TESS Delivers Five New Hot Giant Planets Orbiting Bright Stars from the Full-frame Images

, , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , and

Published 2021 March 25 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Joseph E. Rodriguez et al 2021 AJ 161 194 DOI 10.3847/1538-3881/abe38a

Download Article PDF
DownloadArticle ePub

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

1538-3881/161/4/194

Abstract

We present the discovery and characterization of five hot and warm Jupiters—TOI-628 b (TIC 281408474; HD 288842), TOI-640 b (TIC 147977348), TOI-1333 b (TIC 395171208, BD+47 3521A), TOI-1478 b (TIC 409794137), and TOI-1601 b (TIC 139375960)—based on data from NASA's Transiting Exoplanet Survey Satellite (TESS). The five planets were identified from the full-frame images and were confirmed through a series of photometric and spectroscopic follow-up observations by the TESS Follow-up Observing Program Working Group. The planets are all Jovian size (RP = 1.01–1.77 RJ) and have masses that range from 0.85 to 6.33 MJ. The host stars of these systems have F and G spectral types (5595 ≤ Teff ≤ 6460 K) and are all relatively bright (9.5 < V < 10.8, 8.2 < K < 9.3), making them well suited for future detailed characterization efforts. Three of the systems in our sample (TOI-640 b, TOI-1333 b, and TOI-1601 b) orbit subgiant host stars ($\mathrm{log}$ g < 4.1). TOI-640 b is one of only three known hot Jupiters to have a highly inflated radius (RP > 1.7 RJ, possibly a result of its host star's evolution) and resides on an orbit with a period longer than 5 days. TOI-628 b is the most massive, hot Jupiter discovered to date by TESS with a measured mass of ${6.31}_{-0.30}^{+0.28}$ MJ and a statistically significant, nonzero orbital eccentricity of e = ${0.074}_{-0.022}^{+0.021}$. This planet would not have had enough time to circularize through tidal forces from our analysis, suggesting that it might be remnant eccentricity from its migration. The longest-period planet in this sample, TOI-1478 b (P = 10.18 days), is a warm Jupiter in a circular orbit around a near-solar analog. NASA's TESS mission is continuing to increase the sample of well-characterized hot and warm Jupiters, complementing its primary mission goals.

Export citation and abstract BibTeX RIS

1. Introduction

The discovery of hot Jupiters, combined with the assumption that gas giant planets must form at separations from their host star similar to our own giant planets, indicated that giant planets likely undergo large-scale migration from their formation locations. Various mechanisms have been proposed to place giant planets into very short-period orbits (Goldreich & Tremaine 1980; Lin & Papaloizou 1986; Lin et al. 1996; see Dawson & Johnson 2018 for a detailed review). However, it is not clear which of these mechanisms (if any) are dominant or govern this migration, or whether hot Jupiters can form in situ (Batygin et al. 2016), obviating the need for large-scale migration. One possibility is that giant planets migrate slowly and smoothly within the circumstellar gas–dust disk, resulting in well-aligned, nearly circular orbits (D'Angelo et al. 2003). It is also thought that planetary migration may be heavily influenced by gravitational interactions with other bodies within the system. These interactions result in highly eccentric and misaligned orbits (relative to the rotation axis of the star) and are typically referred to as a type of "high-eccentricity migration" (HEM) or "Kozai–Lidov" (Lidov 1962; Kozai 1962; Rasio & Ford 1996; Wu & Murray 2003; Fabrycky & Tremaine 2007; Nagasawa & Ida 2011; Wu & Lithwick 2011; Naoz 2016). For short-period hot Jupiters, with periods less than about 5 days, the orbits will circularize in only a few billion years, erasing the evidence of HEM. Additionally, these interactions can cause misalignments in the planet's orbital plane (relative to the original disk plane) that can remain present for much longer, and such misalignments can be detected through Doppler spectroscopy, using observations of the Rossiter–McLauglin effect (McLaughlin 1924; Rossiter 1924) or Doppler tomography (e.g., Miller et al. 2010; Johnson et al. 2014; Zhou et al. 2016). Longer-period, hot (P ≳ 5 days) and warm (P > 10 days) Jupiters experience smaller tidal forces, preserving their orbital eccentricities. It is likely that multiple mechanisms shape the short-period giant planet population, and studying these longer-period hot Jupiters may give clues to their common evolutionary pathway.

