that both are stable against evaporation due to their large masses and likely high metallicities in their gaseous envelopes.

keywords:
exoplanets -– stars: planetary systems – planets and satellites: gaseous planets – planets and satellites: detection – planets and satellites: individual: (TOI-2374, TIC 439366538 and TOI-3071, TIC 452006073)
pubyear: 2023pagerange: TOI-2374 b and TOI-3071 b: two metal-rich sub-Saturns well within the Neptunian desertthanks: This paper includes data gathered with the 6.5 meter Magellan Telescopes located at Las Campanas Observatory, Chile.E

1 Introduction

Since the discovery of 51 Pegasi b (Mayor & Queloz, 1995), more than 5500 exoplanets have been confirmed, thanks largely to detections from ground based RV surveys such as HARPS (High Accuracy Radial velocity Planet Searcher; Mayor et al., 2003a) , transiting surveys such as WASP (Wide Angle Search for Planets; Pollacco et al., 2006) and space telescopes such as CoRot (Convection, Rotation, and planetary Transits; Baglin et al., 2008), Kepler (Borucki et al., 2010) and, most recently, the Transiting Exoplanet Survey Satellite (TESS, Ricker et al., 2015). TESS was launched in 2018 with the objective of detecting exoplanets around bright nearby stars by monitoring flux variations of hundreds of thousands of stars in the solar neighborhood, covering 95%similar-toabsentpercent95\sim 95\%∼ 95 % of the sky. Furthermore, it has been optimized to find planets around stars that are bright enough to be able to extract meaningful information from radial velocity (RV) observations from the ground (particularly, planetary mass determination) and conduct atmospheric studies with the James Webb Space Telescope (JWST, Gardner et al., 2006) and other facilities. Since its launch, TESS has confirmed 398 exoplanets with thousands of project candidates still awaiting their nature to be established. 111https://exoplanetarchive.ipac.caltech.edu/

During the lifetime of the Kepler and TESS missions, a notable absence of exoplanets was noticed within a defined region of period-radius and period-mass space, the so-called "Neptunian Desert" (also known as the “hot Neptune desert”, “subJovian desert”, and “evaporation desert”), which roughly encompasses Neptunian-sized planets (approximately 2R<Rp<9R2subscriptRdirect-sumsubscript𝑅𝑝9subscriptRdirect-sum2\,\textrm{R}_{\oplus}<R_{p}<9\,\textrm{R}_{\oplus}2 R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT < italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT < 9 R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT and 0.03MJ<Mp<0.8MJ0.03subscriptMJsubscript𝑀𝑝0.8subscriptMJ0.03\,\textrm{M}_{\rm J}<M_{p}<0.8\,\textrm{M}_{\rm J}0.03 M start_POSTSUBSCRIPT roman_J end_POSTSUBSCRIPT < italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT < 0.8 M start_POSTSUBSCRIPT roman_J end_POSTSUBSCRIPT) with periods up to 5similar-toabsent5\sim 5∼ 5 days (Szabó & Kiss, 2011; Mazeh et al., 2016; Owen & Wu, 2017; Owen & Lai, 2018; Deeg et al., 2023). This should not be due to an observational bias, as planets with short period and intermediate size are easily discovered by transit surveys like Kepler. Several theories have been put forward to explain the existence of the desert and its boundaries in the parameter space (e.g., Owen & Lai, 2018; Vissapragada et al., 2022): the lower boundary could be caused by photoevaporation stripping away their gaseous H/He envelopes, therefore reducing their radii/masses and leaving behind a dense core; while the upper boundary seems to be stable against photoevaporation, and may instead be understood as a “tidal disruption barrier”: in order to form planets below and left of the boundary through high-eccentricity migration, they would have to come so close to their host star and no longer be able to succesfully circularize and stabilize.

Many of these planets are readily observable with missions such as TESS and JWST, and also with ground-based spectroscopic instruments such as HARPS (Mayor et al., 2003b) and ESPRESSO (Pepe et al., 2021) due to their close orbits and short periods. The HARPS-NOMADS programme (PI Armstrong, 1108.C-0697) aims to significantly increase the number of planet confirmations in this parameter space with precise masses and radii. With these parameters we can constrain the densities and thus the internal structures of these planets. The few planets found in this desert are likely to have undergone unusual formation and/or evolutionary processes compared to those in more populated parameter spaces so this analysis could lead towards an improved understanding of the formation and evolution of the Neptunian desert.

Residing in the Neptunian desert with masses between 0.10.10.10.1 and 0.40.40.40.4 MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, hot Saturns are a group whose study can help us further understand the divergent formation pathways of small planets and gas giants: on the one hand, they could be the smallest planets formed via runaway gas accretion. They can therefore provide information on the limits of the core accretion mechanism (Petigura et al., 2018) . On the other hand, in accordance with the trend that associates the presence of short-period planets and of large planets with higher host star metallicity (Mulders et al., 2016; Dong et al., 2018), Petigura et al. (2018) found that stars hosting hot sub-Saturns (planets with radii between 4 and 8 Rsubscript𝑅direct-sumR_{\oplus}italic_R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT and periods between 1 and 10 days) have the highest mean stellar metallicity among all planet hosts. This evidence suggests some kind of mechanism connected to high stellar metallicity that leads to these short-period large planets. Lastly, because their lower surface gravities lead to larger atmospheric scale heights than those of typical hot Jupiters, hot Saturns are some of the best targets for transmission spectroscopy observations (Wakeford et al., 2017).

We present here the detection and characterisation of TOI-2374 b and TOI-3071 b, two hot sub-Saturns transiting a K5V and an F8V star, respectively. The observations leading to the detection and confirmation of the two planets include photometry from TESS and ground-based telescopes, spectroscopy from HARPS and PFS and high-spatial resolution imaging from SOAR. The details of these observations are outlined in Section 2. In Section 3 we present the analysis of these observations, including the determination of stellar parameters and the joint modelling used to constrain the planetary parameters. In Section 4 we present the results and discuss the nature of the two targets, including the position of the planets in the Neptune desert and in mass-radius parameter space. This section also includes an analysis of the internal structure and evaporation history of the planet. Finally, in Section 5 we report the conclusions of our work.

2 Observations

2.1 Photometry

2.1.1 TESS photometry

TESS observed TOI-2374, with TESS Input Catalog (TIC) ID 439366538, in sector 1 (UT 2018 July 25–August 22) with a 30 minute cadence and in sector 28 (UT 2020 July 30–August 26) with a 10 minute cadence, both on camera 1 CCD 4. TOI-2374 b was detected in the Full-frame images (FFIs) by the MIT’s Quick-Look Pipeline (QLP, Huang et al., 2020) and alerted as a TESS object of interest (TOI) on 28 October 2020222https://exofop.ipac.caltech.edu/tess/target.php?id=439366538 (Guerrero et al., 2021). The detection gave a period of 4.31362±0.00001plus-or-minus4.313620.000014.31362\pm 0.000014.31362 ± 0.00001 d, an epoch of 2458326.5553872458326.5553872458326.5553872458326.555387  days in BJD, and a depth of 9.490±0.008plus-or-minus9.4900.0089.490\pm 0.0089.490 ± 0.008 parts-per-thousand (ppt).

TOI-3071 (TIC ID 452006073) was monitored by TESS in sectors 10 (UT 2019 March 26–April 22) and 37 (UT 2021 April 2–28) with camera 2 CCD 2 with a 30 and 10 minute cadence FFI, respectively. The QLP detected the candidate and it was alerted as a TOI on 4 June 2021333https://exofop.ipac.caltech.edu/tess/target.php?id=452006073. The detection gave a period of 1.2671±0.0003plus-or-minus1.26710.00031.2671\pm 0.00031.2671 ± 0.0003 d, an epoch of 2459331.9673±0.0017plus-or-minus2459331.96730.00172459331.9673\pm 0.00172459331.9673 ± 0.0017  days in BJD, and a depth of 2.580±0.002plus-or-minus2.5800.0022.580\pm 0.0022.580 ± 0.002 ppt. It was subsequently observed in sector 64 (UT 2023 April 6-May 4) on camera 2 CCD 1 as a 2-minute cadence target. The Science Processing Operations Center (SPOC) at NASA Ames Research Center conducted a transit with a noise-compensating matched filter (Jenkins, 2002; Jenkins et al., 2010, 2020) and detected the transit signature of TOI-3071 b. An initial limb-darkened model fit was performed (Li et al., 2019) and a suite of diagnostic tests were conducted help make or break the planetary nature of the signal (Twicken et al., 2018). The candidate transit signature passed all these diagnostic tests including the difference image centroiding test, which constrained the location of the host star to within 2.1 +/- 2.6 arc sec of the transit source.

For observations of TOI-2374 from TESS sectors 1 and 28 and observations of TOI-3071 from sectors 10 and 37, we downloaded the light curves computed by the QLP and used the Kepler Spline Simple Aperture Photometry (KSPSAP) flux, which is detrended by applying a high-pass filter to remove low-frequency variability from stellar activity or instrumental noise (Huang et al., 2020). We then used the lightkurve python package (Lightkurve Collaboration et al., 2018) to remove all data points flagged as being affected by excess noise (corresponding to stray light from Earth or the Moon in the camera field of view or scattered light). For TOI-3071 observations from sector 64, we downloaded the photometry provided by the SPOC pipeline, and used the Presearch Data Conditioning Simple Aperture Photometry (PDCSAP), from which common trends and artefacts, as well as an estimated contamination from blended sources, have been removed by the SPOC Presearch Data Conditioning (PDC) algorithm (Twicken et al., 2010; Smith et al., 2012; Stumpe et al., 2012; Stumpe et al., 2014). No further detrending of the light curves was deemed necessary as they are relatively flat across the whole time series. The median-normalised light curves for both targets were used in the joint model outlined in section 3.3.

TIC contamination ratios of 6.65 and 0.62 are reported for TOI-2374 and TOI-3071 respectively. These values are computed as the nominal flux from the contaminants divided by the flux from the source. The large values reported for both targets are most likely due to close sources as can be seen in the TESS Target Pixel Files images shown in Figure 1 (created with tpfplotter444Publicly available at https://github.com/jlillo/tpfplotter, Aller et al. 2020). The TPFs show several sources of potential contamination around the target stars (particularly TOI-2374, which has a very close source two magnitudes brighter). Lightcurves from QLP and SPOC data processing pipelines are approximately corrected for blending from nearby stars. The SPOC pipeline computes a crowding metric using a series of assumptions described in Section 2.3.11 of Stumpe et al. (2012). The QLP uses a different set of assumptions described in step 7 in Huang et al. (2020). To further measure the extent of this blending and contribute to the correction, follow up ground-based photometry was necessary; the details of these observations are described below. Additionally, the TPF for TOI-2374 shows that both the target and the bright companion have the same proper motion direction so they might be bounded companions.

Refer to caption
Figure 1: Target Pixel Files (TPF) for TOI-2374 from TESS S28 (top) and TOI-3071 from S64 (bottom). Targets are marked as a white cross. Other Gaia DR3 sources within a limit of 4 Gaia magnitudes difference are marked as red circles, and are numbered in distance order from the targets. The aperture mask is outlined and shaded in red. This figure was created with tpfplotter (Aller et al., 2020).

2.1.2 Light Curve Follow-up Observations

The TESS pixel scale is 21similar-toabsent21\sim 21\arcsec∼ 21 ″ pixel-1 and photometric apertures typically extend out to roughly 1 arcminute, generally causing multiple stars to blend in the TESS photometric aperture. This is precisely the case for TOI-2374 and TOI-3071 (Fig.1). To determine the true source of the TESS detection, we acquired ground-based time-series follow-up photometry of the field around both TOI-2374 and TOI-3071 as part of the TESS Follow-up Observing Program (TFOP; Collins et al., 2018)555https://tess.mit.edu/followup. The follow-up light curves are also used to confirm the transit depth and thus the TESS photometric deblending factor and refine the TESS ephemeris. We used the TESS Transit Finder, which is a customized version of the Tapir software package (Jensen, 2013), to schedule our transit observations. The observations are summarised below and confirm the transit-like events detected by TESS to be occurring within the small TOI-2374 and TOI-3071 follow-up photometric apertures.

LCOGT

We observed two full transit windows of TOI-2374 b on UTC 2021 May 10 and 2021 May 18 in Sloan isuperscript𝑖i^{\prime}italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and Sloan gsuperscript𝑔g^{\prime}italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT bands, respectively, using the Las Cumbres Observatory Global Telescope (LCOGT; Brown et al., 2013) 1.0 m network nodes at South Africa Astronomical Observatory (SAAO) and Siding Spring Observatory (SSO). The images were calibrated by the standard LCOGT BANZAI pipeline (McCully et al., 2018) and differential photometric data were extracted using AstroImageJ (Collins et al., 2017). We used circular photometric apertures with radius 5.15.15.1\arcsec5.1 ″ for the isuperscript𝑖i^{\prime}italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT band observations and 5.45.45.4\arcsec5.4 ″ for the gsuperscript𝑔g^{\prime}italic_g start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT band observations. The target star apertures excluded all of the flux from the nearest known neighbor in the Gaia DR3 catalog (Gaia DR3 6828814283414902784), which is 22similar-toabsent22\sim 22\arcsec∼ 22 ″ northeast of TOI-2374. Each light curve is included in the global modelling described in section 3.3.

Brierfield Private Observatory

We observed one full transit window of TOI-2374 b on UTC 2021 May 18 in B band using the Brierfield Observatory, located near Bowral, New S. Wales, Australia. The 0.36 m telescope is equipped with a 4096×4096409640964096\times 40964096 × 4096 Moravian 16803 camera. The image scale after binning 2×\times×2 is 1.471.471.47\arcsec1.47 ″ pixel-1, resulting in a 50×50505050\arcmin\times 50\arcmin50 ′ × 50 ′ field of view. The photometric data were extracted using the AstroImageJ (AIJ) software package (Collins et al., 2017). Circular photometric apertures with radius 7.47.47.4\arcsec7.4 ″ were used as they excluded all of the flux from the nearest known neighbor.

PEST

We observed a egress window of TOI-3071 b in Rc band on UTC 2021 June 04 using the Perth Exoplanet Survey Telescope (PEST), located near Perth, Australia. The 0.3 m telescope is equipped with a 5544×3694554436945544\times 36945544 × 3694 QHY183M camera. Images are binned 2×2222\times 22 × 2 in software giving an image scale of 0.70.70.7\arcsec0.7 ″ pixel-1 resulting in a 32×21322132\arcmin\times 21\arcmin32 ′ × 21 ′ field of view. A custom pipeline based on C-Munipack666http://c-munipack.sourceforge.net was used to calibrate the images and extract the differential photometry. We used circular photometric apertures with radius 6.46.46.4\arcsec6.4 ″. The target star aperture included the flux from the nearest known neighbor in the Gaia DR3 catalog (Gaia DR3 5342880462297608192), which is 3.8similar-toabsent3.8\sim 3.8\arcsec∼ 3.8 ″ east of TOI-3071. The PEST observation was not included in the modeling due to the lack of ingress coverage and low signal to noise detection.

2.2 Spectroscopy

In this section we describe in detail the high-precision RV measurements used in this paper. These observations are presented in figures 11 and 12. In section 3.2 we perform a further exploration of this data.

2.2.1 HARPS RVs

We obtained radial velocity (RV) measurements of TOI-2374 and TOI-3071 with the High Accuracy Radial velocity Planet Searcher (HARPS) spectrograph mounted on the ESO 3.6 m telescope at the La Silla Observatory in Chile (Mayor et al., 2003a). Under the HARPS-NOMADS large programme (ID 1108.C-0697, PI: Armstrong), a total of 21 spectra were obtained between 9 April 2022 and 7 July 2022 for TOI-2374 and 14 spectra between 19 March and 2 April 2022 for TOI-3071. The instrument (with resolving power R=115 000𝑅115000R=115\,000italic_R = 115 000) was used in high-accuracy mode (HAM). The exposure time for the TOI-2374 observations was 1800 s. This resulted in a signal to noise ratio (SNR) of between 22 and 35 per pixel. The exposure time for the TOI-3071 observations was 2400 s, resulting in a SNR of between 10 and 25 per pixel. The data were reduced using the standard offline HARPS data reduction pipeline, and a K5 template (TOI-2374) or G2 template (TOI-3071) was used to form a weighted cross-correlation function (CCF) to determine the RV values (Baranne et al., 1996; Pepe et al., 2002). The line bisector (BIS) and full-width at half-maximum (FWHM) were measured using methods by Boisse et al. (2011).

Three bad data points were removed from the TOI-2374 dataset. One of them (BJD = 2459711.8) was reduced incorrectly and the other two (BJD = 2459712.8 and BJD = 2459716.8) contained large errors due to bad weather during the observations. The achieved instrumental precision for the rest of the RV measurements is 4absent4\approx 4≈ 4 m s-1. The observations are presented in tables 4 and 5.

2.2.2 PFS RVs

We observed TOI-2374 with the Planet Finder Spectrograph (PFS; Crane et al. 2006; Crane et al. 2008, 2010) on the Magellan/Clay telescope at Las Campanas Observatory in Chile. We obtained seven observations of this target between 2021 Sep 14 and 2021 Nov 17 through an iodine cell, with relatively short exposure times of 5 minutes. These observations were made in 3x3 binning mode to reduce readout noise. We also obtained an iodine-free template spectrum at higher signal to noise, which we used to extract relative RVs. The RVs were extracted using the pipeline described in Butler et al. (1996), achieving a typical instrumental RV precision of 3absent3\approx 3≈ 3 m s-1. We also quantified potential variations in the spectral line profiles using bisector inverse slopes (BIS) measurements. These were derived using the procedures described in Hartman et al. (2019), applied to the iodine-free orders of each spectrum. The BIS measurements are shifted such that the median bisector span for each order is zero. All measurements are presented in Table 6.

2.2.3 TRES

We obtained two reconnaissance spectra on UT 2021 July 21 and UT 2021 August 5 of TOI-2374 using the Tillinghast Reflector Echelle Spectrograph (TRES, Fűrész, 2008) located at the Fred Lawrence Whipple Observatory (FLWO) in Arizona, USA. TRES has a resolving power of similar-to\sim44,000 and operates in the wavelength range 390-910 nm. The spectra were extracted as described in Buchhave et al. (2010).

2.3 High resolution imaging

Refer to caption
Figure 2: SOAR observation of TOI-2374. Black circles correspond to measured data points and the black lines show the fit in two different separation regimes. The 5σ5𝜎5\sigma5 italic_σ detection sensitivity is plotted with the speckle imaging auto-correlation functions inset.
Refer to caption
Figure 3: SOAR observation of TOI-3071. Black circles correspond to measured data points and the black lines show the fit in two different separation regimes. The 5σ5𝜎5\sigma5 italic_σ detection sensitivity is plotted with the speckle imaging auto-correlation functions inset.

High-angular resolution imaging is needed to search for nearby sources that can contaminate the TESS photometry, resulting in an underestimated planetary radius, or be the source of astrophysical false positives, such as background eclipsing binaries (Daemgen et al., 2009; Lillo-Box et al., 2012, for example). We searched for stellar companions to TOI-2374 with speckle imaging on the 4.1-m Southern Astrophysical Research (SOAR) telescope (Tokovinin, 2018) on 3 December 2020 UT, observing in Cousins I-band, a similar visible bandpass as TESS. This observation was sensitive to a 4.7-magnitude fainter star at an angular distance of 1 arcsec from the target. More details of the observations within the SOAR TESS survey are available in Ziegler et al. (2020). The 5σ5𝜎5\sigma5 italic_σ detection sensitivity and speckle auto-correlation functions from the observation are shown in Figure 2. No nearby stars were detected within 3″of TOI-2374 in the SOAR observations.

TOI-3071 was similarly observed by SOAR on 20 March 2022, sensitive to a 5.5-magnitude fainter star at a 1 arcsec separation from the target. The 5σ5𝜎5\sigma5 italic_σ detection sensitivity and speckle auto-correlation functions from this observation are shown in Figure 3. We detected no nearby stars within 3″of TOI-3071 in the SOAR observations.

3 Analysis

3.1 Stellar parameters

Because the final physical parameters of the planets depend directly on the values of the stellar parameters, here we perform several independent methods to measure and derive a range of stellar parameters for TOI-3071 and TOI-2374.

3.1.1 Spectroscopic parameters

The stellar spectroscopic parameters (Teffsubscript𝑇effT_{\mathrm{eff}}italic_T start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, logg𝑔\log groman_log italic_g, microturbulence, [Fe/H]) were estimated using the ARES+MOOG methodology from the respective combined HARPS spectrum of each star. The methodology is described in detail in Sousa et al. (2021); Sousa (2014); Santos et al. (2013). We used the latest version of ARES 777The last version, ARES v2, can be downloaded at https://github.com/sousasag/ARES (Sousa et al., 2007, 2015) to consistently measure the equivalent widths (EW) on the list of iron lines. For TOI-3071 we used the list of lines presented in Sousa et al. (2008), while for TOI-2374 we used the list of lines presented in Tsantaki et al. (2013) which is more appropriate for stars cooler than 5200 K. In the analysis we use a minimization process to find the ionization and excitation equilibrium to converge for the best set of spectroscopic parameters. This process makes use of a grid of Kurucz model atmospheres (Kurucz, 1993) and the radiative transfer code MOOG (Sneden, 1973). We also derived a more accurate trigonometric surface gravity using recent GAIA data following the same procedure as described in Sousa et al. (2021). We estimated rotational velocity vsini𝑣𝑖v\sin iitalic_v roman_sin italic_i values by performing spectral synthesis (with the same code and model atmospheres as for stellar parameters) of two iron lines in the 6705Å region. We fix the macroturbulence velocity with the empirical formula by Doyle et al. (2014), which provides vmac=4.6 km s-1 for TOI-3071. Cool stars are outside of such calibration, hence we adopted a value vmac=2 km s-1 for TOI-2374. Both stars seem to be slow rotators with vsini𝑣𝑖v\sin iitalic_v roman_sin italic_i = 2.0 km s-1 for TOI-3071 and vsini𝑣𝑖v\sin iitalic_v roman_sin italic_i <<< 0.5 km s-1 for TOI-2374. The stellar parameters are presented in tables 1 and 2.

Independently, we derived the stellar atmospheric parameters for TOI-2374 from the TRES spectra using the Stellar Parameter Classification (SPC, Buchhave et al., 2012) tool. SPC cross correlates an observed spectrum against a grid of synthetic spectra based on Kurucz atmospheric models (Kurucz, 1992). With this method we obtained Teff=4937±50subscript𝑇effplus-or-minus493750T_{\mathrm{eff}}=4937\pm 50italic_T start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 4937 ± 50 K, logg=4.61±0.10𝑔plus-or-minus4.610.10\log g=4.61\pm 0.10roman_log italic_g = 4.61 ± 0.10 and [Fe/H]=0.24±0.08delimited-[]FeHplus-or-minus0.240.08[\mathrm{Fe/H}]=0.24\pm 0.08[ roman_Fe / roman_H ] = 0.24 ± 0.08, which are in agreement (within errors) with the results derived from the HARPS spectra.

3.1.2 Abundances

Stellar abundances of the elements were derived using the same tools and models as for stellar parameter determination under the assumption of local thermodynamic equilibrium. For the derivation of chemical abundances of refractory elements we closely followed the methods described in e.g. Adibekyan et al. (2012); Adibekyan et al. (2015); Delgado Mena et al. (2014); Delgado Mena et al. (2017). Abundances of the volatile elements, C and O, were derived following the method of Delgado Mena et al. (2021) and Bertran de Lis et al. (2015). Nine out of the 14 individual spectra of TOI-3071 were contaminated around the 6300.3 Å oxygen line and had to be discarded before being coadded, leading to a lower S/N spectrum and thus a higher error in oxygen abundance. These two elements could not be derived for TOI-2374 because their lines are very weak in cool stars spectra. All the [X/H] ratios are obtained by doing a differential analysis with respect to a high S/N solar (Vesta) spectrum from HARPS. The abundances of the elements are shown in table 7 in the appendix.

Table 1: An overview of stellar properties for TOI-2374.
Property [unit] Value Source
Identifiers
Name TOI-2374
TIC ID 439366538 TICv8.2
Gaia ID 6828814283414902912 Gaia DR3
Astrometric properties
R.A (J2000.0) 21:17:59.59 Gaia DR3
Dec (J2000.0) -22:02:59.36 Gaia DR3
Parallax [mas] 7.3592 ±plus-or-minus\pm± 0.018 Gaia DR3
Distance [pc] 134.655 ±plus-or-minus\pm± 0.8695 TICv8.2
Photometric properties
T mag 11.2057±0.0061plus-or-minus11.20570.006111.2057\pm 0.006111.2057 ± 0.0061 TICv8.2
V mag 12.095±0.080plus-or-minus12.0950.08012.095\pm 0.08012.095 ± 0.080 TICv8.2
G mag 11.818±0.003plus-or-minus11.8180.00311.818\pm 0.00311.818 ± 0.003 Gaia DR3
J mag 10.387±0.022plus-or-minus10.3870.02210.387\pm 0.02210.387 ± 0.022 2MASS
K mag 9.793±0.023plus-or-minus9.7930.0239.793\pm 0.0239.793 ± 0.023 2MASS
Atmosfpheric properties
Teffsubscript𝑇effT_{\mathrm{eff}}italic_T start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT [K] 4802±97plus-or-minus4802974802\pm 974802 ± 97 This work
logg𝑔\log groman_log italic_g [cgs] 4.52±0.05plus-or-minus4.520.054.52\pm 0.054.52 ± 0.05 Gaia DR3
logg𝑔\log groman_log italic_g [cgs] 4.14±0.24plus-or-minus4.140.244.14\pm 0.244.14 ± 0.24 This work
vmicsubscript𝑣micv_{\mathrm{mic}}italic_v start_POSTSUBSCRIPT roman_mic end_POSTSUBSCRIPT [km s-1] 0.62±0.18plus-or-minus0.620.180.62\pm 0.180.62 ± 0.18 This work
[Fe/H]delimited-[]FeH[\mathrm{Fe/H}][ roman_Fe / roman_H ] [dex] 0.15±0.04plus-or-minus0.150.040.15\pm 0.040.15 ± 0.04 This work
vsini𝑣𝑖v\sin iitalic_v roman_sin italic_i [km s-1] <0.5absent0.5<0.5< 0.5 This work
logRHKsubscriptsuperscript𝑅HK\log R^{\prime}_{\rm HK}roman_log italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_HK end_POSTSUBSCRIPT 4.98±0.06plus-or-minus4.980.06-4.98\pm 0.06- 4.98 ± 0.06 This work
Physical parameters
Rssubscript𝑅sR_{\mathrm{s}}italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT [Rsubscript𝑅direct-productR_{\odot}italic_R start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT] 0.69±0.02plus-or-minus0.690.020.69\pm 0.020.69 ± 0.02 This work
Mssubscript𝑀sM_{\mathrm{s}}italic_M start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT [Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT] 0.75±0.01plus-or-minus0.750.010.75\pm 0.010.75 ± 0.01 This work
Protsubscript𝑃rotP_{\rm rot}italic_P start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT [days] 50.2±2.8plus-or-minus50.22.850.2\pm 2.850.2 ± 2.8 This work
τ𝜏\tauitalic_τ [Gyr] 4.10±0.63plus-or-minus4.100.634.10\pm 0.634.10 ± 0.63 This work
  • Sources: TICv8.2 (Stassun et al., 2019), 2MASS (Skrutskie et al., 2006), Gaia Data Release 3 (Gaia Collaboration et al., 2023).

Table 2: An overview of stellar properties for TOI-3071.
Property [unit] Value Source
Identifiers
Name TOI-3071
TIC ID 452006073 TIC v8.2
Gaia ID 5342880462319699200 Gaia DR3
Astrometric properties
R.A (J2000.0) 11:33:06.98 Gaia DR3
Dec (J2000.0) -56:30:12.27 Gaia DR3
Parallax [mas] 2.0101 ±plus-or-minus\pm± 0.0097 Gaia DR3
Distance [pc] 485.219 ±plus-or-minus\pm± 8.8845 TICv8.2
Photometric properties
T mag 11.8706±0.0061plus-or-minus11.87060.006111.8706\pm 0.006111.8706 ± 0.0061 TICv8.2
V mag 12.383±0.057plus-or-minus12.3830.05712.383\pm 0.05712.383 ± 0.057 TICv8.2
G mag 12.267±0.002plus-or-minus12.2670.00212.267\pm 0.00212.267 ± 0.002 Gaia DR3
J mag 11.314±0.024plus-or-minus11.3140.02411.314\pm 0.02411.314 ± 0.024 TICv8.2
K mag 11.005±0.021plus-or-minus11.0050.02111.005\pm 0.02111.005 ± 0.021 TICv8.2
Atmosfpheric properties
Teffsubscript𝑇effT_{\mathrm{eff}}italic_T start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT [K] 6177±62plus-or-minus6177626177\pm 626177 ± 62 This work
logg𝑔\log groman_log italic_g [cgs] 4.37±0.10plus-or-minus4.370.104.37\pm 0.104.37 ± 0.10 This work
logg𝑔\log groman_log italic_g [cgs] 4.32±0.02plus-or-minus4.320.024.32\pm 0.024.32 ± 0.02 Gaia DR3
vmicsubscript𝑣micv_{\mathrm{mic}}italic_v start_POSTSUBSCRIPT roman_mic end_POSTSUBSCRIPT [km s-1] 1.29±0.02plus-or-minus1.290.021.29\pm 0.021.29 ± 0.02 This work
[Fe/H]delimited-[]FeH[\mathrm{Fe/H}][ roman_Fe / roman_H ] [dex] 0.35±0.04plus-or-minus0.350.040.35\pm 0.040.35 ± 0.04 This work
vsini𝑣𝑖v\sin iitalic_v roman_sin italic_i [km s-1] 2.0±0.5plus-or-minus2.00.52.0\pm 0.52.0 ± 0.5 This work
logRHKsubscriptsuperscript𝑅HK\log R^{\prime}_{\rm HK}roman_log italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_HK end_POSTSUBSCRIPT 5.15±0.26plus-or-minus5.150.26-5.15\pm 0.26- 5.15 ± 0.26 This work
Phyisical parameters
Rssubscript𝑅sR_{\mathrm{s}}italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT [Rsubscript𝑅direct-productR_{\odot}italic_R start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT] 1.31±0.04plus-or-minus1.310.041.31\pm 0.041.31 ± 0.04 This work
Mssubscript𝑀sM_{\mathrm{s}}italic_M start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT [Msubscript𝑀direct-productM_{\odot}italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT] 1.29±0.02plus-or-minus1.290.021.29\pm 0.021.29 ± 0.02 This work
Protsubscript𝑃rotP_{\rm rot}italic_P start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT [days] 26.6±6.4plus-or-minus26.66.426.6\pm 6.426.6 ± 6.4 This work
τ𝜏\tauitalic_τ [Gyr] 4.91±0.83plus-or-minus4.910.834.91\pm 0.834.91 ± 0.83 This work (gyro)
τ𝜏\tauitalic_τ [Gyr] 0.95±0.14plus-or-minus0.950.140.95\pm 0.140.95 ± 0.14 This work (clocks &\&& PARAM)
  • Sources: TICv8.2 (Stassun et al., 2019), Gaia Data Release 3 (Gaia Collaboration et al., 2023).

3.1.3 Physical parameters

From the parameters obtained in the spectroscopic analyisis we estimated the stellar radius and mass using the calibrations presented in Torres et al. (2010) (see tables 1 and 2). In addition, we performed an analysis of the broadband spectral energy distribution (SED) of the star together with the Gaia DR3 parallax, in order to determine an empirical measurement of the stellar radius (Stassun & Torres, 2016; Stassun et al., 2017; Stassun et al., 2018). We pulled the JHKS𝐽𝐻subscript𝐾𝑆JHK_{S}italic_J italic_H italic_K start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT magnitudes from 2MASS, the W1–W3 magnitudes from WISE, and the GBPGRPsubscript𝐺BPsubscript𝐺RPG_{\rm BP}G_{\rm RP}italic_G start_POSTSUBSCRIPT roman_BP end_POSTSUBSCRIPT italic_G start_POSTSUBSCRIPT roman_RP end_POSTSUBSCRIPT magnitudes from Gaia. We also utilized the absolute flux-calibrated Gaia spectrum where available. Together, the available photometry spans the full stellar SED over the wavelength range 0.4–10 μ𝜇\muitalic_μm (see figures 4 and 5). We performed a fit using PHOENIX stellar atmosphere models (Husser, T.-O. et al., 2013), adopting from the spectroscopic analysis the effective temperature (Teffsubscript𝑇effT_{\rm eff}italic_T start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT), metallicity ([Fe/H]), and surface gravity (logg𝑔\log groman_log italic_g). We fitted for the extinction AVsubscript𝐴𝑉A_{V}italic_A start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT, limited to the maximum line-of-sight value from the Galactic dust maps of Schlegel et al. (1998). The resulting fits (figures 4 and 5) have AV=0.02±0.02subscript𝐴𝑉plus-or-minus0.020.02A_{V}=0.02\pm 0.02italic_A start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT = 0.02 ± 0.02 and 0.20±0.03plus-or-minus0.200.030.20\pm 0.030.20 ± 0.03 for TOI-2374 and TOI-3071, respectively, with a reduced χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of 1.3 and 1.4, respectively. Integrating the (unreddened) model SED gives the bolometric flux at Earth, Fbol=4.981±0.058×1010subscript𝐹bolplus-or-minus4.9810.058superscript1010F_{\rm bol}=4.981\pm 0.058\times 10^{-10}italic_F start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT = 4.981 ± 0.058 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg s-1 cm-2 and 3.291±0.038×1010plus-or-minus3.2910.038superscript10103.291\pm 0.038\times 10^{-10}3.291 ± 0.038 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT erg s-1 cm-2, respectively. Taking the Fbolsubscript𝐹bolF_{\rm bol}italic_F start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT together with the Gaia parallax directly gives the bolometric luminosity, Lbol=0.2867±0.0034subscript𝐿bolplus-or-minus0.28670.0034L_{\rm bol}=0.2867\pm 0.0034italic_L start_POSTSUBSCRIPT roman_bol end_POSTSUBSCRIPT = 0.2867 ± 0.0034 L and 2.539±0.032plus-or-minus2.5390.0322.539\pm 0.0322.539 ± 0.032 L, respectively. The Stefan-Boltzmann relation then gives the stellar radius, R=0.774±0.025subscript𝑅plus-or-minus0.7740.025R_{\star}=0.774\pm 0.025italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT = 0.774 ± 0.025 R and 1.393±0.030plus-or-minus1.3930.0301.393\pm 0.0301.393 ± 0.030 R, respectively. These estimates differ by 2.6σ2.6𝜎2.6\sigma2.6 italic_σ and 1.7σ1.7𝜎1.7\sigma1.7 italic_σ with the ones obtained from the spectroscopic analysis, respectively. Given that the two sets disagree by less than 3σ3𝜎3\sigma3 italic_σ and that the former analysis works with the full spectrum, we decided to use the ones estimated from the calibrations in Torres et al. (2010) as input for the joint model described in section 3.3.

Refer to caption
Figure 4: Spectral energy distribution of TOI-2374. Red symbols represent the observed photometric measurements, where the horizontal bars represent the effective width of the passband. Blue symbols are the model fluxes from the best-fit PHOENIX atmosphere model (black).
Refer to caption
Figure 5: Spectral energy distribution of TOI-3071. Red symbols represent the observed photometric measurements, where the horizontal bars represent the effective width of the passband. Blue symbols are the model fluxes from the best-fit PHOENIX atmosphere model (black). The absolute flux-calibrated Gaia spectrum is shown as a grey swathe in the inset figure.

We used the relations stated in Noyes et al. (1984) to derive chromospheric activity index logRHKsubscriptsuperscript𝑅HK\log R^{\prime}_{\rm HK}roman_log italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_HK end_POSTSUBSCRIPT values from the HARPS measurements of the Mount-Wilson S-index. With these values we obtained estimates of the rotational period of the stars Protsubscript𝑃rotP_{\rm rot}italic_P start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT. For both stars, Protsubscript𝑃rotP_{\rm rot}italic_P start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT is consistent with the maximum rotation period calculated with the vsini𝑣𝑖v\sin iitalic_v roman_sin italic_i estimates, Prot/sini>70subscript𝑃rot𝑖70P_{\rm rot}/\sin i>70italic_P start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT / roman_sin italic_i > 70 d for TOI-2374 and 33±8plus-or-minus33833\pm 833 ± 8 d for TOI-3071. We used Eq. 3 and Eq. 16 in Barnes (2007) to calculate the gyrochronological stellar ages from the rotation periods, giving 4.34 ±plus-or-minus\pm± 0.67 Gyr and 4.9 ±plus-or-minus\pm± 0.8 Gyr for TOI-2374 and TOI-3071, respectively.