While some ground-based transit surveys that were dedicated to discovering hot Jupiters had near 24 hr coverage (Bakos et al. 2013), in general they struggled to discover planets with periods ≳5 days due to the poor duty cycle from weather and only being able to observe at night (Gaudi et al. 2005). Additionally, many of the first hot Jupiters discovered were assumed to reside in circular orbits when analyzing the observations, an assumption that may confuse current efforts to understand migration. Fortunately, NASA's Transiting Exoplanet Survey Satellite (TESS) mission was launched in 2018 April and completed its primary 2 yr long mission in 2020 July (Ricker et al. 2015). TESS was awarded a 27 month first extended mission in which it will not only reobserve some areas covered in the primary mission but also observe most of the ecliptic plane, nearly completing coverage of the entire sky. TESS has a minimum observing baseline of ∼27 days, and from recent occurrence-rate studies, TESS planet searches will be mostly complete for hot Jupiters with periods ⪅10 days (Zhou et al. 2019). Therefore, TESS provides a great resource for the discovery and confirmation of new longer-period hot and warm Jupiters (5 < P < 15 days), where eccentricities from migration would not be completely erased by tidal forces. TESS has already discovered a number of statistically significant hot Jupiters with highly eccentric (e > 0.2) orbits like HD 2685 b (Jones et al. 2019), TOI-172 b (Rodriguez et al. 2019), TOI-150 b (Kossakowski et al. 2019), TIC 237913194 b (Schlecker et al. 2020), and TOI-559 b (Ikwut-Ukwa et al. 2021). Additionally, TESS recently confirmed that the hot Jupiter HD 118203 b, an RV-identified planet with high eccentricity discovered with the radial velocity method (da Silva et al. 2006), transits its host star (Pepper et al. 2020).

TESS will also provide the ability to study hot-Jupiter reinflation since its high photometric precision will allow it to discover giant planets around larger, more evolved host stars. As a star evolves off the main sequence, the stellar irradiation received by warm Jupiters is similar to that of a hot Jupiter. Therefore, discovering gas giant planets orbiting evolved stars at longer periods (10 s of days) can test whether this increased irradiation causes the same inflation seen for short-period hot Jupiters (Lopez & Fortney 2016). Most warm Jupiters orbiting main-sequence stars show little to no inflation (Demory & Seager 2011), suggesting that this energy must be transferred deep into the planet's interior (Liu et al. 2008; Spiegel & Burrows 2013), and as the star evolves, these warm Jupiters may reinflate from the increased irradiation. Recent discoveries of hot Jupiters orbiting evolved stars are suggestive of reinflation (Almenara et al. 2015; Grunblatt et al. 2016; Hartman & Bakos 2016; Stevens et al. 2017), and TESS has already found a few hot and warm Jupiters orbiting evolved stars (Brahm et al. 2019; Huber et al. 2019; Nielsen et al. 2019; Rodriguez et al. 2019; Wang et al. 2019; Sha et al. 2021).

In this paper we present the discovery of TOI-628 b, TOI-640 b, TOI-1333 b, TOI-1478 b, and TOI-1601 b, five new hot and warm Jupiters from NASA's TESS mission. All five planets were discovered from an analysis of the 30 minute cadence full-frame images (FFIs) and first identified as TESS Objects of Interest (TOIs) by the TESS Science Office. 71 These five new systems increase the known sample of well-characterized hot Jupiters, particularly those with longer orbital periods (>5 days). In Section 2 we present our time-series photometric and spectroscopic observations obtained by the TESS Follow-up Observing Program (TFOP) Working Group (WG) and describe the high-spatial-resolution imaging of all five targets, specifically on TOI-1333 and its two nearby companions. Our methodology for our global modeling using EXOFASTv2 (Eastman et al. 2019) is summarized in Section 3. We place these systems in context with the known population of hot and warm Jupiters and discuss the impact of TESS on the discovery of giant planets in Section 7, and our conclusions are given in Section 5.