Moreover, we obtained two other independent estimates of the stellar ages τ𝜏\tauitalic_τ: first, we used the web interface PARAM 1.3 888http://stev.oapd.inaf.it/cgi-bin/param_1.3, which estimates the basic intrinsic parameters of stars given their photometric and spectroscopic data following the method described in da Silva et al. (2006). This resulted in stellar age estimates of 4.8 ±plus-or-minus\pm± 4.2 Gyr and 0.8 ±plus-or-minus\pm± 0.6 Gyr for TOI-2374 and TOI-3071, respectively. As expected, the age determination based on evolutionary models is extremely uncertain for main-sequence stars. The age of TOI-2374 formally agrees with the one from gyrochronology. On the other hand, PARAM provides a much younger age for TOI-3071. We then used the chemical abundances of some elements to derive ages through the so-called chemical clocks (i.e. certain chemical abundance ratios which have a strong correlation for age). We applied the 3D formulas described in Table 10 of Delgado Mena et al. (2019), which also consider the variation in age produced by the effective temperature and iron abundance. The chemical clocks [Y/Mg], [Y/Zn], [Y/Ti], [Y/Si], [Y/Al], [Sr/Ti], [Sr/Mg] and [Sr/Si] were used from which we obtain a weighted average age of 1.1 ±plus-or-minus\pm± 0.2 Gyr for TOI-3071. This age is again much younger than the age from gyrochornology. The ages from chemical clocks for cool stars such as TOI-2374 must be taken with caution since the calibrations are obtained using hotter stars. Therefore, we choose to apply the 2D formulas considering the abundance ratios and the metallicity (see Table 8 in Delgado Mena et al., 2019) which provide a weighted average age estimate of 3.3 ±plus-or-minus\pm± 0.7 Gyr. Ages from each individual clock are shown in Table 8. We include in Table 1 the geometric average of the three stellar age estimates for TOI-2374. The error is given by the deviation of the three values. In Table 2 we report two values for the age of TOI-3071, corresponding to the two disagreen results: the gyrochronological age and the weighted average of the ages derived with PARAM and with the chemical clocks.

Refer to caption
Figure 6: GLS Periodograms (Zechmeister & Kürster, 2009) for the TOI-2374 HARPS data (top) and HARPS + PFS data (bottom). The highest peak in the RV periodogram, corresponding to the orbital period of TOI-2374 b is denoted by a dashed vertical red line. The 0.1, 0.05, and 0.01 per cent False Alarm Probabilities (FAP) are calculated using the approach of Baluev (2008) and are shown as dashed horizontal lines. In the HARPS + PFS figures, the second periodogram corresponds to the the radial velocity residuals after the best fit model for a Keplerian orbit has been removed.
Refer to caption
Figure 7: GLS Periodograms (Zechmeister & Kürster, 2009) for the TOI-3071 HARPS data. The highest peak in the RV periodogram, corresponding to the orbital period of TOI-3071 b is denoted by a dashed vertical red line. The 0.1, 0.05, and 0.01 per cent False Alarm Probabilities (FAP) are calculated using the approach of Baluev (2008) and are shown as dashed horizontal lines. The second periodogram corresponds to the the radial velocity residuals after the best fit model for a Keplerian orbit has been removed.
Refer to caption
Figure 8: Light curves of TOI-2374. TESS light curve for Sectors 1 and 28 (grey dots), with time given as Barycentric Julian Date (BJD). The top plots show the TESS QLP KSPSAP light curves from sector 1 (30 minute cadence) and sector 28 (10 min cadence). Overplotted in green is the best-fit solution to the global model resulting from the analysis described in section 3.3. The middle plots show this same data phase folded. The binned points indicate the mean of each bin, with error bars representing the standard error of the mean with the binned data shown in black. Residuals of this fit can be found in the bottom plots.
Refer to caption
Figure 9: Light curves of TOI-2374 from ground-based telescopes. From left to right, the light curves correspond to the Brierfield Observatory, the LCO-SAAO, the LCO-SSO with ip filter, and the LCO-SSO with gp filter. The data are shown as grey dots, with binned values in black. Overplotted in red is the best fit model from the joint fit described in section 3.3. The bottom plots show the residuals of this fit for each lightcurve.
Refer to caption
Figure 10: Light curves of TOI-3071. TESS light curve for Sectors 10, 37 and 64 (grey dots), with time given as Barycentric Julian Date (BJD). The top plots show the TESS QLP KSPSAP light curves from sector 10 (30 minute cadence) and sector 37 (10 minute cadence) and SPOC PDCSAP light curves from sector 64 (2 minute cadence). Overplotted in blue is the best-fit solution to the global model resulting from the analysis described in section 3.3. The middle plots show this same data phase folded. The binned points indicate the mean of each bin, with error bars representing the standard error of the mean with the binned data shown in black. Residuals of this fit can be found in the bottom plots.
Refer to caption
Figure 11: Precise RV measurements of TOI-2374. HARPS measurements are shown as red circles and PFS measurements as blue diamonds. The model plotted is the MCMC median of the global model, corresponding to the orbital period of TOI-2374b. Empirically derived linear and quadratic trends and per-instrument offset γ𝛾\gammaitalic_γ have been subtracted from the raw RV measurements. The error bars represent the reported uncertainty and the empirically derived per-instrument jitter, added in quadrature. These RV measurements are also listed in tables 5 (HARPS) and 6 (PFS). Top: RV measurements plotted against time. Bottom: phase-folded RV measurements.
Refer to caption
Figure 12: Precise RV measurements of TOI-3071. The model plotted is the MCMC median of the global model, corresponding to the orbital period of TOI-3071b. Empirically derived linear and quadratic trends and per-instrument offset γ𝛾\gammaitalic_γ have been subtracted from the raw RV measurements. The error bars represent the reported uncertainty and the empirically derived per-instrument jitter, added in quadrature. These RV measurements are also listed in Table 4. Top: RV measurements plotted against time. Bottom: phase-folded RV measurements.

3.2 RV analysis

We carried out a preliminary stage of RV data exploration in two parts: periodogram analysis through the DACE platform999https://dace.unige.ch and testing of linear RV models with different stellar activity indicators as model covariates. These examinations showed no indication of stellar activity in either of the systems. However, the time span is short for TOI-3071.

Despite having a relatively small number of points, we were able to find the planets of both systems independently of photometry. Generalised Lomb-Scargle Periodograms (GLS, Zechmeister & Kürster, 2009) of the RV data and stellar activity indicators are shown in Figures 6 (TOI-2374) and 7 (TOI-3071). The RV periodograms present peaks above the 0.01 per cent False Alarm Probability at the expected planetary periods. After removing the best fit model for a Keplerian orbit, the periodograms of the RV residuals show no further significant peaks, therefore ruling out the detection of a second planet in each system from these time series.

When looking at periodograms of TOI-3071 stellar activity indicators, we found that the bisector span, contrast and Ca-index show power at low frequency. However, as the time span is relatively short (13.9 days), we cannot constrain the period. We also found one peak in the Hαsubscript𝐻𝛼H_{\alpha}italic_H start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT periodogram (and, to a lesser extent, in the Na periodogram) that is similar to the planetary period but is not large enough to be deemed significant. No other stellar activity indicator shows signs of significant frequencies in this system. As to TOI-2374, there is a significant peak around 1 day in the Na periodogram, which corresponds to a 1 d alias of a low frequency signal, which disappears when we substract a long-term drift from the timeseries. Other than that, this star does not show signs of any other activity.

We computed the Pearson’s R statistic for the RV data and the stellar activity indicators and found no significant correlation between those variables for any system. As for the correlations between the stellar activity indicators and the RV residuals after removing the best fit for a Keplerian orbit, the highest value obtained was R0.61similar-toabsent0.61\sim 0.61∼ 0.61 between TOI-2374 RV residuals and its FWHM, but it strongly depends on a single point (BJD = 2459750.72), without which the correlation disappeared and the Keplerian model stays essentially unchanged. The same situation is seen between the TOI-3071 RV residuals and its contrast, which shared a R of 0.61similar-toabsent0.61\sim 0.61∼ 0.61, but strongly depends on four points. No other indicator shows significant correlation with the RV residuals for none of the systems.

To further examine the effects of stellar activity on the RV data, we tested if the RV curves of our joint model described in section 3.3 should include linear terms corresponding to any of the stellar activity indicators. For this we built simple linear models where the dependent variable was the radial velocity and the covariates were each of the available stellar activity indicators, placed individually in each linear model. We performed an F-test for each model compared to a constant RV model and none of them obtained a p-value smaller than 0.01. We then incorporated into each model a circular orbit with a period fixed to the planetary period found and we repeated the F-test comparing them to the model that only included the circular orbit. Again, none of the stellar activity indicators obtained a p-value smaller than 0.01. This shows that it is not necessary to incorporate linear terms with these covariates to explain the RV data.

3.3 Joint modelling

For each of the TOI-2374 and the TOI-3071 planetary systems, we used the exoplanet package (Foreman-Mackey et al., 2020) to model the RV data and light curves simultaneously. exoplanet utilises the probabilistic programming package PyMC3 (Salvatier et al., 2016) and the light curve modelling package Starry (Luger et al., 2019). The radial velocity mean for each spectroscopic instrument was subtracted from the data. QLP and ground-based light curves were already normalized from the pipelines to have an out-of-transit flux of one. The TESS-SPOC light curves were normalized by dividing them by the median of the flux. All timestamps were converted to the time system used by TESS, i.e. BJD - 2457000 (BJD-TDB).

To model the planetary transits, we used a limb-darkened transit model utilising the quadratic limb-darkening parameterisation in Kipping (2013) and a Keplerian orbit model, which was parameterised for the planet in terms of the orbital period P𝑃Pitalic_P, the epoch t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the eccentricity e𝑒eitalic_e, the argument of periastron ω𝜔\omegaitalic_ω, the impact parameter b𝑏bitalic_b, the stellar radius Rsubscript𝑅R_{\star}italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT and the stellar mass Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT. These parameters were then input into light curve models created with Starry, alongside parameter Rpsubscript𝑅𝑝R_{p}italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (planetary radius), hyperparameter texpsubscript𝑡expt_{\mathrm{exp}}italic_t start_POSTSUBSCRIPT roman_exp end_POSTSUBSCRIPT (the exposure time of the instrument) and the time series of the data t𝑡titalic_t. The RV model was computed from the Keplerian orbit model by adding the semi-amplitude K𝐾Kitalic_K of the RV signal as a parameter. Based on the analysis described in the previous section, we modeled only one planet per system and found no need to include a Gaussian process as a model for the noise given that periodogram analysis of RV and LC data showed no indication of stellar activity or signals from stellar rotation at sufficiently significant frequencies. Instead, we assumed the errors are normal and included an additional white noise term for every instrument, which was added as a free nonnegative jitter term in quadrature with the reported errors.

All prior distributions set on the parameters fit in this model are given in tables 9 and 10. We put Gaussian priors on the stellar radius and mass informed by the values reported in section 3.1.3. We used values given by the QLP detection for the epoch and period to inform our Gaussian priors for log-P𝑃Pitalic_P and t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in both planets, and depth values from the same pipeline to inform the Gaussian prior on log-k𝑘kitalic_k, where k𝑘kitalic_k is the planet-to-star radius ratio k=RpR𝑘subscript𝑅𝑝subscript𝑅k=\frac{R_{p}}{R_{\star}}italic_k = divide start_ARG italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG start_ARG italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG, with a mean of 0.5log0.50.5\log0.5 roman_log-depth. We inflated the widths of these priors to σ=1𝜎1\sigma=1italic_σ = 1. Because it is well known that sampling directly the eccentricity and the argument of periastron can be problematic for most MCMC samplers (Parmentier & Crossfield, 2018), we sampled for esinω𝑒𝜔\sqrt{e}\sin\omegasquare-root start_ARG italic_e end_ARG roman_sin italic_ω and ecosω𝑒𝜔\sqrt{e}\cos\omegasquare-root start_ARG italic_e end_ARG roman_cos italic_ω with a uniform prior within a unit disk. This leads to a uniform prior on e𝑒eitalic_e as noted by Anderson et al. (2011). For the impact parameter we chose a uniform prior between 0 and 1+k1𝑘1+k1 + italic_k. For each lightcurve, we put a Gaussian prior on the transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (the light curve flux level out of transit) with a mean of 1 and standard deviation of 0.1. For the jitter parameter of each instrument we used a wide Gaussian prior, the mean of which was the log of the error median on each light curve. We used independent limb-darkening parameters for each filter, parameterised following Kipping (2013). For each spectroscopic instrument, we introduced a constant offset term to the reported RVs. We also sampled for a linear and quadratic trend in the RV data. We used uniform priors for log-K𝐾Kitalic_K and for the log-jitter of each RV instrument.

We added one dilution factor for each TESS sector as in Almenara et al. (2022).101010For the light curve model, we used M=f0(ftT+fc)/(ft+fc)𝑀subscript𝑓0subscript𝑓𝑡𝑇subscript𝑓𝑐subscript𝑓𝑡subscript𝑓𝑐M=f_{0}(f_{t}T+f_{c})/(f_{t}+f_{c})italic_M = italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_f start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_T + italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) / ( italic_f start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ), with M𝑀Mitalic_M being the model to be compared with the observations, f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT being the transit normalisation factor, ftsubscript𝑓𝑡f_{t}italic_f start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT being the flux of the target star in the aperture, T𝑇Titalic_T being the transit model, and fcsubscript𝑓𝑐f_{c}italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT being the flux of the contaminant star in the aperture. The dilution factor is defined as d=fc/(ft+fc)𝑑subscript𝑓𝑐subscript𝑓𝑡subscript𝑓𝑐d=f_{c}/(f_{t}+f_{c})italic_d = italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / ( italic_f start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ). Here, we use this parameter to complement the contamination correction carried out by the pipelines, so we refer to it dilution correction factor d𝑑ditalic_d. We used a Gaussian prior for d𝑑ditalic_d with a mean of 0 and standard deviation of 0.2. In this way, positive dilution correction values would correspond to a pipeline undercorrection, while negative values indicate an overcorrection. For the ground-based light curves we fixed the dilution correction factor to zero, thus assuming that these light curves are not contaminated at all.

We used exoplanet to maximise the log probability of the model. Two data points from TESS sector 28 were sigma-clipped during the optimization. The fit values obtained were then used as the starting point of the PyMC3 sampler, which draws samples from the posterior using a variant of Hamiltonian Monte Carlo, the No-U-Turn Sampler (NUTS). We allowed for 10000 burn-in samples which were discarded, and then 30000 steps with 10 chains. The MCMC chains for all model parameteres have Gelman-Rubin statistics (rank normalized split-R^^𝑅\hat{R}over^ start_ARG italic_R end_ARG) (Vehtari et al., 2021) very close to unity (<1.01absent1.01<1.01< 1.01), effective sample sizes (bulk-ESS) larger than 9100 and tail-ESS larger than 4600, so we have a good level of confidence that the chains mixed well and we have a stable estimate of uncertainty. The median model and the interval between the 16th and 84th percentiles are shown in figures 812. We present the parameters’ marginal posterior sample median for the TOI-2374 and TOI-3071 systems from the joint fits in table 3.

4 Results & discussion

Table 3: Planetary and system parameters of TOI-2374 b and TOI-3071 b. The values given are the medians, 16th and 84th percentiles of the MCMC-derived marginalized posterior distribution. Further parameters from the joint fit model can be found in tables 9 and 10.
Parameter (unit) Value Source
TOI-2374 b
Period P𝑃Pitalic_P (days) 4.31361±0.00001plus-or-minus4.313610.000014.31361\pm 0.000014.31361 ± 0.00001 Joint fit
Full transit duration Tdursubscript𝑇𝑑𝑢𝑟T_{dur}italic_T start_POSTSUBSCRIPT italic_d italic_u italic_r end_POSTSUBSCRIPT (hours) 1.38±0.07plus-or-minus1.380.071.38\pm 0.071.38 ± 0.07 Joint fit (derived)
Epoch t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (BJD-2457000) 1326.564±0.002plus-or-minus1326.5640.0021326.564\pm 0.0021326.564 ± 0.002 Joint fit
Radius Rpsubscript𝑅𝑝R_{p}italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (R) 6.81±0.30plus-or-minus6.810.306.81\pm 0.306.81 ± 0.30 Joint fit
Planet-to-star radius ratio Rp/Rsubscript𝑅𝑝subscript𝑅R_{p}/R_{\star}italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT - 0.091±0.002plus-or-minus0.0910.0020.091\pm 0.0020.091 ± 0.002 Joint fit
Impact parameter b𝑏bitalic_b - 0.6510.063+0.047subscriptsuperscript0.6510.0470.0630.651^{+0.047}_{-0.063}0.651 start_POSTSUPERSCRIPT + 0.047 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.063 end_POSTSUBSCRIPT Joint fit
Inclination i𝑖iitalic_i () 87.50.2+0.3subscriptsuperscript87.50.30.287.5^{+0.3}_{-0.2}87.5 start_POSTSUPERSCRIPT + 0.3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.2 end_POSTSUBSCRIPT Joint fit
Eccentricity e𝑒eitalic_e - 0.13absentsuperscript0.13\leq 0.13^{*}≤ 0.13 start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT Joint fit
Radial velocity semi-amplitude K𝐾Kitalic_K (ms-1) 27±2plus-or-minus27227\pm 227 ± 2 Joint fit
esinω𝑒𝜔\sqrt{e}\sin{\omega}square-root start_ARG italic_e end_ARG roman_sin italic_ω - 0.200.09+0.06subscriptsuperscript0.200.060.090.20^{+0.06}_{-0.09}0.20 start_POSTSUPERSCRIPT + 0.06 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.09 end_POSTSUBSCRIPT Joint fit
ecosω𝑒𝜔\sqrt{e}\cos{\omega}square-root start_ARG italic_e end_ARG roman_cos italic_ω - 0.100.17+0.13subscriptsuperscript0.100.130.170.10^{+0.13}_{-0.17}0.10 start_POSTSUPERSCRIPT + 0.13 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.17 end_POSTSUBSCRIPT Joint fit
Minimum mass Mpsinisubscript𝑀𝑝𝑖M_{p}\sin{i}italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_sin italic_i (M) 56.594.26+3.49subscriptsuperscript56.593.494.2656.59^{+3.49}_{-4.26}56.59 start_POSTSUPERSCRIPT + 3.49 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 4.26 end_POSTSUBSCRIPT Joint fit (derived)
Mass Mpsubscript𝑀𝑝M_{p}italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (M) 56.644.27+3.49subscriptsuperscript56.643.494.2756.64^{+3.49}_{-4.27}56.64 start_POSTSUPERSCRIPT + 3.49 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 4.27 end_POSTSUBSCRIPT Joint fit (derived)
Bulk density ρ𝜌\rhoitalic_ρ (g cm-3) 0.980.13+0.15subscriptsuperscript0.980.150.130.98^{+0.15}_{-0.13}0.98 start_POSTSUPERSCRIPT + 0.15 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.13 end_POSTSUBSCRIPT Joint fit (derived)
Semi-major axis a𝑎aitalic_a (AU) 0.0471±0.0002plus-or-minus0.04710.00020.0471\pm 0.00020.0471 ± 0.0002 Joint fit (derived)
System scale a/R𝑎subscript𝑅a/R_{\star}italic_a / italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT - 14.7±0.4plus-or-minus14.70.414.7\pm 0.414.7 ± 0.4 Joint fit (derived)
Equilibrium temperature Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT (albedo ζ=0.0𝜁0.0\zeta=0.0italic_ζ = 0.0) (K) 885±22plus-or-minus88522885\pm 22885 ± 22 Joint fit (derived)
Equilibrium temperature Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT (ζ=0.3𝜁0.3\zeta=0.3italic_ζ = 0.3) (K) 810±20plus-or-minus81020810\pm 20810 ± 20 Joint fit (derived)
Equilibrium temperature Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT (ζ=0.5𝜁0.5\zeta=0.5italic_ζ = 0.5) (K) 744±18plus-or-minus74418744\pm 18744 ± 18 Joint fit (derived)
TOI-3071 b
Period P𝑃Pitalic_P (days) 1.266938±0.000002plus-or-minus1.2669380.0000021.266938\pm 0.0000021.266938 ± 0.000002 Joint fit
Full transit duration Tdursubscript𝑇𝑑𝑢𝑟T_{dur}italic_T start_POSTSUBSCRIPT italic_d italic_u italic_r end_POSTSUBSCRIPT (hours) 1.850.07+0.08subscriptsuperscript1.850.080.071.85^{+0.08}_{-0.07}1.85 start_POSTSUPERSCRIPT + 0.08 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.07 end_POSTSUBSCRIPT Joint fit (derived)
Epoch t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (BJD-2457000) 1594.608±0.002plus-or-minus1594.6080.0021594.608\pm 0.0021594.608 ± 0.002 Joint fit
Radius Rpsubscript𝑅𝑝R_{p}italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (R) 7.160.51+0.57subscriptsuperscript7.160.570.517.16^{+0.57}_{-0.51}7.16 start_POSTSUPERSCRIPT + 0.57 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.51 end_POSTSUBSCRIPT Joint fit
Planet-to-star radius ratio Rp/Rsubscript𝑅𝑝subscript𝑅R_{p}/R_{\star}italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT - 0.0500.003+0.003subscriptsuperscript0.0500.0030.0030.050^{+0.003}_{-0.003}0.050 start_POSTSUPERSCRIPT + 0.003 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.003 end_POSTSUBSCRIPT Joint fit
Impact parameter b𝑏bitalic_b - 0.560.08+0.07subscriptsuperscript0.560.070.080.56^{+0.07}_{-0.08}0.56 start_POSTSUPERSCRIPT + 0.07 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.08 end_POSTSUBSCRIPT Joint fit
Inclination i𝑖iitalic_i () 82.11.1+1.3subscriptsuperscript82.11.31.182.1^{+1.3}_{-1.1}82.1 start_POSTSUPERSCRIPT + 1.3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1.1 end_POSTSUBSCRIPT Joint fit
Eccentricity e𝑒eitalic_e - 0.09absentsuperscript0.09\leq 0.09^{*}≤ 0.09 start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT Joint fit
Radial velocity semi-amplitude K𝐾Kitalic_K (ms-1) 33.7±1.7plus-or-minus33.71.733.7\pm 1.733.7 ± 1.7 Joint fit
esinω𝑒𝜔\sqrt{e}\sin{\omega}square-root start_ARG italic_e end_ARG roman_sin italic_ω - 0.110.07+0.10subscriptsuperscript0.110.100.07-0.11^{+0.10}_{-0.07}- 0.11 start_POSTSUPERSCRIPT + 0.10 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.07 end_POSTSUBSCRIPT Joint fit
ecosω𝑒𝜔\sqrt{e}\cos{\omega}square-root start_ARG italic_e end_ARG roman_cos italic_ω - 0.070.14+0.16subscriptsuperscript0.070.160.140.07^{+0.16}_{-0.14}0.07 start_POSTSUPERSCRIPT + 0.16 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.14 end_POSTSUBSCRIPT Joint fit
Minimum mass Mpsinisubscript𝑀𝑝𝑖M_{p}\sin{i}italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_sin italic_i (M) 67.6±3.5plus-or-minus67.63.567.6\pm 3.567.6 ± 3.5 Joint fit (derived)
Mass Mpsubscript𝑀𝑝M_{p}italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (M) 68.2±3.5plus-or-minus68.23.568.2\pm 3.568.2 ± 3.5 Joint fit (derived)
Bulk density ρ𝜌\rhoitalic_ρ (g cm-3) 1.020.22+0.26subscriptsuperscript1.020.260.221.02^{+0.26}_{-0.22}1.02 start_POSTSUPERSCRIPT + 0.26 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.22 end_POSTSUBSCRIPT Joint fit (derived)
Semi-major axis a𝑎aitalic_a (AU) 0.0249±0.0001plus-or-minus0.02490.00010.0249\pm 0.00010.0249 ± 0.0001 Joint fit (derived)
System scale a/R𝑎subscript𝑅a/R_{\star}italic_a / italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT - 4.11±0.13plus-or-minus4.110.134.11\pm 0.134.11 ± 0.13 Joint fit (derived)
Equilibrium temperature Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT (albedo ζ=0.0𝜁0.0\zeta=0.0italic_ζ = 0.0) (K) 2155±40plus-or-minus2155402155\pm 402155 ± 40 Joint fit (derived)
Equilibrium temperature Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT (ζ=0.3𝜁0.3\zeta=0.3italic_ζ = 0.3) (K) 1971±37plus-or-minus1971371971\pm 371971 ± 37 Joint fit (derived)
Equilibrium temperature Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT (ζ=0.5𝜁0.5\zeta=0.5italic_ζ = 0.5) (K) 1812±34plus-or-minus1812341812\pm 341812 ± 34 Joint fit (derived)
  • Tdursubscript𝑇𝑑𝑢𝑟T_{dur}italic_T start_POSTSUBSCRIPT italic_d italic_u italic_r end_POSTSUBSCRIPT is approximated by Pπsin1[Ra(1+k)2b2sini]1e21+esinω𝑃𝜋superscript1subscript𝑅𝑎superscript1𝑘2superscript𝑏2𝑖1superscript𝑒21𝑒𝜔\frac{P}{\pi}\sin^{-1}\left[\frac{R_{\star}}{a}\frac{\sqrt{(1+k)^{2}-b^{2}}}{% \sin{i}}\right]\frac{\sqrt{1-e^{2}}}{1+e\sin{\omega}}divide start_ARG italic_P end_ARG start_ARG italic_π end_ARG roman_sin start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ divide start_ARG italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG italic_a end_ARG divide start_ARG square-root start_ARG ( 1 + italic_k ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG roman_sin italic_i end_ARG ] divide start_ARG square-root start_ARG 1 - italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG 1 + italic_e roman_sin italic_ω end_ARG where k=Rp/R𝑘subscript𝑅𝑝subscript𝑅k=R_{p}/R_{\star}italic_k = italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT (Winn, 2014).

  • Computed as the 95%percent9595\%95 % HDI𝐻𝐷𝐼HDIitalic_H italic_D italic_I upper bound for the posterior of e𝑒eitalic_e. Because the orbit is essentially circular, ω𝜔\omegaitalic_ω is not well-determined.

The combined analysis of high resolution spectroscopy and space- and ground-based photometry allowed the determination of the masses and radii of planetary companions for the two target systems. We determined that both planets are sub-Saturns with mass and radius precision of 47%4percent74-7\%4 - 7 % (see Table 3). From this we inferred bulk densities of 0.980.13+0.15subscriptsuperscript0.980.150.130.98^{+0.15}_{-0.13}0.98 start_POSTSUPERSCRIPT + 0.15 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.13 end_POSTSUBSCRIPT g cm-3 for TOI-2374 b and 1.020.22+0.26subscriptsuperscript1.020.260.221.02^{+0.26}_{-0.22}1.02 start_POSTSUPERSCRIPT + 0.26 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.22 end_POSTSUBSCRIPT g cm-3 for TOI-3071 b. These parameters place both planets within in the Neptunian desert, with TOI-3071 b being much deeper in than TOI-2374 b (Fig. 13). We found the eccentricity of both targets to be consistent with 0. The 95%percent9595\%95 % highest density interval (HDI) for the eccentricity ranges from 0 to 0.13 and 0.09 for TOI-2374 b and TOI-3071 b, respectively. We did not find any statistically significant long-term trends in the RV measurements of either star.

We note that the MCMC posterior distributions for the TESS light curves dilution correction factors are narrower than the priors (see figures 19 and 20 and tables 9 and 10 in the Appendix). For TOI-2374, the posterior modes of the QLP dilution correction factors are systematically negative, hinting to an overestimation of the dilution factor estimated by this pipeline. On the other hand, in TOI-3071, the posterior mean of the QLP dilution correction factor is positive, suggesting that for this star, the correction procedure underestimates the dilution. Finally, our model also hints to an overestimation of the dilution factor in the SPOC lightcurve of TOI-3071, as the posterior mean of d𝑑ditalic_d is negative. Exploring the reasons of these behaviours is outside the scope of this article. Additionaly, in all cases the 83%percent\%% HDI contains the null correction factor value. In any case, our model accounts for the uncertainty in the dilution factors estimated by the pipelines on the determination of the planetary radii.

Refer to caption
Figure 13: TOI-2374 b (black marker) and TOI-3071 b (red marker) in the context of the Neptunian Desert regions in Mass vs Period (top panel) and Radius vs Period (bottom panel). The dashed blue lines show the delineation of the Neptunian desert from Mazeh et al. (2016), whereas the horizontal dotted blue lines show the updated lower limits for periods 2absent2\leq 2≤ 2 days from Deeg et al. (2023). Known planets were sourced from the NASA exoplanet archive (https://exoplanetarchive.ipac.caltech.edu/) on 28 May 2024.

4.1 Comparison to other exoplanets

In order to understand how TOI-2374 b and TOI-3071 b fit into the landscape of known planets, we compare them to planets of similar size, mass, and period (P<20𝑃20P<20italic_P < 20 d𝑑ditalic_d; 40<MP<10040subscript𝑀𝑃10040<M_{P}<10040 < italic_M start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT < 100 Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT; ρ<2𝜌2\rho~{}<~{}2italic_ρ < 2 gcm3𝑔𝑐superscript𝑚3g~{}cm^{-3}italic_g italic_c italic_m start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT) with mass precision better than 30%percent3030\%30 % and radius precision better than 20%percent2020\%20 % (top panel of Fig. 14). This includes all “hot” giant planets below Saturn’s mass. We use the parameters from the planetary systems table on the NASA Exoplanet Archive 111111https://exoplanetarchive.ipac.caltech.edu/ accessed on 28 May 2024. The empirical R-M relation for volatile-rich planets (ρ<3.3𝜌3.3\rho<3.3italic_ρ < 3.3 gcm3𝑔𝑐superscript𝑚3g~{}cm^{-3}italic_g italic_c italic_m start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT) found by Otegi et al. (2020) is also shown in the diagram. The bottom panel of Fig. 14 shows the equilibrium temperature (Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT) vs radius plane for the same subset of exoplanets, but extending the mass range to 10<MP<30010subscript𝑀𝑃30010<M_{P}<30010 < italic_M start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT < 300 Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT. Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT is calculated from the planetary orbital semi-major axes a𝑎aitalic_a and the host-star effective temperature Tsubscript𝑇T_{\star}italic_T start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT and radius Rsubscript𝑅R_{\star}italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT assuming full day-night heat distribution, according to

Teq=TR2a(1ζ)1/4subscript𝑇eqsubscript𝑇subscript𝑅2𝑎superscript1𝜁14T_{\rm eq}=T_{\star}\sqrt{\frac{R_{\star}}{2a}}(1-\zeta)^{1/4}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT = italic_T start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT square-root start_ARG divide start_ARG italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_a end_ARG end_ARG ( 1 - italic_ζ ) start_POSTSUPERSCRIPT 1 / 4 end_POSTSUPERSCRIPT (1)

where ζ𝜁\zetaitalic_ζ is the Bond albedo of the planet. For this analysis we assumed ζ=0.5𝜁0.5\zeta=0.5italic_ζ = 0.5, similar to that of Jupiter (Li et al., 2018), to calculate Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT for TOI-2374 b, TOI-3071 b and every other planet in the dataset (even if they already had a reported Teqsubscript𝑇eqT_{\rm eq}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT).

Refer to caption
Figure 14: Top: mass-radius diagram of the confirmed planet population with P<20𝑃20P<20italic_P < 20 d𝑑ditalic_d, masses 40<MP<10040subscript𝑀𝑃10040<M_{P}<10040 < italic_M start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT < 100 Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT and bulk densities ρ<2𝜌2\rho<2italic_ρ < 2 gcm3𝑔𝑐superscript𝑚3g~{}cm^{-3}italic_g italic_c italic_m start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, with radius determined to better than 20%percent2020\%20 % precision and masses determined to better than 30%percent3030\%30 % precision. The dashed lines show constant density values. For comparison, the position of Saturn in the plot is represented with a white diamond marker. The bulk densities for each planet are color coded. The R-M empirical relation found by Otegi et al. (2020) is shown by a red dashed line with the 1σ1𝜎1\sigma1 italic_σ regions colored. Bottom: equilibrium temperatures and radii of the same set of exoplanets, extending the mass range to 10<MP<30010subscript𝑀𝑃30010<M_{P}<30010 < italic_M start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT < 300 Msubscript𝑀direct-sumM_{\oplus}italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT. The bulk densities for each planet are color coded and the markers are sized according to their mass.

Both planets are hot and highly irradiated, with Teq744subscript𝑇eq744T_{\rm eq}\approx 744italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT ≈ 744 K for TOI-2374 b and Teq1812subscript𝑇eq1812T_{\rm eq}\approx 1812italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT ≈ 1812 K for TOI-3071 b. TOI-2374 b is slightly below the usual inflated hot-Jupiter cutoff of F2×108subscript𝐹2superscript108F_{*}\approx 2\times 10^{8}italic_F start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≈ 2 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT erg/s/cm2 (Miller & Fortney, 2011), and it is therefore not expected to be anomalously inflated. On the other hand, TOI-3071 b is well inside the hot-Jupiter instellation regime. However, previous studies suggest that planets with masses below about 0.4MJ0.4subscript𝑀𝐽0.4M_{J}0.4 italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT show a weaker response to high instellation fluxes compared to more massive planets, and are less likely to be strongly inflated (Thorngren & Fortney, 2018; Sestovic et al., 2018). Therefore, standard evolution models of giant planets may reflect the radius more accurately than for higher mass planets, for which the inflation is known to be underestimated. From Fig. 14 it is clear that TOI-3071 b has a similar radius to previously detected giant exoplanets with similar masses, despite having a much higher equilibrium temperature. Besides hydrogen and helium, the planet is therefore expected to contain a large amount of heavier elements.

4.2 Inferred heavy-element masses

Both planets have available masses, radii and ages, which make them viable targets for the estimation of their heavy-element masses (e.g., Miller & Fortney, 2011; Thorngren et al., 2016; Müller & Helled, 2023b). The bulk heavy-element mass of a giant planet is a key property since it can be used to test planet formation models and provide additional constraints on the possible formation pathways (e.g., Guillot et al., 2006; Johansen & Lambrechts, 2017; Hasegawa et al., 2018). Unfortunately, it cannot be measured directly and must be inferred from a combination of measurements and giant-planet evolution models. To estimate the composition of the two planets, we modeled their evolution with a modified version of the stellar evolution code Modules for Experiments in Stellar Astrophysics (MESA; Paxton et al. (2011); Paxton et al. (2013, 2015, 2018, 2019); Jermyn et al. (2023)). Müller et al. (2020b, a) modified the equation of state of MESA to make it suitable to model giant planets that contain a significant amount of heavy elements. Here, we used a recently updated version (Müller & Helled, 2024), which uses a new hydrogen-helium equation of state that includes non-ideal interactions in the hydrogen-helium mixture (Chabrier & Debras, 2021). These effects have been shown to lead to a significant change in the cooling of planets (Howard & Guillot, 2023). We assumed a core-envelope structure and a 50-50 water-rock composition for the heavy elements. The radiative opacity was from Freedman et al. (2014), and to account for the instellation fluxes we used a gray atmospheric model developed for irradiated planets (Guillot, 2010; Parmentier & Guillot, 2014).

We first calculated the planetary evolution assuming that they have the same composition as their stars. Both planets are much smaller compared to these predictions, demonstrating that they must have super-stellar metallicities (see solid lines in Fig. 21 in Appendix).

To quantify the metal enrichment of the planets, we used the standard Monte-Carlo approach (see, e.g., Müller & Helled, 2023a, for a review): sample planets were created by drawing from the observation prior distributions for the mass, radius, and stellar age. For the mass and radius, we used normal distributions informed by the parameters’ marginal posterior sample median and standard deviation from the joint fit (Sect. 3.3), and for the age we used a uniform distribution based on the estimates from the stellar parameters analysis (Sect. 3.1.3). The evolution models were then used to calculate the cooling for different heavy-element masses Mzsubscript𝑀𝑧M_{z}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT to find the one that agrees with the observations. This process is repeated until there is enough data to estimate the posterior distribution for the heavy-element mass. For TOI-3071b we used two different stellar age priors (from chemical clocks and gyrochronology) and then combined their posteriors for the final result assuming that the priors are equally likely. The MCMC approach required the calculation of a large number of evolution models. Since this would have been too slow, we followed a similar approach as outlined in Müller & Helled (2021). For each planet, we calculated a grid of evolution models (in the planet mass-metallicity space) and then used 2d piecewise monotonic cubic interpolation to generate new cooling tracks.

The resulting posterior distributions of the inferred heavy-element masses are shown in Fig. 15. We find heavy-element masses (mass fractions) of Mz=33.3±3.8Msubscript𝑀𝑧plus-or-minus33.33.8subscript𝑀direct-sumM_{z}=33.3\pm 3.8M_{\oplus}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 33.3 ± 3.8 italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT (Z=0.59±0.05𝑍plus-or-minus0.590.05Z=0.59\pm 0.05italic_Z = 0.59 ± 0.05) for TOI-2374 b and Mz=45.1±5.5Msubscript𝑀𝑧plus-or-minus45.15.5subscript𝑀direct-sumM_{z}=45.1\pm 5.5M_{\oplus}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 45.1 ± 5.5 italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT (Z=0.66±0.07)Z=0.66\pm 0.07)italic_Z = 0.66 ± 0.07 ) for TOI-3071 b. For comparison, Saturn, the most similar planet in the solar system, has an estimated bulk heavy-element mass (mass fraction) of Mz30Msimilar-to-or-equalssubscript𝑀𝑧30subscript𝑀direct-sumM_{z}\simeq 30M_{\oplus}italic_M start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≃ 30 italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT (Z0.3similar-to-or-equals𝑍0.3Z\simeq 0.3italic_Z ≃ 0.3). The exact value is unknown and depends on several assumptions when modeling Saturn’s interior. Both targets are metal-rich, with TOI-3071b being more enriched both in absolute and relative terms (see Figure 22 for posterior distributions of the heavy-element masses in mass fractions). TOI-3071b in particular has an extremely high bulk metallicity and is about 60-70% heavy elements (water and rocks).

The fact that these planets have such high metallicities challenges standard formation models which generally tend to predict lower heavy-element masses (Pollack et al., 1996; Helled et al., 2014; Johansen & Lambrechts, 2017; Bitsch et al., 2018). Interestingly, both planets correspond to masses below Saturn’s mass, which was recently suggested to be the transition mass to become giant planets (Helled, 2023). In that view, the formation of a giant planet can last for a few million years where gas accretion is being delayed due to continuous accretion of planetesimals (Alibert et al., 2018; Helled et al., 2022b). Such a formation scenario also suggests that planets below about Saturn’s mass are expected to be metal-rich in composition, consistent with our estimates for TOI-2374 b and TOI-3071 b. An alternative explanation is a post-formation enrichment due to planetesimal accretion or collisions between planetary embryos (Mousis et al., 2009; Shibata et al., 2020; Ginzburg & Chiang, 2020). However, such collisions have difficulty enriching planets with tens of Earth masses of heavy elements.