2. Observations and Archival Data

We used a series of photometric and spectroscopic observations to confirm and characterize the new systems, including high-spatial-resolution imaging to rule out false positives, confirm them as bona fide planets, and measure key parameters such as orbital eccentricity and the planet's density. See Table 1 for a list of the literature kinematics and magnitudes for each target.

2.1. TESS Photometry

The initial detection of the new planets came from data collected by the TESS mission. TESS images the sky with a 24° × 96° field of view and observes the same stars for about a month before moving on to observe a different region. During its primary mission, TESS saved and downloaded images of a smaller number of preselected stars every two minutes, while downloading coadded images of its entire field of view every 30 minutes. None of the planets described in this paper orbit stars that were preselected for two-minute cadence observations, so we use data from the 30 minute cadence FFIs.

After the data were transferred from the orbiting spacecraft back to Earth, the FFIs were calibrated using the TICA software (M. Fausnaugh et al. 2021, in preparation), and light curves for a set of stars complete down to TESS-band magnitude = 13.5 were extracted with the MIT Quick Look Pipeline (QLP; Huang et al. 2020b). The QLP extracts light curves from the FFIs using a difference image analysis technique. After removing the effects of scattered earthshine and moonshine from the images, the QLP subtracts a high-quality reference image from each individual science frame and measures difference fluxes within sets of photometric apertures surrounding each star in the image. These difference fluxes are then converted to absolute brightness measurements by adding back the median flux expected from each star based on its TESS-band magnitude. The QLP light curves have been used to discover dozens of planets (e.g., Huang et al. 2018; Rodriguez et al. 2019; Huang et al. 2020a) and a few thousand planet candidates (Guerrero et al. 2021). Additional information and a description of the QLP procedures are given by Huang et al. (2020b). The QLP light curves are shown in Figure 1. All five of these light curves were flattened using Keplerspline, 72 a spline fitting routine to divide out the best-fit stellar variability (Vanderburg & Johnson 2014). The spacing of the spline break points for each system was determined by minimizing the Bayesian information criterion following the methodology from Shallue & Vanderburg (2018). After removing the stellar variability, we keep all data from one full transit duration before the transit until one full transit duration after the transit, and we discard the remaining baseline data (which contains very little useful information but is computationally expensive to model). We use these light-curve segments for the global fitting of each system (see Section 3).

Figure 1.

Figure 1. Raw TESS QLP 30 minute light curves for (top left) TOI-628, TOI-640 (top right), TOI-1333 (middle left), TOI-1478 (middle right), and TOI-1601 (bottom). Transits highlighted in gray were excluded from the global fit since they were flagged as bad quality by the QLP pipeline (Huang et al. 2020b).