Based on the models from Santos et al. (2015, 2017); Adibekyan et al. (2021), we used the stellar abundances to calculate the summed mass fraction of all heavy elements available as planetary building blocks. The values are ZTOI2374=0.018±0.03subscript𝑍𝑇𝑂𝐼2374plus-or-minus0.0180.03Z_{TOI-2374}=0.018\pm 0.03italic_Z start_POSTSUBSCRIPT italic_T italic_O italic_I - 2374 end_POSTSUBSCRIPT = 0.018 ± 0.03 and ZTOI3071b=0.024±0.03subscript𝑍𝑇𝑂𝐼3071𝑏plus-or-minus0.0240.03Z_{TOI-3071b}=0.024\pm 0.03italic_Z start_POSTSUBSCRIPT italic_T italic_O italic_I - 3071 italic_b end_POSTSUBSCRIPT = 0.024 ± 0.03, respectively. For comparison, the same models yield Z=0.013±0.01subscript𝑍direct-productplus-or-minus0.0130.01Z_{\odot}=0.013\pm 0.01italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.013 ± 0.01 for solar abundances. Both mass fractions are above the solar value, which hints at a possible connection between the stellar abundances and the high inferred bulk metallicities.

Finally, we note that measurements of the atmospheric chemical composition could be used to further constrain the planetary formation, evolution, and interior (e.g., Burrows, 2014; Thorngren & Fortney, 2019; Helled et al., 2022a).

Refer to caption
Figure 15: Inferred heavy-element mass posteriors for the two planets. For TOI-3071b, the two curves show the results for the two different stellar age estimates (chemical clocks and gyrochronology), and the filled area is their combined posterior. Both planets are heavy-element rich, with a bulk heavy-element fraction of Z0.6greater-than-or-equivalent-to𝑍0.6Z\gtrsim 0.6italic_Z ≳ 0.6.

4.3 Evaporation history

4.3.1 Internal structure

To model the evaporation histories of the two planets, we assumed a simpler two-layer internal structure with a rocky core surrounded by a pure H/He envelope. This model can be described with four quantities: the rocky core mass Mcorecore{}_{\text{core}}start_FLOATSUBSCRIPT core end_FLOATSUBSCRIPT and core radius Rcorecore{}_{\text{core}}start_FLOATSUBSCRIPT core end_FLOATSUBSCRIPT, and the gaseous envelope thickness Renvenv{}_{\text{env}}start_FLOATSUBSCRIPT env end_FLOATSUBSCRIPT and envelope mass fraction fenvsubscriptfenv\text{f}_{\text{env}}f start_POSTSUBSCRIPT env end_POSTSUBSCRIPT, defined as the ratio of envelope mass to planet mass Menv/MpsubscriptMenvsubscriptMp\text{M}_{\text{env}}/\text{M}_{\text{p}}M start_POSTSUBSCRIPT env end_POSTSUBSCRIPT / M start_POSTSUBSCRIPT p end_POSTSUBSCRIPT. In order to determine these parameters, we adopted the empirical mass-radius relation by Otegi et al. (2020) for the rocky core, and the envelope structure model by Chen & Rogers (2016) for the gaseous atmosphere, which is based on MESA simulations. Uisng this model, we thus determined envelope mass fractions of 38±5plus-or-minus38538\pm 538 ± 5 % for TOI-2374 b and 26±6plus-or-minus26626\pm 626 ± 6 % for TOI-3071 b. These correspond to core (heavy-element) masses of M=core34.7±3.7{}_{\text{core}}=34.7\pm 3.7start_FLOATSUBSCRIPT core end_FLOATSUBSCRIPT = 34.7 ± 3.7 for TOI-2374 b, and M=core50.5±4.9{}_{\text{core}}=50.5\pm 4.9start_FLOATSUBSCRIPT core end_FLOATSUBSCRIPT = 50.5 ± 4.9 for TOI-3071 b, both consistent with the heavy-element masses determined in Sect. 4.2 within 1σ1𝜎1\sigma1 italic_σ.

Refer to caption
Figure 16: Left panel: spin period evolution of TOI-2374 using the models by Johnstone et al. (2021). The model mean is shown as a solid black line, and the 2σ2𝜎2\sigma2 italic_σ spread as a grey shaded region. TOI-2374 is shown as a blue circle. Right panel: evolution models of the X-ray luminosity following the left panel. The current X-ray luminosity of the star predicted by the model is shown as a blue circle.
Refer to caption
Figure 17: Left panel: spin period evolution of TOI-3071 using the models by Johnstone et al. (2021), following Figure 16. Due to the disagreement between the age estimated from gyrochronology (blue circle) and the age from the chemical clocks and PARAM isochrones (red circle), we plotted an additional spin evolution model fitted to the age from chemical clocks (red line). Right panel: X-ray evolution of TOI-3071 using the models by Johnstone et al. (2021), following Figure 16. The X-ray evolution model fitted to the age from the chemical clocks is shown as a red line.

4.3.2 Evolution modelling

Simulating the evolution of a planet’s atmosphere under photoevaporation requires knowledge of its X-ray irradiation history. We can estimate the X-ray and EUV emission (together, XUV) of a star from its spin period, as the two are related via rotation–activity relations (e.g. Wright et al., 2011; Pizzolato et al., 2003). We adopted the stellar rotation evolution models by Johnstone et al. (2021), which simulate the evolution of the angular momentum of stars constrained by observations of young star clusters. Moreover, to estimate the EUV component, we adopted the scaling relation by King & Wheatley (2021), which are based on Solar TIMED/SEE data. We plot the evolution of the spin period and corresponding XUV emission of TOI-2374 and TOI-3071 on Figures 16 and 17, respectively.

Regarding TOI-2374, we found the rotational models by Johnstone et al. (2021) are consistent with the measured age and spin period of the star.

For TOI-3071, we considered the two age estimates separately: an age of 1 Gyr from the chemical clocks and PARAM isochrones, and an age of 5 Gyr from gyrochronology (see Sect. 3.1.3).

We found the choice of age had a strong impact on the X-ray emission history of the star. The younger age is more consistent with a short spin period of 2-4 d, whereas the older age from gyrochronology is more consistent with a slower spin period of 20-30 d, as the star has had more time to spin down. The spin period of the star, however, is constrained to 26 d, and so adopting the younger age would imply that the star must be unusually slowly rotating and X-ray faint (Fig. 17, red line).

For that reason, we adopted two scenarios for the X-ray emission history of TOI-3071: (1) the high activity scenario where we adopted the predicted evolution from the models by Johnstone et al. (2021), consistent with the older gyrochronological age (see Fig. 17, black line); and (2) the low activity scenario, where we adopted a fainter emission history model consistent with the younger age from the 3D chemical clocks (Fig. 17, red line).

Refer to caption
Figure 18: Left panel: Past and future evolution of the atmosphere of TOI-2374 b under photoevaporation. The solid blue line shows the expected evolution of the planet, whereas the shaded region represents possible evaporation histories taking into account the uncertainties in the planet’s mass, radius, and XUV irradiation. The planet is plotted as a white circle. Right panel: Plot of the past and future evolution of the atmosphere of TOI-3071 b under photoevaporation, following the left panel. The area shaded in blue shows evolution under an X-ray irradiation history consistent with the stellar age derived from gyrochronology (5 Gyr), whereas the area shaded in red shows evolution under a fainter irradiation history consistent with the stellar age derived from chemical clocks (1 Gyr).

We then simulated the past and future evaporation histories of the planets with the photoevolver code (Fernández Fernández et al., 2023). The simulation is built upon three ingredients: (1) the XUV emission history of the star, which determines the X-ray irradiation on the planet, (2) a mass loss formulation, which determines the amount of gas lost due to XUV irradiation, and (3) an envelope structure model, which relates the mass of the gaseous atmosphere to its thickness. Therefore, on each time step, the X-ray luminosity of the star at that age is used to calculate the mass loss rate, which is then used to recalculate the mass and size of the planet.

We adopted the models by Johnstone et al. (2021) for the stellar XUV emission history (described above), and the model by Chen & Rogers (2016) for the structure of the gaseous atmosphere, described in Section 4.3.1. Regarding the mass loss formulation, we adopted the model by Kubyshkina et al. (2018). We evolved the planets’ atmospheres using the 4th order Runge-Kutta algorithm (Dormand & Prince, 1980) as the integration method with a variable time step no larger than 1 Myr. We ran the simulation backwards to the age of 10 Myr and forwards to 10 Gyr. In the case of the F-type TOI-3071, we stop the simulation at 6 Gyr, the approximate age at which the star leaves the main sequence (Choi et al., 2016).

Moreover, for each of the two planets, we also explored a range of possible evaporation histories taking into account the uncertainties on the planet’s mass, radius, and XUV irradiation. To maximise the evaporation rate, we picked a radius on the higher end of its 1σ1𝜎1\sigma1 italic_σ uncertainty and a mass on the lower end of its 1σ1𝜎1\sigma1 italic_σ uncertainty, which minimises the density and maximises the escape rates. We also picked the highest XUV emission history allowed by the model’s 2σ2𝜎2\sigma2 italic_σ spread, as shown on Figures 16 and 17. Likewise, to minimise the evaporation rate, we picked the opposite parameters: values for the mass and radius within their 1σ1𝜎1\sigma1 italic_σ errors that maximise density under the faintest XUV irradiation history allowed by the model’s 2σ2𝜎2\sigma2 italic_σ spread.

4.3.3 Simulation results

The evaporation histories of TOI-2374 b and TOI-3071 b are plotted on Figure 18 (left and right panels, respectively).

We found that TOI-2374 b is stable against evaporation, as its envelope mass fraction changes only by a few percent across its lifetime. The evolution of its radius is thus largely driven by thermal contraction. We also constrained its current mass loss rates to 0.40.8×1080.40.8superscript1080.4-0.8\times 10^{8}0.4 - 0.8 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT g s-1.

On the other hand, TOI-3071 b is more susceptible to escape despite its higher mass. This is due to being twice as close to its star compared to TOI-2374 b, as well as the higher X-ray luminosity of its F-type host star. Under the higher X-ray irradiation history consistent with the gyrochronological age (5 Gyr), our simulations show that the planet is reaching the point of losing its atmosphere completely when its star starts evolving off the main sequence (see Fig. 18), with an envelope mass fraction of 10-20% at 6 Gyr down from 40-60% initially. Under this X-ray irradiation history, the planet currently experiences mass loss rates of 7.540.7×1087.540.7superscript1087.5-40.7\times 10^{8}7.5 - 40.7 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT g s-1.

However, under the fainter irradiation history consistent with the much younger age derived from chemical clocks (1 Gyr), TOI-3071 b experiences lower escape rates throughout its life and is stable against evaporation. We constrained its current mass loss rates to 0.92.6×1080.92.6superscript1080.9-2.6\times 10^{8}0.9 - 2.6 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT g s-1, one order of magnitude lower compared to the high activity scenario.

In Sect. 4.2, we presented more detailed internal structure models to estimate the bulk metallicity of the two planets. These models cannot predict the atmospheric composition, which can only be determined by observations. However, since both are metal-rich, they are also likely to have high-metallicity atmospheres. Such high-metallicity atmospheres experience much lower escape rates as heavier species require more energy to remove (Wilson et al., 2022; Owen & Jackson, 2012). Since we only considered a pure H/He composition in our atmospheric evolution models, our results in Fig. 18 constitute an upper bound to the evaporation histories of the planets.

For TOI-2374,b, the atmosphere is stable against evaporation throughout its life, even when assuming a pure hydrogen envelope. TOI-3071,b experiences moderate escape rates and is significantly affected by evaporation only as its host star transitions out of the main sequence, risking total atmospheric loss. However, heavy-mass planets like TOI-3071 b could present a metallicity gradient in their gaseous envelopes, with a relatively pure H/He upper atmosphere that progresses towards metal-rich layers deeper in the atmosphere. If the planet experienced stronger escape rates earlier in its life, either from photoevaporation or Roche-lobe overflow, the pure H/He upper layers of the envelope would be stripped off first, exposing the deeper metal-rich layers which would then curtail atmospheric escape (Owen & Murray-Clay, 2018; Hoyer et al., 2023).

Therefore, the presence of heavy species in the hydrogen-rich envelope of TOI-3071 b at the present time would lead to diminished escape rates, making its atmosphere more resistant to evaporation and thus likely stable. As a result, we find the two planets are not significantly affected by photoevaporation and are thus able to hold onto their gaseous envelopes throughout their lifetimes.

4.4 Prospects for atmospheric characterization

The transmission spectroscopy metric (TSM) is defined as

TSM=(scale factor)×Rp3TeqMpR2×10mJ/5𝑇𝑆𝑀(scale factor)superscriptsubscript𝑅𝑝3subscript𝑇eqsubscript𝑀𝑝superscriptsubscript𝑅2superscript10subscript𝑚𝐽5TSM=\text{(scale factor)}\times\frac{R_{p}^{3}T_{\rm eq}}{M_{p}R_{\star}^{2}}% \times 10^{-m_{J}/5}italic_T italic_S italic_M = (scale factor) × divide start_ARG italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG × 10 start_POSTSUPERSCRIPT - italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT / 5 end_POSTSUPERSCRIPT (2)

where mJsubscript𝑚𝐽m_{J}italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT is the apparent magnitude of the host star in the J band and the scale factor is 1.151.151.151.15 for 4<Rp<104subscript𝑅𝑝104<R_{p}<104 < italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT < 10 Rsubscript𝑅direct-sumR_{\oplus}italic_R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT (Kempton et al., 2018).

We calculated the targets’ TSM using the equilibrium temperatures for albedos of ζ=0.5𝜁0.5\zeta=0.5italic_ζ = 0.5 and obtained a value of 84±13plus-or-minus841384\pm 1384 ± 13 for TOI-2374 b and 36±9plus-or-minus36936\pm 936 ± 9 for TOI-3071 b. The former is slightly below the threshold of 90 set by Kempton et al. (2018) for sub-jovians to be considered high quality atmospheric characterization targets. However, if we assume lower albedos of 0.30.30.30.3 and 0.00.00.00.0, we obtain TSM values of 91±15plus-or-minus911591\pm 1591 ± 15 and 100±16plus-or-minus10016100\pm 16100 ± 16 for TOI-2374 b, respectively. For TOI-3071 b the TSM is always lower than 50. As mentioned above, atmospheric measurements of these targets could provide key information to further constrain the planetary formation, evolution and interior structure.

5 Conclusion

We have presented the discovery and characterisation of two new transiting hot Saturns orbiting the K-type and F-type stars TOI-2374 and TOI-3071, respectively. Our analysis includes photometry from TESS, follow up ground-based photometry and spectroscopy from HARPS and PFS. The combination of this data in a joint light curve and radial velocity model allowed the confirmation of the planets and the determination of their orbital and physical properties.

Both planets are hot sub-Saturns in close, virtually circular orbits. They have similar masses and radii and are located within the Neptunian desert. Due to its closer orbit, TOI-3071 b is much deeper in the desert than TOI-2374 b. This, and the fact that TOI-3071 is a much hotter star and emits higher XUV flux, leads to TOI-3071 b having a much higher equilibrium temperature, higher even than that of all previously detected giant exoplanets with similar masses or radii. Both planets are very metal rich, with TOI-3071 b being more enriched both in absolute and relative terms. By studying the evolution of the planets’ atmospheres under photoevaporation we find that both are stable against evaporation.

Both planets are amenable to further follow-up with JWST, as despite the relatively low TSM of TOI-3071 b, it is a fairly bright (11.87 T mag) target undergoing atmospheric evaporation. Meanwhile, TOI-2374 b has a higher TSM and lies close to the Continuous Viewing Zone of both TESS and JWST for ease of observation scheduling. Measuring the atmospheric chemical compositions would prove very useful to further constrain the formation, evolution and interior of the planets.

Acknowledgements

VA is supported by FCT - Fundação para a Ciência e a Tecnologia through national funds by the following grants: UIDP/04434/2020 (https://doi.org/10.54499/UIDP/04434/2020); UIDB/04434/2020 (https://doi.org/10.54499/UIDB/04434/2020); 2022.06962.PTDC (http://doi.org/10.54499/2022.06962.PTDC).

KAC and CNW acknowledge support from the TESS mission via subaward s3449 from MIT.

FB carried out this work within the framework of the National Centre of Competence in Research PlanetS supported by the Swiss National Science Foundation.

XD acknowledges the support from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement SCORE No 851555) and from the Swiss National Science Foundation under the grant SPECTRE (No 200021_215200)

JL-B was partly funded by grants LCF/BQ/PI20/11760023, Ramón y Cajal fellowship with code RYC2021-031640-I, and the Spanish MCIN/AEI/10.13039/501100011033 grant PID2019-107061GB-C61.

NCS is co-funded by the European Union (ERC, FIERCE, 101052347). Views and opinions expressed are however those of the author(s) only and do not necessarily reflect those of the European Union or the European Research Council. Neither the European Union nor the granting authority can be held responsible for them. This work was supported by FCT - Fundação para a Ciência e a Tecnologia through national funds and by FEDER through COMPETE2020 - Programa Operacional Competitividade e Internacionalização by these grants: UIDB/04434/2020; UIDP/04434/2020.

This work makes use of observations from the LCOGT network. Part of the LCOGT telescope time was granted by NOIRLab through the Mid-Scale Innovations Program (MSIP). MSIP is funded by NSF.

This research has made use of the Exoplanet Follow-up Observation Program (ExoFOP; DOI: 10.26134/ExoFOP5) website, which is operated by the California Institute of Technology, under contract with the National Aeronautics and Space Administration under the Exoplanet Exploration Program.

This research has made use of the NASA Exoplanet Archive, which is operated by the California Institute of Technology, under contract with the National Aeronautics and Space Administration under the Exoplanet Exploration Program.

This paper made use of data collected by the TESS mission and are publicly available from the Mikulski Archive for Space Telescopes (MAST) operated by the Space Telescope Science Institute (STScI).

We acknowledge the use of public TESS data from pipelines at the TESS Science Office and at the TESS Science Processing Operations Center.

Resources supporting this work were provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Division at Ames Research Center for the production of the SPOC data products.

Funding for the TESS mission is provided by NASA’s Science Mission Directorate. KAC and CNW acknowledge support from the TESS mission via subaward s3449 from MIT.

This research was funded in part by the UKRI, (Grants ST/X001121/1, EP/X027562/1). For the purpose of open access, the author has applied a Creative Commons Attribution (CC BY) licence to any Author Accepted Manuscript version arising from this submission.