Standard image High-resolution image

Because the QLP uses aperture photometry on difference frames using a median-combined reference image, it does not always measure absolute transit depths, only difference fluxes. Transit depths measured by QLP, especially in crowded fields, might be dependent on the accuracy of the TESS-band magnitudes from the TESS Input Catalog (Stassun et al. 2018) to correct for contamination due to blending of nearby stars within the aperture. To check that the transit depths measured by QLP for these targets are reasonably accurate, we performed a spot check for TOI-1601 with light curves extracted using a more traditional simple aperture photometry method (Vanderburg et al. 2019). We did not deblend the photometry in this test because TOI-1601 is in a relatively sparse field, as shown from our high-resolution imaging (see Section 2.6.2). We ran an independent EXOFASTv2 fit for TOI-1601 b (swapping this light curve for the QLP one) following the strategy discussed in Section 3, and the results were consistent to within <1σ uncertainties. Additionally, for each system, we used the follow-up ground-based photometry within the global analysis to also constrain the depth, providing independent constraints that can be used to confirm the QLP depths (since the TFOP photometry is at a higher angular resolution than TESS). Within our global fit, we checked on any unknown contamination by fitting for a dilution term on the TESS bandpass to account for any difference compared to the SG1 photometry. In all cases other than TOI-1478, the fitted dilution was consistent with zero and well within our Gaussian 10% prior around zero, showing clear consistency between TESS and the TFOP seeing-limited photometry. TOI-1478 showed a significant Required dilution on the order of 12%. To properly account for this, we removed the TESS dilution prior (see Section 3 for details), allowing it to be a free parameter, to properly correct for this within the fit.

We searched the nonflattened QLP light curves for rotation-based modulations using the VARTOOLS Lomb–Scargle function (Hartman & Bakos 2016). Specifically, we searched from 0.1 to 30 days and detect a clear, strong periodicity at 5.296 days for TOI-1333, significantly different from the orbital period of its planetary companion (Pb = 4.72 days). This same periodicity is observed in our ground-based photometry, ruling out any systematics in the TESS observations. We see some tentative evidence of a periodicity at 10–11 days for TOI-628, but it was only observed in one TESS sector.

2.2. WASP Photometry

The WASP transit search consisted of two wide-field arrays of eight cameras, with SuperWASP on La Palma covering the northern sky and WASP-South in South Africa covering the south (Pollacco et al. 2006). Each camera used a 200 mm, f/1.8 lens with a broadband filter spanning 400–700 nm, backed by 2048 × 2048 CCDs giving a plate scale of 137 pixel−1. Observations then rastered available fields with a typical 15 minute cadence.

We searched the WASP data for any rotational modulations using the methods from Maxted et al. (2011). TOI-640 was observed for spans of 150 nights in each of four years. The data from 2008, 2009, and 2010 show no significant modulation. The data from 2007, however, show significant power at a period of 62 ± 5 days, with an amplitude of 3 mmag and an estimated false-alarm probability below 1%. Since this is seen in only one season, and given that the data span only 2.5 cycles, we do not regard this detection as fully reliable. For TOI-1333, the WASP data span 130 days in 2007 and show a clear modulation with a period of 15.9 ± 0.3 days, an amplitude of 19 mmag, and a false-alarm probability below 1%. A similar periodicity is also seen in the TESS data and in the Kilodegree Extremely Little Telescope (KELT) data (see below) but at one-third of this period at 5.3 days. No significant periodicity was detected for TOI-1478 or TOI-1601.

Given knowledge of the TESS detections, transits of three of the systems described here can be found readily in the WASP data. TOI-640 was observed between 2006 and 2012, accumulating 23,000 data points. The WASP search algorithm (Collier Cameron et al. 2007) finds the transit with a period of 5.003773 ± 0.000041 and a midtransit epoch (TC ) of 2454822.00318 ± 0.00411 HJDTDB. This detection had been overlooked by WASP vetters owing to the near-integer day period (5.00 days), since the dominant red noise in the WASP data is at multiples of a day. TOI-1478 was observed between 2009 and 2012, accumulating 9000 data points, less than usual for WASP since the field is near the crowded Galactic plane. It had not been flagged as a WASP candidate, but the search algorithm finds the transit with a period of 10.18051 ± 0.00017 days and a TC of 2455696.36710 ± 0.00492 HJDTDB. TOI-1601 was observed over 2006 and 2007, accumulating 10,400 data points. The search algorithm finds the transit and gets a period of 5.33197 ± 0.00010 and a TC of 2454186.65253 ± 0.01283 HJDTDB. We use these TC values as priors for the EXOFASTv2 global fits of TOI-640 b, TOI-1478 b, and TOI-1601 b. We see an ∼40% reduction in uncertainty on the period of the planet when including the WASP TC prior.