Data Availability

The TESS data are available from the Mikulski Archive for Space Telescopes (MAST), at https://heasarc.gsfc.nasa.gov/docs/tess/data-access.html. The other photometry from LCOGT, Briefrield Private Observatory and PEST, and the high-resolution imaging data, are available for public download from the ExoFOP-TESS archive at https://exofop.ipac.caltech.edu/tess/target.php?id=439366538 for TOI-2374 and at https://exofop.ipac.caltech.edu/tess/target.php?id=452006073 for TOI-3071. The PFS RV data is also available here. The HARPS RV data is shown in tables 4 and 5 and the full HARPS data products can be found on the ESO archive. The model code underlying this article will be shared on reasonable request to the corresponding author.

References

  • Adibekyan et al. (2012) Adibekyan V. Z., Sousa S. G., Santos N. C., Delgado Mena E., González Hernández J. I., Israelian G., Mayor M., Khachatryan G., 2012, A&A, 545, A32
  • Adibekyan et al. (2015) Adibekyan V., et al., 2015, A&A, 583, A94
  • Adibekyan et al. (2021) Adibekyan V., et al., 2021, Science, 374, 330
  • Alibert et al. (2018) Alibert Y., et al., 2018, Nature Astronomy, 2, 873
  • Aller et al. (2020) Aller A., Lillo-Box J., Jones D., Miranda L. F., Barceló Forteza S., 2020, A&A, 635, A128
  • Almenara et al. (2022) Almenara J. M., et al., 2022, Photodynamical analysis of the nearly resonant planetary system WASP-148 - Accurate transit-timing variations and mutual orbital inclination, doi:10.1051/0004-6361/202142964, https://doi.org/10.1051/0004-6361/202142964
  • Anderson et al. (2011) Anderson D. R., et al., 2011, ApJ, 726, L19
  • Baglin et al. (2008) Baglin A., Auvergne M., Barge P., Deleuil M., Michel E., 2008, Proceedings of the International Astronomical Union, 4, 71–81
  • Baluev (2008) Baluev R. V., 2008, MNRAS, 385, 1279
  • Baranne et al. (1996) Baranne A., et al., 1996, A&AS, 119, 373
  • Barnes (2007) Barnes S. A., 2007, ApJ, 669, 1167
  • Bertran de Lis et al. (2015) Bertran de Lis S., Delgado Mena E., Adibekyan V. Z., Santos N. C., Sousa S. G., 2015, A&A, 576, A89
  • Bitsch et al. (2018) Bitsch B., Morbidelli A., Johansen A., Lega E., Lambrechts M., Crida A., 2018, A&A, 612, A30
  • Boisse et al. (2011) Boisse I., Bouchy F., Hébrard G., Bonfils X., Santos N., Vauclair S., 2011, A&A, 528, A4
  • Borucki et al. (2010) Borucki W. J., et al., 2010, Science, 327, 977
  • Brown et al. (2013) Brown T. M., et al., 2013, PASP, 125, 1031
  • Buchhave et al. (2010) Buchhave L. A., et al., 2010, ApJ, 720, 1118
  • Buchhave et al. (2012) Buchhave L. A., et al., 2012, Nature, 486, 375
  • Burrows (2014) Burrows A. S., 2014, Proceedings of the National Academy of Science, 111, 12601
  • Butler et al. (1996) Butler R. P., Marcy G. W., Williams E., McCarthy C., Dosanjh P., Vogt S. S., 1996, PASP, 108, 500
  • Chabrier & Debras (2021) Chabrier G., Debras F., 2021, ApJ, 917, 4
  • Chen & Rogers (2016) Chen H., Rogers L. A., 2016, ApJ, 831, 180
  • Choi et al. (2016) Choi J., Dotter A., Conroy C., Cantiello M., Paxton B., Johnson B. D., 2016, ApJ, 823, 102
  • Collins et al. (2017) Collins K. A., Kielkopf J. F., Stassun K. G., Hessman F. V., 2017, AJ, 153, 77
  • Collins et al. (2018) Collins K., Quinn S. N., Latham D. W., Christiansen J., Ciardi D., Dragomir D., Crossfield I., Seager S., 2018, in American Astronomical Society Meeting Abstracts #231. p. 439.08
  • Crane et al. (2006) Crane J. D., Shectman S. A., Butler R. P., 2006, in McLean I. S., Iye M., eds, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Vol. 6269, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series. p. 626931, doi:10.1117/12.672339
  • Crane et al. (2008) Crane J. D., Shectman S. A., Butler R. P., Thompson I. B., Burley G. S., 2008, in McLean I. S., Casali M. M., eds, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Vol. 7014, Ground-based and Airborne Instrumentation for Astronomy II. p. 701479, doi:10.1117/12.789637
  • Crane et al. (2010) Crane J. D., Shectman S. A., Butler R. P., Thompson I. B., Birk C., Jones P., Burley G. S., 2010, in McLean I. S., Ramsay S. K., Takami H., eds, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Vol. 7735, Ground-based and Airborne Instrumentation for Astronomy III. p. 773553, doi:10.1117/12.857792
  • Daemgen et al. (2009) Daemgen S., Hormuth F., Brandner W., Bergfors C., Janson M., Hippler S., Henning T., 2009, Astronomy and Astrophysics, v.498, 567-574 (2009), 498
  • Deeg et al. (2023) Deeg H. J., et al., 2023, A&A, 677, A12
  • Delgado Mena et al. (2014) Delgado Mena E., et al., 2014, A&A, 562, A92
  • Delgado Mena et al. (2017) Delgado Mena E., Tsantaki M., Adibekyan V. Z., Sousa S. G., Santos N. C., González Hernández J. I., Israelian G., 2017, A&A, 606, A94
  • Delgado Mena et al. (2019) Delgado Mena E., et al., 2019, A&A, 624, A78
  • Delgado Mena et al. (2021) Delgado Mena E., Adibekyan V., Santos N. C., Tsantaki M., González Hernández J. I., Sousa S. G., Bertrán de Lis S., 2021, A&A, 655, A99
  • Dong et al. (2018) Dong S., Xie J.-W., Zhou J.-L., Zheng Z., Luo A., 2018, Proceedings of the National Academy of Sciences, 115, 266
  • Dormand & Prince (1980) Dormand J., Prince P., 1980, Journal of Computational and Applied Mathematics, 6, 19
  • Doyle et al. (2014) Doyle A. P., Davies G. R., Smalley B., Chaplin W. J., Elsworth Y., 2014, MNRAS, 444, 3592
  • Fűrész (2008) Fűrész G., 2008, PhD thesis, University of Szeged, Hungary
  • Fernández Fernández et al. (2023) Fernández Fernández J., Wheatley P. J., King G. W., 2023, MNRAS, 522, 4251
  • Foreman-Mackey et al. (2020) Foreman-Mackey D., Czekala I., Luger R., Agol E., Barentsen G., Barclay T., 2020, exoplanet-dev/exoplanet v0.2.6, doi:10.5281/zenodo.1998447, https://doi.org/10.5281/zenodo.1998447
  • Freedman et al. (2014) Freedman R. S., Lustig-Yaeger J., Fortney J. J., Lupu R. E., Marley M. S., Lodders K., 2014, ApJS, 214, 25
  • Gaia Collaboration et al. (2023) Gaia Collaboration et al., 2023, A&A, 674, A1
  • Gardner et al. (2006) Gardner J. P., et al., 2006, Space Science Reviews, 123, 485–606
  • Ginzburg & Chiang (2020) Ginzburg S., Chiang E., 2020, MNRAS, 498, 680
  • Guerrero et al. (2021) Guerrero N. M., et al., 2021, ApJS, 254, 39
  • Guillot (2010) Guillot T., 2010, A&A, 520, A27
  • Guillot et al. (2006) Guillot T., Santos N. C., Pont F., Iro N., Melo C., Ribas I., 2006, A&A, 453, L21
  • Hartman et al. (2019) Hartman J. D., et al., 2019, The Astronomical Journal, 157, 55
  • Hasegawa et al. (2018) Hasegawa Y., Bryden G., Ikoma M., Vasisht G., Swain M., 2018, ApJ, 865, 32
  • Helled (2023) Helled R., 2023, A&A, 675, L8
  • Helled et al. (2014) Helled R., et al., 2014, in Beuther H., Klessen R. S., Dullemond C. P., Henning T., eds, Protostars and Planets VI. pp 643–665 (arXiv:1311.1142), doi:10.2458/azu_uapress_9780816531240-ch028
  • Helled et al. (2022a) Helled R., et al., 2022a, Experimental Astronomy, 53, 323
  • Helled et al. (2022b) Helled R., et al., 2022b, Icarus, 378, 114937
  • Howard & Guillot (2023) Howard S., Guillot T., 2023, A&A, 672, L1
  • Hoyer et al. (2023) Hoyer S., et al., 2023, A&A, 675, A81
  • Huang et al. (2020) Huang C. X., et al., 2020, Research Notes of the American Astronomical Society, 4, 204
  • Husser, T.-O. et al. (2013) Husser, T.-O. Wende-von Berg, S. Dreizler, S. Homeier, D. Reiners, A. Barman, T. Hauschildt, P. H. 2013, A&A, 553, A6
  • Jenkins (2002) Jenkins J. M., 2002, ApJ, 575, 493
  • Jenkins et al. (2010) Jenkins J. M., et al., 2010, in Radziwill N. M., Bridger A., eds, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Vol. 7740, Software and Cyberinfrastructure for Astronomy. p. 77400D, doi:10.1117/12.856764
  • Jenkins et al. (2020) Jenkins J. M., Tenenbaum P., Seader S., Burke C. J., McCauliff S. D., Smith J. C., Twicken J. D., Chandrasekaran H., 2020, Kepler Data Processing Handbook: Transiting Planet Search, Kepler Science Document KSCI-19081-003
  • Jensen (2013) Jensen E., 2013, Tapir: A web interface for transit/eclipse observability, Astrophysics Source Code Library (ascl:1306.007)
  • Jermyn et al. (2023) Jermyn A. S., et al., 2023, ApJS, 265, 15
  • Johansen & Lambrechts (2017) Johansen A., Lambrechts M., 2017, Annual Review of Earth and Planetary Sciences, 45, 359
  • Johnstone et al. (2021) Johnstone C. P., Bartel M., Güdel M., 2021, A&A, 649, A96
  • Kempton et al. (2018) Kempton E. M.-R., et al., 2018, Publications of the Astronomical Society of the Pacific, 130, 114401
  • King & Wheatley (2021) King G. W., Wheatley P. J., 2021, MNRAS, 501, L28
  • Kipping (2013) Kipping D. M., 2013, mnras, 435, 2152
  • Kubyshkina et al. (2018) Kubyshkina D., et al., 2018, ApJ, 866, L18
  • Kurucz (1992) Kurucz R. L., 1992, in Barbuy B., Renzini A., eds, The Stellar Populations of Galaxies. Springer Netherlands, Dordrecht, pp 225–232
  • Kurucz (1993) Kurucz R. L., 1993, SYNTHE spectrum synthesis programs and line data
  • Li et al. (2018) Li L., et al., 2018, Nature Communications, 9, 3709
  • Li et al. (2019) Li J., Tenenbaum P., Twicken J. D., Burke C. J., Jenkins J. M., Quintana E. V., Rowe J. F., Seader S. E., 2019, PASP, 131, 024506
  • Lightkurve Collaboration et al. (2018) Lightkurve Collaboration et al., 2018, Lightkurve: Kepler and TESS time series analysis in Python, Astrophysics Source Code Library, record ascl:1812.013 (ascl:1812.013)
  • Lillo-Box et al. (2012) Lillo-Box J., Barrado D., Bouy H., 2012, Astronomy &amp; Astrophysics, 546, A10
  • Luger et al. (2019) Luger R., Agol E., Foreman-Mackey D., Fleming D. P., Lustig-Yaeger J., Deitrick R., 2019, aj, 157, 64
  • Mayor & Queloz (1995) Mayor M., Queloz D., 1995, Nature, 378, 355
  • Mayor et al. (2003a) Mayor M., et al., 2003a, The Messenger, 114, 20
  • Mayor et al. (2003b) Mayor M., et al., 2003b, The Messenger, 114, 20
  • Mazeh et al. (2016) Mazeh T., Holczer T., Faigler S., 2016, A&A, 589, A75
  • McCully et al. (2018) McCully C., Volgenau N. H., Harbeck D.-R., Lister T. A., Saunders E. S., Turner M. L., Siiverd R. J., Bowman M., 2018, in Proc. SPIE. p. 107070K (arXiv:1811.04163), doi:10.1117/12.2314340
  • Miller & Fortney (2011) Miller N., Fortney J. J., 2011, ApJ, 736, L29
  • Mousis et al. (2009) Mousis O., Marboeuf U., Lunine J. I., Alibert Y., Fletcher L. N., Orton G. S., Pauzat F., Ellinger Y., 2009, ApJ, 696, 1348
  • Mulders et al. (2016) Mulders G. D., Pascucci I., Apai D., Frasca A., Molenda-Żakowicz J., 2016, The Astronomical Journal, 152, 187
  • Müller & Helled (2021) Müller S., Helled R., 2021, MNRAS, 507, 2094
  • Müller & Helled (2023a) Müller S., Helled R., 2023a, Frontiers in Astronomy and Space Sciences, 10, 1179000
  • Müller & Helled (2023b) Müller S., Helled R., 2023b, A&A, 669, A24
  • Müller et al. (2020a) Müller S., Helled R., Cumming A., 2020a, A&A, 638, A121
  • Müller et al. (2020b) Müller S., Ben-Yami M., Helled R., 2020b, ApJ, 903, 147
  • Müller & Helled (2024) Müller S., Helled R., 2024, Can Jupiter’s atmospheric metallicity be different from the deep interior? (arXiv:2403.16273)
  • Noyes et al. (1984) Noyes R. W., Weiss N. O., Vaughan A. H., 1984, ApJ, 287, 769
  • Otegi et al. (2020) Otegi J. F., Bouchy F., Helled R., 2020, A&A, 634, A43
  • Owen & Jackson (2012) Owen J. E., Jackson A. P., 2012, MNRAS, 425, 2931
  • Owen & Lai (2018) Owen J. E., Lai D., 2018, MNRAS, 479, 5012
  • Owen & Murray-Clay (2018) Owen J. E., Murray-Clay R., 2018, MNRAS, 480, 2206
  • Owen & Wu (2017) Owen J. E., Wu Y., 2017, ApJ, 847, 29
  • Parmentier & Crossfield (2018) Parmentier V., Crossfield I. J. M., 2018, Exoplanet Phase Curves: Observations and Theory. Springer International Publishing, Cham, pp 1419–1440, doi:10.1007/978-3-319-55333-7_116, https://doi.org/10.1007/978-3-319-55333-7_116
  • Parmentier & Guillot (2014) Parmentier V., Guillot T., 2014, A&A, 562, A133
  • Paxton et al. (2011) Paxton B., Bildsten L., Dotter A., Herwig F., Lesaffre P., Timmes F., 2011, ApJS, 192, 3
  • Paxton et al. (2013) Paxton B., et al., 2013, ApJS, 208, 4
  • Paxton et al. (2015) Paxton B., et al., 2015, ApJS, 220, 15
  • Paxton et al. (2018) Paxton B., et al., 2018, ApJS, 234, 34
  • Paxton et al. (2019) Paxton B., et al., 2019, ApJS, 243, 10
  • Pepe et al. (2002) Pepe F., et al., 2002, The Messenger, 110, 9
  • Pepe et al. (2021) Pepe F., et al., 2021, A&A, 645, A96
  • Petigura et al. (2018) Petigura E. A., et al., 2018, The Astronomical Journal, 155, 89
  • Pizzolato et al. (2003) Pizzolato N., Maggio A., Micela G., Sciortino S., Ventura P., 2003, A&A, 397, 147
  • Pollacco et al. (2006) Pollacco D. L., et al., 2006, PASP, 118, 1407
  • Pollack et al. (1996) Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J., Podolak M., Greenzweig Y., 1996, Icarus, 124, 62
  • Ricker et al. (2015) Ricker G. R., et al., 2015, Journal of Astronomical Telescopes, Instruments, and Systems, 1, 014003
  • Salvatier et al. (2016) Salvatier J., Wiecki T. V., Fonnesbeck C., 2016, PeerJ Computer Science, 2, e55
  • Santos et al. (2013) Santos N. C., et al., 2013, A&A, 556, A150
  • Santos et al. (2015) Santos N. C., et al., 2015, A&A, 580, L13
  • Santos et al. (2017) Santos N. C., et al., 2017, A&A, 608, A94
  • Schlegel et al. (1998) Schlegel D. J., Finkbeiner D. P., Davis M., 1998, ApJ, 500, 525
  • Sestovic et al. (2018) Sestovic M., Demory B.-O., Queloz D., 2018, A&A, 616, A76
  • Shibata et al. (2020) Shibata S., Helled R., Ikoma M., 2020, A&A, 633, A33
  • Skrutskie et al. (2006) Skrutskie M. F., et al., 2006, AJ, 131, 1163
  • Smith et al. (2012) Smith J. C., et al., 2012, PASP, 124, 1000
  • Sneden (1973) Sneden C. A., 1973, PhD thesis, THE UNIVERSITY OF TEXAS AT AUSTIN.
  • Sousa (2014) Sousa S. G., 2014, ARES + MOOG: A Practical Overview of an Equivalent Width (EW) Method to Derive Stellar Parameters. Springer International Publishing, p. 297–310, doi:10.1007/978-3-319-06956-2_26, http://dx.doi.org/10.1007/978-3-319-06956-2_26
  • Sousa et al. (2007) Sousa S. G., Santos N. C., Israelian G., Mayor M., Monteiro M. J. P. F. G., 2007, A&A, 469, 783
  • Sousa et al. (2008) Sousa S. G., et al., 2008, A&A, 487, 373
  • Sousa et al. (2015) Sousa S. G., Santos N. C., Adibekyan V., Delgado-Mena E., Israelian G., 2015, A&A, 577, A67
  • Sousa et al. (2021) Sousa S. G., et al., 2021, arXiv e-prints, p. arXiv:2109.04781
  • Stassun & Torres (2016) Stassun K. G., Torres G., 2016, AJ, 152, 180
  • Stassun et al. (2017) Stassun K. G., Collins K. A., Gaudi B. S., 2017, AJ, 153, 136
  • Stassun et al. (2018) Stassun K. G., Corsaro E., Pepper J. A., Gaudi B. S., 2018, AJ, 155, 22
  • Stassun et al. (2019) Stassun K. G., et al., 2019, AJ, 158, 138
  • Stumpe et al. (2012) Stumpe M. C., et al., 2012, PASP, 124, 985
  • Stumpe et al. (2014) Stumpe M. C., Smith J. C., Catanzarite J. H., Van Cleve J. E., Jenkins J. M., Twicken J. D., Girouard F. R., 2014, PASP, 126, 100
  • Szabó & Kiss (2011) Szabó G. M., Kiss L. L., 2011, ApJ, 727, L44
  • Thorngren & Fortney (2018) Thorngren D. P., Fortney J. J., 2018, AJ, 155, 214
  • Thorngren & Fortney (2019) Thorngren D., Fortney J. J., 2019, ApJ, 874, L31
  • Thorngren et al. (2016) Thorngren D. P., Fortney J. J., Murray-Clay R. A., Lopez E. D., 2016, ApJ, 831, 64
  • Tokovinin (2018) Tokovinin A., 2018, PASP, 130, 035002
  • Torres et al. (2010) Torres G., Andersen J., Giménez A., 2010, A&ARv, 18, 67
  • Tsantaki et al. (2013) Tsantaki M., Sousa S. G., Adibekyan V. Z., Santos N. C., Mortier A., Israelian G., 2013, A&A, 555, A150
  • Twicken et al. (2010) Twicken J. D., Chandrasekaran H., Jenkins J. M., Gunter J. P., Girouard F., Klaus T. C., 2010, in Radziwill N. M., Bridger A., eds, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series Vol. 7740, Software and Cyberinfrastructure for Astronomy. p. 77401U, doi:10.1117/12.856798
  • Twicken et al. (2018) Twicken J. D., et al., 2018, PASP, 130, 064502
  • Vehtari et al. (2021) Vehtari A., Gelman A., Simpson D., Carpenter B., Bürkner P.-C., 2021, Bayesian Analysis, 16
  • Vissapragada et al. (2022) Vissapragada S., et al., 2022, AJ, 164, 234
  • Wakeford et al. (2017) Wakeford H. R., et al., 2017, The Astronomical Journal, 155, 29
  • Wilson et al. (2022) Wilson T. G., et al., 2022, MNRAS, 511, 1043
  • Winn (2014) Winn J. N., 2014, Transits and Occultations (arXiv:1001.2010)
  • Wright et al. (2011) Wright N. J., Drake J. J., Mamajek E. E., Henry G. W., 2011, ApJ, 743, 48
  • Zechmeister & Kürster (2009) Zechmeister M., Kürster M., 2009, A&A, 496, 577
  • Ziegler et al. (2020) Ziegler C., Tokovinin A., Briceño C., Mang J., Law N., Mann A. W., 2020, AJ, 159, 19
  • da Silva et al. (2006) da Silva L., et al., 2006, A&A, 458, 609

Appendix A Spectroscopy

The HARPS RV data (described in Section 2.2.1) are presented in tables 4 (TOI-3071) anb 5 (TOI-2374). The PFS RV data (described in Section 2.2.2) are presented in Table 6.