2.3. KELT Photometry

To complement the TESS photometry, we analyzed observations of these five TOIs from the KELT survey 73 (Pepper et al. 2007, 2012, 2018). For a full description of the KELT observing strategy and reduction process, see Siverd et al. (2012) and Kuhn et al. (2016). KELT has two fully robotic telescopes, each of which uses a Mamiya 645 80 mm f/1.9 lens with 42 mm aperture and Apogee 4 k × 4 k CCD on a Paramount ME mount. This setup provides a 26° × 26° field of view with a 23'' pixel scale. The two telescopes are located at Winer Observatory in Sonoita, AZ and at the South African Astronomical Observatory (SAAO) in Sutherland, South Africa. KELT covers ∼85% of the entire sky, and the observing strategy results in a 20–30 minute cadence. Some of the KELT light curves are publicly available through the NASA Exoplanet Archive. KELT observations were available for TOI-628, TOI-1333, TOI-1478, and TOI-1601. KELT-South observed TOI-628 2828 times and TOI-1478 4632 times from 2010 to 2015. KELT-North observed TOI-1333 from 2012 to 2014 and TOI-1601 from 2006 to 2014, acquiring 2580 and 8520 observations, respectively. Since KELT has been observing since 2006 in some cases, the observations significantly extend the baseline of the photometry and can provide a strong constraint on the ephemeris of each system. Following the strategy described in Siverd et al. (2012) and Kuhn et al. (2016), we also search the KELT light curves for transits of each planet. Unfortunately, no significant signs of the known planetary transits were found, likely due to the poor duty cycle for longer orbital periods.

Following the approach of Stassun et al. (1999) and Oelkers et al. (2018), we executed a search for periodic signals most likely to come from the rotation period of the star. For these stars, we postprocessed the light-curve data using the Trend-Filtering Algorithm (Kovács et al. 2005) to remove common systematics. We then searched for candidate rotation signals using a modified version of the Lomb–Scargle period-finder algorithm (Lomb 1976; Scargle 1982). We searched for periods between a minimum period of 0.5 days and a maximum period of 50 days using 2000 frequency steps. 74 We masked periods between 0.5 and 0.505 days and 0.97–1.04 days to avoid the most common detector aliases associated with KELT's observational cadence and its interaction with the periods for the solar and sidereal day. For each star, we selected the highest statistically significant peak of the power spectrum (hereafter γ) as the candidate period.

We then executed a bootstrap analysis, using 1000 Monte Carlo iterations, where the dates of the observations were not changed but the magnitude values of the light curve were randomized, following the work of Henderson & Stassun (2012). We recalculated the Lomb–Scargle power spectrum for each iteration and recorded the maximum peak power (hereafter ${\gamma }_{\mathrm{sim}}$) of all iterations. If the highest power spectrum peak was larger than the maximum simulated peak ($\gamma \gt {\gamma }_{\mathrm{sim}}$) after 1000 iterations, we considered the periodic signal to be a candidate rotation period.

We find TOI-1333 to have a strongly significant (γ > 50) candidate rotation period at 5.3 days, which is consistent with TESS (see Section 2.1), and TOI-1601 and TOI-1478 to have weakly significant (γ > 10) candidate rotation periods of 9.3 and 16.6 days, respectively. However, we do not see these periodicities in the TESS photometry, and they are likely aliases of the KELT observing strategy. KELT did not obtain observations of TOI-640.