Table 4: TOI-3071 HARPS radial velocities.
Time RV σRVsubscript𝜎RV\sigma_{\textrm{RV}}italic_σ start_POSTSUBSCRIPT RV end_POSTSUBSCRIPT FWHM Bisector Contrast
(BJD-2457000) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (kms1𝑘𝑚superscript𝑠1kms^{-1}italic_k italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT)
2657.758 9353.8 4.5 7.945(11) 45(6) 43.199052(1)
2658.831 9375.9 4.8 7.938(11) 48(7) 43.167730(1)
2660.804 9390.7 3.8 7.957(11) 43(6) 43.186370(1)
2664.689 9410.5 3.1 7.958(11) 45(4) 43.381653(1)
2664.818 9415.3 4.3 7.985(11) 33(6) 43.070969(1)
2665.622 9358.9 2.9 7.954(11) 21(4) 43.494550(1)
2665.806 9391.0 4.3 7.960(11) 23(6) 43.497998(1)
2666.774 9347.8 3.7 7.941(11) 40(5) 43.568806(1)
2667.597 9402.3 3.9 7.958(11) 30(5) 43.480382(1)
2667.759 9376.5 3.9 7.938(11) 38(5) 43.521020(1)
2668.766 9408.2 5.1 7.953(11) 30(7) 43.469858(1)
2669.802 9416.1 4.2 7.951(11) 26(6) 43.498341(1)
2670.753 9366.7 3.8 7.950(11) 20(5) 43.505071(1)
2671.658 9358.0 3.3 7.961(11) 24(4) 43.514566(1)
  • The full HARPS data products can be found on the ESO archive

Table 5: TOI-2374 HARPS radial velocities.
Time RV σRVsubscript𝜎RV\sigma_{\textrm{RV}}italic_σ start_POSTSUBSCRIPT RV end_POSTSUBSCRIPT FWHM Bisector Contrast
(BJD-2457000) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (kms1𝑘𝑚superscript𝑠1kms^{-1}italic_k italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT)
2678.897 14983.6 3.4 6.6264(88) 35(5) 51.054114(1)
2679.890 15008.6 3.7 6.2425(88) 39(5) 51.041247(1)
2681.905 14961.4 3.6 6.2461(88) 34(5) 51.061773(1)
2682.913 14963.8 2.8 6.2338(88) 35(4) 50.923438(1)
2711.852 14006.3 1.7 9.193(13) 81(2) 21.564691(1)
2712.822 14938 28 6.2944(88) -22(39) 52.935152(1)
2714.869 15012.4 2.8 6.2385(88) 42(4) 50.732359(1)
2715.792 14969.3 2.7 6.2179(88) 26(4) 50.174915(1)
2716.806 14902 20 6.3119(89) 42(28) 51.119746(1)
2723.924 14994.7 3.8 6.2257(88) 27(5) 51.121218(1)
2725.869 14958.4 2.6 6.2410(88) 25(4) 51.017728(1)
2726.847 14997.4 2.9 6.2614(88) 23(4) 51.007037(1)
2727.894 15006.0 2.7 6.2535(88) 30(4) 51.031929(1)
2728.850 14961.7 5.0 6.2662(88) 17(7) 51.068559(1)
2750.724 14974.5 7.8 6.3386(90) 59(11) 50.319108(1)
2759.766 14958.7 3.5 6.2457(88) 24(5) 51.046751(1)
2760.767 14983.7 7.9 6.2861(88) 25(11) 51.429800(1)
2760.788 14978.0 4.6 6.2310(88) 34(6) 51.064180(1)
2761.722 14992.3 6.6 6.2332(88) 39(9) 51.422302(1)
2766.754 14997.3 5.0 6.2391(88) 18(7) 51.411205(1)
2767.864 14959.4 3.2 6.2342(88) 23(4) 51.197217(1)
  • The full HARPS data products can be found on the ESO archive

Table 6: TOI-2374 PFS radial velocities.
Time ΔΔ\Deltaroman_ΔRV σRVsubscript𝜎RV\sigma_{\textrm{RV}}italic_σ start_POSTSUBSCRIPT RV end_POSTSUBSCRIPT Bisector σbissubscript𝜎bis\sigma_{\textrm{bis}}italic_σ start_POSTSUBSCRIPT bis end_POSTSUBSCRIPT
(BJD-2457000) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) (ms1𝑚superscript𝑠1ms^{-1}italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT)
2471.603 4.45 2.3 46.6 31.4
2535.533 -20.59 3.3 94.4 118.8
2472.547 43.89 2.06 100.5 24.9
2504.567 -2.89 5.02 -179.6 173.2
2509.534 0 3.41 -69.6 59.0
2473.515 36.13 2.25 -43.4 32.1
2534.528 -5.46 2.67 -36.9 53.6

Appendix B Stellar abundances and chemical clocks

The stellar abundances of the elements derived in Sect. 3.1.2 are shown in Table 7 and the stellar ages derived from chemical clocks as described in Sect. 3.1.3 are shown in Table 8.

Table 7: Stellar abundances weighted average and standard deviation, derived with the logg from Gaia DR3. Li is depleted in both targets. The abundances of carbon and oxygen for TOI-2374 cannot be determined for this cool star with our methodology. For TOI-3071, the abundance of oxygen has a high error because the lines are very weak and hard to measure in its spectrum with low S/N. Moreover, one of the lines was contaminated by a telluric line in many of the spectra so it was measured using a small number of spectra.
Star TOI-2374 TOI-3071
[Fe/H] 0.15 ±plus-or-minus\pm± 0.04 0.35 ±plus-or-minus\pm± 0.04
[O/H] - 0.332 ±plus-or-minus\pm± 0.154
[C/H] - 0.241 ±plus-or-minus\pm± 0.026
[Cu/H] 0.293 ±plus-or-minus\pm± 0.100 0.358 ±plus-or-minus\pm± 0.028
[Zn/H] 0.088 ±plus-or-minus\pm± 0.079 0.284 ±plus-or-minus\pm± 0.058
[Sr/H] 0.248 ±plus-or-minus\pm± 0.167 0.272 ±plus-or-minus\pm± 0.077
[Y/H] 0.090 ±plus-or-minus\pm± 0.155 0.327 ±plus-or-minus\pm± 0.023
[Zr/H] 0.321 ±plus-or-minus\pm± 0.109 0.339 ±plus-or-minus\pm± 0.025
[Ba/H] 0.026 ±plus-or-minus\pm± 0.087 0.280 ±plus-or-minus\pm± 0.039
[Ce/H] 0.120 ±plus-or-minus\pm± 0.117 0.219 ±plus-or-minus\pm± 0.036
[Nd/H] 0.323 ±plus-or-minus\pm± 0.084 0.209 ±plus-or-minus\pm± 0.034
[Mg/H] 0.07 ±plus-or-minus\pm± 0.09 0.30 ±plus-or-minus\pm± 0.06
[Al/H] 0.17 ±plus-or-minus\pm± 0.07 0.30 ±plus-or-minus\pm± 0.08
[Si/H] 0.21 ±plus-or-minus\pm± 0.09 0.33 ±plus-or-minus\pm± 0.04
[Ti/H] 0.20 ±plus-or-minus\pm± 0.12 0.33 ±plus-or-minus\pm± 0.04
[Ni/H] - 0.36 ±plus-or-minus\pm± 0.02
Table 8: Stellar ages derived from chemical clocks (see Sect. 3.1.3).
TOI-2374: 2D clocks Age (Gyr)
[Y/Zn] 3.6911800 ±plus-or-minus\pm± 3.7638734
[Zn/Fe] 2.8571398 ±plus-or-minus\pm± 2.1241537
[Y/al] 4.4640000 ±plus-or-minus\pm± 2.7333606
[Y/Mg] 2.6525000 ±plus-or-minus\pm± 3.7252200
TOI-3071: 3D clocks Age (Gyr)
[Y/Zn] 1.0152100 ±plus-or-minus\pm± 1.0256072
[Y/Ti] 1.1382002 ±plus-or-minus\pm± 0.87969329
[Y/Mg] 0.94835018 ±plus-or-minus\pm± 0.98228770
[Sr/Ti] 1.7918402 ±plus-or-minus\pm± 1.4564712
[Sr/Mg] 1.4527202 ±plus-or-minus\pm± 1.3351688
[Y/Si] 0.93273022 ±plus-or-minus\pm± 0.85117129
[Sr/Si] 1.1112002 ±plus-or-minus\pm± 1.3381209
[Y/Al] 1.4601001 ±plus-or-minus\pm± 0.80915273

Appendix C The joint fit

The full set of priors and fit values from our joint fit model described in Section 3.3 are presented in tables 9 and 10. The posterior distributions for the dilution correction factors are shown in figures 19 and 20.

Table 9: Prior distributions used in our joint fit model for the TOI-2374 system, fully described in Section 3.3, and the fit values resulting from the model. The priors are created using distributions in PyMC3 with the relevant inputs to each distribution described in the table footer. The fit values are given as the median values of the samples, and the uncertainties as the 16th and 84th percentiles. Further (derived) system parameters can be found in Table 3.
Parameter (unit) Prior Distribution Fit value
Planet
log(P)𝑃\log{(P)}roman_log ( italic_P ) (days) 𝒩(1.461777,1.0)𝒩1.4617771.0\mathcal{N}(1.461777,1.0)caligraphic_N ( 1.461777 , 1.0 ) 1.461776±0.000002plus-or-minus1.4617760.0000021.461776\pm 0.0000021.461776 ± 0.000002
Epoch t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (BJD-2457000) 𝒩(1326.555387,1.0)𝒩1326.5553871.0\mathcal{N}(1326.555387,1.0)caligraphic_N ( 1326.555387 , 1.0 ) 1326.563±0.002plus-or-minus1326.5630.0021326.563\pm 0.0021326.563 ± 0.002
log(Rp)subscript𝑅𝑝\log{(R_{p})}roman_log ( italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) (R) 𝒩(2.6998,1.0)𝒩superscript2.69981.0\mathcal{N}(-2.6998^{*},1.0)caligraphic_N ( - 2.6998 start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , 1.0 ) 2.40±0.03plus-or-minus2.400.03-2.40\pm 0.03- 2.40 ± 0.03
esinω𝑒𝜔\sqrt{e}\sin{\omega}square-root start_ARG italic_e end_ARG roman_sin italic_ω 𝒰(1,1)𝒰11\mathcal{U}(-1,1)caligraphic_U ( - 1 , 1 ) 0.200.09+0.06subscriptsuperscript0.200.060.090.20^{+0.06}_{-0.09}0.20 start_POSTSUPERSCRIPT + 0.06 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.09 end_POSTSUBSCRIPT
ecosω𝑒𝜔\sqrt{e}\cos{\omega}square-root start_ARG italic_e end_ARG roman_cos italic_ω 𝒰(1,1)𝒰11\mathcal{U}(-1,1)caligraphic_U ( - 1 , 1 ) 0.100.17+0.13subscriptsuperscript0.100.130.170.10^{+0.13}_{-0.17}0.10 start_POSTSUPERSCRIPT + 0.13 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.17 end_POSTSUBSCRIPT
Star
Mass Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT (M) 𝒩(0.75,0.01,0.0,3.0)subscript𝒩0.750.010.03.0\mathcal{N_{B}}(0.75,0.01,0.0,3.0)caligraphic_N start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT ( 0.75 , 0.01 , 0.0 , 3.0 ) 0.75±0.01plus-or-minus0.750.010.75\pm 0.010.75 ± 0.01
Radius Rsubscript𝑅R_{\star}italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT (R) 𝒩(0.69,0.02,0.0,3.0)subscript𝒩0.690.020.03.0\mathcal{N_{B}}(0.69,0.02,0.0,3.0)caligraphic_N start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT ( 0.69 , 0.02 , 0.0 , 3.0 ) 0.69±0.02plus-or-minus0.690.020.69\pm 0.020.69 ± 0.02
Limb darkening coefficients
TESS filter u1subscript𝑢1u_{1}italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT Kipping (2013) 0.70.4+0.5subscriptsuperscript0.70.50.40.7^{+0.5}_{-0.4}0.7 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.4 end_POSTSUBSCRIPT
TESS filter u2subscript𝑢2u_{2}italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Kipping (2013) 0.00.4+0.5subscriptsuperscript0.00.50.40.0^{+0.5}_{-0.4}0.0 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.4 end_POSTSUBSCRIPT
B filter u1subscript𝑢1u_{1}italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT Kipping (2013) 0.90.6+0.5subscriptsuperscript0.90.50.60.9^{+0.5}_{-0.6}0.9 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.6 end_POSTSUBSCRIPT
B filter u2subscript𝑢2u_{2}italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Kipping (2013) 0.1±0.5plus-or-minus0.10.5-0.1\pm 0.5- 0.1 ± 0.5
ip filter u1subscript𝑢1u_{1}italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT Kipping (2013) 0.5±0.3plus-or-minus0.50.30.5\pm 0.30.5 ± 0.3
ip filter u2subscript𝑢2u_{2}italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Kipping (2013) 0.2±0.4plus-or-minus0.20.40.2\pm 0.40.2 ± 0.4
gp filter u1subscript𝑢1u_{1}italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT Kipping (2013) 0.90.4+0.3subscriptsuperscript0.90.30.40.9^{+0.3}_{-0.4}0.9 start_POSTSUPERSCRIPT + 0.3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.4 end_POSTSUBSCRIPT
gp filter u2subscript𝑢2u_{2}italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Kipping (2013) 0.20.3+0.4subscriptsuperscript0.20.40.3-0.2^{+0.4}_{-0.3}- 0.2 start_POSTSUPERSCRIPT + 0.4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.3 end_POSTSUBSCRIPT
Photometry
TESSS1 transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.00023±0.00008plus-or-minus1.000230.000081.00023\pm 0.000081.00023 ± 0.00008
TESSS28 transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.00011±0.00008plus-or-minus1.000110.000081.00011\pm 0.000081.00011 ± 0.00008
Brierfield transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.0000±0.0005plus-or-minus1.00000.00051.0000\pm 0.00051.0000 ± 0.0005
LCO-SAAO transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.0001±0.0001plus-or-minus1.00010.00011.0001\pm 0.00011.0001 ± 0.0001
LCO-SSO (ip filter) transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.0000±0.0002plus-or-minus1.00000.00021.0000\pm 0.00021.0000 ± 0.0002
LCO-SSO (gp filter) transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.0000±0.0002plus-or-minus1.00000.00021.0000\pm 0.00021.0000 ± 0.0002
TESSS1 dilution correction factor d𝑑ditalic_d 𝒩(0.0,0.2)𝒩0.00.2\mathcal{N}(0.0,0.2)caligraphic_N ( 0.0 , 0.2 ) 0.070.09+0.08subscriptsuperscript0.070.080.09-0.07^{+0.08}_{-0.09}- 0.07 start_POSTSUPERSCRIPT + 0.08 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.09 end_POSTSUBSCRIPT
TESSS28 dilution correction factor d𝑑ditalic_d 𝒩(0.0,0.2)𝒩0.00.2\mathcal{N}(0.0,0.2)caligraphic_N ( 0.0 , 0.2 ) 0.040.08+0.06subscriptsuperscript0.040.060.08-0.04^{+0.06}_{-0.08}- 0.04 start_POSTSUPERSCRIPT + 0.06 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.08 end_POSTSUBSCRIPT
Brierfield dilution correction factor d𝑑ditalic_d 0.00.00.00.0 (fixed) -
LCO-SAAO dilution correction factor d𝑑ditalic_d 0.00.00.00.0 (fixed) -
LCO-SSO (ip filter) dilution correction factor d𝑑ditalic_d 0.00.00.00.0 (fixed) -
LCO-SSO (gp filter) dilution correction factor d𝑑ditalic_d 0.00.00.00.0 (fixed) -
TESSS1 log (Jitter) 𝒩(6.3000,0.01)𝒩superscript6.30000.01\mathcal{N}(-6.3000^{\dagger},0.01)caligraphic_N ( - 6.3000 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 0.01 ) 6.32±0.01plus-or-minus6.320.01-6.32\pm 0.01- 6.32 ± 0.01
TESSS28 log (Jitter) 𝒩(5.8934,0.01)𝒩superscript5.89340.01\mathcal{N}(-5.8934^{\dagger},0.01)caligraphic_N ( - 5.8934 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 0.01 ) 5.94±0.01plus-or-minus5.940.01-5.94\pm 0.01- 5.94 ± 0.01
Brierfield log (Jitter) 𝒩(5.8482,0.01)𝒩superscript5.84820.01\mathcal{N}(-5.8482^{\dagger},0.01)caligraphic_N ( - 5.8482 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 0.01 ) 5.85±0.01plus-or-minus5.850.01-5.85\pm 0.01- 5.85 ± 0.01
LCO-SAAO log (Jitter) 𝒩(6.7355,0.01)𝒩superscript6.73550.01\mathcal{N}(-6.7355^{\dagger},0.01)caligraphic_N ( - 6.7355 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 0.01 ) 6.74±0.01plus-or-minus6.740.01-6.74\pm 0.01- 6.74 ± 0.01
LCO-SSO (ip filter) log (Jitter) 𝒩(6.8753,0.01)𝒩superscript6.87530.01\mathcal{N}(-6.8753^{\dagger},0.01)caligraphic_N ( - 6.8753 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 0.01 ) 6.88±0.01plus-or-minus6.880.01-6.88\pm 0.01- 6.88 ± 0.01
LCO-SSO (gp filter) log (Jitter) 𝒩(6.7380,0.01)𝒩superscript6.73800.01\mathcal{N}(-6.7380^{\dagger},0.01)caligraphic_N ( - 6.7380 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 0.01 ) 6.74±0.01plus-or-minus6.740.01-6.74\pm 0.01- 6.74 ± 0.01
RVs
log(K)𝐾\log{(K)}roman_log ( italic_K ) (m s-1) 𝒰(0.0,5.0)𝒰0.05.0\mathcal{U}(0.0,5.0)caligraphic_U ( 0.0 , 5.0 ) 3.300.08+0.06subscriptsuperscript3.300.060.083.30^{+0.06}_{-0.08}3.30 start_POSTSUPERSCRIPT + 0.06 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.08 end_POSTSUBSCRIPT
HARPS offset (m s-1) 𝒩(0.0,100.0)𝒩0.0100.0\mathcal{N}(0.0,100.0)caligraphic_N ( 0.0 , 100.0 ) 7.2±4.4plus-or-minus7.24.47.2\pm 4.47.2 ± 4.4
PFS offset (m s-1) 𝒩(0.0,100.0)𝒩0.0100.0\mathcal{N}(0.0,100.0)caligraphic_N ( 0.0 , 100.0 ) 14.810.8+11.5subscriptsuperscript14.811.510.8-14.8^{+11.5}_{-10.8}- 14.8 start_POSTSUPERSCRIPT + 11.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 10.8 end_POSTSUBSCRIPT
Linear trend (m s-1 d-1) 𝒩(0.0,100.0)𝒩0.0100.0\mathcal{N}(0.0,100.0)caligraphic_N ( 0.0 , 100.0 ) 0.12±0.06plus-or-minus0.120.06-0.12\pm 0.06- 0.12 ± 0.06
Quadratic trend (m s-1 d-2) 𝒩(0.0,10.0)𝒩0.010.0\mathcal{N}(0.0,10.0)caligraphic_N ( 0.0 , 10.0 ) 0.0004±0.0003plus-or-minus0.00040.00030.0004\pm 0.00030.0004 ± 0.0003
HARPS log(Jitter)Jitter\log{(\rm{Jitter)}}roman_log ( roman_Jitter ) (m s-1) 𝒰(5.0,3.0)𝒰5.03.0\mathcal{U}(-5.0,3.0)caligraphic_U ( - 5.0 , 3.0 ) 1.03.1+0.6subscriptsuperscript1.00.63.11.0^{+0.6}_{-3.1}1.0 start_POSTSUPERSCRIPT + 0.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 3.1 end_POSTSUBSCRIPT
PFS log(Jitter)Jitter\log{(\rm{Jitter)}}roman_log ( roman_Jitter ) (m s-1) 𝒰(5.0,3.0)𝒰5.03.0\mathcal{U}(-5.0,3.0)caligraphic_U ( - 5.0 , 3.0 ) 2.0±0.4plus-or-minus2.00.42.0\pm 0.42.0 ± 0.4
  • Distributions:

  • 𝒩(μ,σ)𝒩𝜇𝜎\mathcal{N}(\mu,\sigma)caligraphic_N ( italic_μ , italic_σ ): a normal distribution with a mean μ𝜇\muitalic_μ and a standard deviation σ𝜎\sigmaitalic_σ;

  • 𝒩(μ,σ,a,b)subscript𝒩𝜇𝜎𝑎𝑏\mathcal{N_{B}}(\mu,\sigma,a,b)caligraphic_N start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT ( italic_μ , italic_σ , italic_a , italic_b ): a bounded normal distribution with a mean μ𝜇\muitalic_μ, a standard deviation σ𝜎\sigmaitalic_σ, a lower bound a𝑎aitalic_a, and an upper bound b𝑏bitalic_b (bounds optional);

  • 𝒰(a,b)𝒰𝑎𝑏\mathcal{U}(a,b)caligraphic_U ( italic_a , italic_b ): a uniform distribution between a𝑎aitalic_a and b𝑏bitalic_b;

  • Distributions for limb darkening coefficients u1subscript𝑢1u_{1}italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and u2subscript𝑢2u_{2}italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are built into the exoplanet package and based on Kipping (2013).

  • Prior values:

  • equivalent to 0.5(log(D))+log(R)0.5𝐷subscript𝑅0.5(\log{(D)})+\log{(R_{\star})}0.5 ( roman_log ( italic_D ) ) + roman_log ( italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) where D𝐷Ditalic_D is the transit depth (ppm multiplied by 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT) and Rsubscript𝑅R_{\star}italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the mean of the prior on the stellar radius (R);

  • equivalent to the log of the error median on the data of the correspondent instrument.

Table 10: Same as table 9 but for the TOI-3071 system. Further (derived) system parameters can be found in Table 3.
Parameter (unit) Prior Distribution Fit value
Planet
log(P)𝑃\log{(P)}roman_log ( italic_P ) (days) 𝒩(0.2367297,1.0)𝒩0.23672971.0\mathcal{N}(0.2367297,1.0)caligraphic_N ( 0.2367297 , 1.0 ) 0.236603±0.000002plus-or-minus0.2366030.0000020.236603\pm 0.0000020.236603 ± 0.000002
Epoch t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (BJD-2457000) 𝒩(1594.611536,1.0)𝒩1594.6115361.0\mathcal{N}(1594.611536,1.0)caligraphic_N ( 1594.611536 , 1.0 ) 1594.608±0.002plus-or-minus1594.6080.0021594.608\pm 0.0021594.608 ± 0.002
log(Rp)subscript𝑅𝑝\log{(R_{p})}roman_log ( italic_R start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) (R) 𝒩(2.709956,1.0)𝒩superscript2.7099561.0\mathcal{N}(-2.709956^{*},1.0)caligraphic_N ( - 2.709956 start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , 1.0 ) 2.99±0.07plus-or-minus2.990.07-2.99\pm 0.07- 2.99 ± 0.07
esinω𝑒𝜔\sqrt{e}\sin{\omega}square-root start_ARG italic_e end_ARG roman_sin italic_ω 𝒰(1,1)𝒰11\mathcal{U}(-1,1)caligraphic_U ( - 1 , 1 ) 0.110.07+0.10subscriptsuperscript0.110.100.07-0.11^{+0.10}_{-0.07}- 0.11 start_POSTSUPERSCRIPT + 0.10 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.07 end_POSTSUBSCRIPT
ecosω𝑒𝜔\sqrt{e}\cos{\omega}square-root start_ARG italic_e end_ARG roman_cos italic_ω 𝒰(1,1)𝒰11\mathcal{U}(-1,1)caligraphic_U ( - 1 , 1 ) 0.070.14+0.16subscriptsuperscript0.070.160.140.07^{+0.16}_{-0.14}0.07 start_POSTSUPERSCRIPT + 0.16 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.14 end_POSTSUBSCRIPT
Star
Mass Msubscript𝑀M_{\star}italic_M start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT (M) 𝒩(1.29,0.02,0.0,3.0)subscript𝒩1.290.020.03.0\mathcal{N_{B}}(1.29,0.02,0.0,3.0)caligraphic_N start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT ( 1.29 , 0.02 , 0.0 , 3.0 ) 1.29±0.02plus-or-minus1.290.021.29\pm 0.021.29 ± 0.02
Radius Rsubscript𝑅R_{\star}italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT (R) 𝒩(1.31,0.04,0.0,3.0)subscript𝒩1.310.040.03.0\mathcal{N_{B}}(1.31,0.04,0.0,3.0)caligraphic_N start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT ( 1.31 , 0.04 , 0.0 , 3.0 ) 1.31±0.04plus-or-minus1.310.041.31\pm 0.041.31 ± 0.04
Limb darkening coefficients
TESS filter u1subscript𝑢1u_{1}italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT Kipping (2013) 0.280.20+0.31subscriptsuperscript0.280.310.200.28^{+0.31}_{-0.20}0.28 start_POSTSUPERSCRIPT + 0.31 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.20 end_POSTSUBSCRIPT
TESS filter u2subscript𝑢2u_{2}italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Kipping (2013) 0.310.37+0.34subscriptsuperscript0.310.340.370.31^{+0.34}_{-0.37}0.31 start_POSTSUPERSCRIPT + 0.34 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.37 end_POSTSUBSCRIPT
Photometry
TESSS10 transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.00013±0.00004plus-or-minus1.000130.000041.00013\pm 0.000041.00013 ± 0.00004
TESSS37 transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.00010±0.00004plus-or-minus1.000100.000041.00010\pm 0.000041.00010 ± 0.00004
TESSS64 transit normalisation factor f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 𝒩(1.0,0.1)𝒩1.00.1\mathcal{N}(1.0,0.1)caligraphic_N ( 1.0 , 0.1 ) 1.00018±0.00003plus-or-minus1.000180.000031.00018\pm 0.000031.00018 ± 0.00003
TESSS10 dilution correction factor d𝑑ditalic_d 𝒩(0.0,0.2)𝒩0.00.2\mathcal{N}(0.0,0.2)caligraphic_N ( 0.0 , 0.2 ) 0.1±0.1plus-or-minus0.10.10.1\pm 0.10.1 ± 0.1
TESSS37 dilution correction factor d𝑑ditalic_d 𝒩(0.0,0.2)𝒩0.00.2\mathcal{N}(0.0,0.2)caligraphic_N ( 0.0 , 0.2 ) 0.1±0.1plus-or-minus0.10.10.1\pm 0.10.1 ± 0.1
TESSS64 dilution correction factor d𝑑ditalic_d 𝒩(0.0,0.2)𝒩0.00.2\mathcal{N}(0.0,0.2)caligraphic_N ( 0.0 , 0.2 ) 0.2±0.1plus-or-minus0.20.1-0.2\pm 0.1- 0.2 ± 0.1
TESSS10 log (Jitter) 𝒩(6.67,1.0)𝒩superscript6.671.0\mathcal{N}(-6.67^{\dagger},1.0)caligraphic_N ( - 6.67 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 1.0 ) 8.70.5+0.4subscriptsuperscript8.70.40.5-8.7^{+0.4}_{-0.5}- 8.7 start_POSTSUPERSCRIPT + 0.4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.5 end_POSTSUBSCRIPT
TESSS37 log (Jitter) 𝒩(6.02,1.0)𝒩superscript6.021.0\mathcal{N}(-6.02^{\dagger},1.0)caligraphic_N ( - 6.02 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 1.0 ) 8.30.5+0.4subscriptsuperscript8.30.40.5-8.3^{+0.4}_{-0.5}- 8.3 start_POSTSUPERSCRIPT + 0.4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.5 end_POSTSUBSCRIPT
TESSS64 log (Jitter) 𝒩(5.52,1.0)𝒩superscript5.521.0\mathcal{N}(-5.52^{\dagger},1.0)caligraphic_N ( - 5.52 start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT , 1.0 ) 8.70.4+0.3subscriptsuperscript8.70.30.4-8.7^{+0.3}_{-0.4}- 8.7 start_POSTSUPERSCRIPT + 0.3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.4 end_POSTSUBSCRIPT
HARPS RVs
log(K)𝐾\log{(K)}roman_log ( italic_K ) (m s-1) 𝒰(0.0,5.0)𝒰0.05.0\mathcal{U}(0.0,5.0)caligraphic_U ( 0.0 , 5.0 ) 3.52±0.05plus-or-minus3.520.053.52\pm 0.053.52 ± 0.05
Offset (m s-1) 𝒩(0.0,100.0)𝒩0.0100.0\mathcal{N}(0.0,100.0)caligraphic_N ( 0.0 , 100.0 ) 0.7±1.6plus-or-minus0.71.60.7\pm 1.60.7 ± 1.6
Linear trend (m s-1 d-1) 𝒩(0.0,10.0)𝒩0.010.0\mathcal{N}(0.0,10.0)caligraphic_N ( 0.0 , 10.0 ) 0.45±0.28plus-or-minus0.450.280.45\pm 0.280.45 ± 0.28
Quadratic trend (m s-1 d-2) 𝒩(0.0,1.0)𝒩0.01.0\mathcal{N}(0.0,1.0)caligraphic_N ( 0.0 , 1.0 ) 0.09±0.07plus-or-minus0.090.07-0.09\pm 0.07- 0.09 ± 0.07
log(Jitter)Jitter\log{(\rm{Jitter)}}roman_log ( roman_Jitter ) (m s-1) 𝒰(5.0,3.0)𝒰5.03.0\mathcal{U}(-5.0,3.0)caligraphic_U ( - 5.0 , 3.0 ) 2.1±2.0plus-or-minus2.12.0-2.1\pm 2.0- 2.1 ± 2.0
  • Distributions:

  • 𝒩(μ,σ)𝒩𝜇𝜎\mathcal{N}(\mu,\sigma)caligraphic_N ( italic_μ , italic_σ ): a normal distribution with a mean μ𝜇\muitalic_μ and a standard deviation σ𝜎\sigmaitalic_σ;

  • 𝒩(μ,σ,a,b)subscript𝒩𝜇𝜎𝑎𝑏\mathcal{N_{B}}(\mu,\sigma,a,b)caligraphic_N start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT ( italic_μ , italic_σ , italic_a , italic_b ): a bounded normal distribution with a mean μ𝜇\muitalic_μ, a standard deviation σ𝜎\sigmaitalic_σ, a lower bound a𝑎aitalic_a, and an upper bound b𝑏bitalic_b (bounds optional);

  • 𝒰(a,b)𝒰𝑎𝑏\mathcal{U}(a,b)caligraphic_U ( italic_a , italic_b ): a uniform distribution between a𝑎aitalic_a and b𝑏bitalic_b;

  • Distributions for limb darkening coefficients u1subscript𝑢1u_{1}italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and u2subscript𝑢2u_{2}italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are built into the exoplanet package and based on Kipping (2013).

  • Prior values:

  • equivalent to 0.5(log(D))+log(R)0.5𝐷subscript𝑅0.5(\log{(D)})+\log{(R_{\star})}0.5 ( roman_log ( italic_D ) ) + roman_log ( italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ) where D𝐷Ditalic_D is the transit depth (ppm multiplied by 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT) and Rsubscript𝑅R_{\star}italic_R start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the mean of the prior on the stellar radius (R);

  • equivalent to the log of the error median on the data of the correspondent instrument.

Refer to caption
Figure 19: Posterior distributions for the dilution correction factors of the TOI-2374 TESS light curves contrasted with the prior distributions.
Refer to caption
Figure 20: Posterior distributions for the dilution correction factors of the TOI-3071 light curves contrasted with the prior distributions.

Appendix D HEAVY ELEMENT MASS

The cooling curves for TOI-2374 b and TOI-3071 b described in section 4.2 are shown in Fig. 21, together with the planet radii and stellar ages from observations. The inferred heavy-element mass fraction posteriors described in the same section are shown in Figure 22

Refer to caption
Figure 21: Cooling curves for TOI-2374 b (dark blue) and TOI-3071 b (yellow). The solid lines assume they have the same composition as their stars. The shaded regions are cooling curves using the best-fit inferred compositions within 1σ1𝜎1\sigma1 italic_σ. The error bars show the observational data. Both planets are much smaller than the predictions from the evolution models, suggesting super-stellar bulk metallicities.
Refer to caption
Figure 22: Inferred heavy-element mass fraction posteriors for the two planets. For TOI-3071b, the two curves show the results for the two different stellar age estimates (chemical clocks and gyrochronology), and the filled area is their combined posterior.

Appendix E Author affiliations

1International Center for Advanced Studies (ICAS) and ICIFI (CONICET), ECyT-UNSAM, Campus Miguelete, 25 de Mayo y Francia, (1650) Buenos Aires, Argentina
2Department of Physics, University of Warwick, Gibbet Hill Road, Coventry CV4 7AL, UK
3Centre for Exoplanets and Habitability, University of Warwick, Gibbet Hill Road, Coventry CV4 7AL, UK
4Department of Astrophysics, University of Zürich, Winterthurerstrasse 190, 8057 Zürich, Switzerland
5Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, CAUP, Rua das Estrelas, 4150-762 Porto, Portugal
6Departamento de Física e Astronomia, Faculdade de Ciências, Universidade do Porto, Rua do Campo Alegre, 4169-007 Porto, Portugal
7Department of Physics and Astronomy, Vanderbilt University, Nashville, TN 37235, USA
8Center for Astrophysics | Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA
9Department of Astrophysical Sciences, Princeton University, 4 Ivy Lane, Princeton, NJ 08544, USA
10Observatoire de Genève, Université de Genève, Chemin Pegasi 51, 1290 Versoix, Switzerland
11Earth and Planets Laboratory, Carnegie Institution for Science, 5241 Broad Branch Road, NW, Washington, DC 20015, USA
12Observatories of the Carnegie Institution for Science, 813 Santa Barbara Street, Pasadena, CA 91101, USA
13NASA Ames Research Center, Moffett Field, CA 94035, USA
14Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
15Centro de Astrobiología (CAB, CSIC-INTA), Depto. de Astrofísica, ESAC campus, 28692, Villanueva de la Cañada (Madrid), Spain
16NASA Exoplanet Science Institute, IPAC, California Institute of Technology, Pasadena, CA 91125 USA
17University Observatory Munich, Ludwig-Maximilians-Universität, Scheinerstr. 1, 81679 Munich, Germany
18European Southern Observatory, Karl-Schwarzschild-Str. 2, 85748 Garching, Germany
19Department of Physics and Astronomy, McMaster University, 1280 Main Street West, Hamilton, Ontario, L8S 4L8, Canada
20Université Grenoble Alpes, CNRS, IPAG, 38000 Grenoble, France
21Las Campanas Observatory, Carnegie Institution of Washington, Colina el Pino, Casilla 601 La Serena, Chile
22Brierfield Observatory Bowral, NSW, 2576, Australia 23Department of Physics and Kavli Institute for Astrophysics and Space Research, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
24Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
25Department of Aeronautics and Astronautics, MIT, 77 Massachusetts Avenue, Cambridge, MA 02139, USA
26Kotizarovci Observatory, Sarsoni 90, 51216 Viskovo, Croatia
27Perth Exoplanet Survey Telescope, Perth, 6152 Western Australia
28Planetary Discoveries, 28935 Via Adelena, Santa Clarita, CA 91354, USA
29SETI Institute, Mountain View, CA 94043 USA
30 Department of Physics, Engineering and Astronomy, Stephen F. Austin State University, 1936 North St, Nacogdoches, TX 75962, USA