Table 1. Literature and Measured Properties

Other Identifiers
  TOI-628TOI-640TOI-1333TOI-1478TOI-1601
  TIC 281408474TIC 147977348TIC 395171208TIC 409794137TIC 139375960
  HD 288842BD+47 3521A
 TYCHO-2TYC 0146-01523-1TYC 7099-00846-1TYC 3595-01186-1TYC 5440-01407-1TYC 2836-00689-1
 2MASSJ06370314+0146031J06385630-3638462J21400351+4824243J08254410-1333356J02332674+4100483
 TESS Sector66,715,167,818
ParameterDescriptionValueValueValueValueValueSource
αJ2000 a R.A. (R.A.)06:37:03.1360706:38:56.3074221:40:03.5039808:25:44.1070802:33:26.746831
δJ2000 a Decl. (decl.)01:46:03.19552−36:38:46.1442548:24:24.52541−13:33:35.42756+41:00:48.368931
BT Tycho BT mag.10.782 ± 0.05111.178 ± 0.059.933 ± 0.02411.618 ± 0.08211.521 ± 0.0672
VT Tycho VT mag.10.176 ± 0.04110.574 ± 0.0439.487 ± 0.02110.805 ± 0.06710.710 ± 0.0512
G Gaia G mag.10.0579 ± 0.0210.4006 ± 0.029.35 ± 0.0210.66 ± 0.0210.53 ± 0.021
BP Gaia BP mag.10.38 ± 0.0210.68 ± 0.029.59 ± 0.0211.03 ± 0.0210.86 ± 0.021
RP Gaia RP mag.9.61 ± 0.029.99 ± 0.028.99 ± 0.0210.16 ± 0.0210.06 ± 0.021
T TESS mag.9.6565 ± 0.006610.0367 ± 0.0069.03527 ± 0.006110.2042 ± 0.00610.1035 ± 0.00653
J 2MASS J mag.9.170 ± 0.0419.519 ± 0.0248.485 ± 0.0279.590 ± 0.0239.505 ± 0.0224
H 2MASS H mag.8.895 ± 0.0579.327 ± 0.0268.397 ± 0.0439.255 ± 0.0269.266 ± 0.0214
KS 2MASS KS mag.8.811 ± 0.029.243 ± 0.0238.272 ± 0.0249.201 ± 0.0219.19 ± 0.024
WISE1WISE1 mag.8.76 ± 0.039.213 ± 0.037.706 ± 0.0139.15 ± 0.039.16 ± 0.035
WISE2WISE2 mag.8.79 ± 0.039.240 ± 0.037.594 ± 0.0129.19 ± 0.039.21 ± 0.035
WISE3WISE3 mag.8.59 ± 0.039.239 ± 0.038.035 ± 0.0199.18 ± 0.039.18 ± 0.0345
WISE4WISE4 mag.8.23 ± 0.198.043 ± 0.1638.80 ± 0.335
μα Gaia DR2 proper motion−1.437 ± 0.117−3.872 ± 0.040−9.810 ± 0.050−8.277 ± 0.05821.708 ± 0.0951
  in R.A. (mas yr−1)  
μδ Gaia DR2 proper motion−1.082 ± 0.0874.927 ± 0.045−10.501 ± 0.04667.943 ± 0.046−0.874 ± 0.0901
  in decl. (mas yr−1)  
$v\sin {i}_{\star }$ Rotational velocity (km s−1) 6.9 ± 0.56.1 ± 0.514.2 ± 0.54.3 ± 0.56.4 ± 0.5Sections 2.5.2, 2.5.2 and 2.5.3
vmac Macroturbulent broadening (km s−1) 5.4 ± 0.76.32 ± 1.377.4 ± 1.84.9 ± 0.56.3 ± 0.6Sections 2.5.2, 2.5.2 and 2.5.3
π b Gaia parallax (mas)5.601 ± 0.1032.925 ± 0.0334.989 ± 0.0386.542 ± 0.0472.974 ± 0.0801
PRot Rotation period (days)  5.3 ± 0.159  Sections 2.1, 2.2, and 2.3

Notes. The uncertainties of the photometry have a systematic error floor applied.

a R.A. and decl. are in epoch J2000. The coordinates come from Vizier, where the Gaia R.A. and decl. have been precessed and corrected to J2000 from epoch J2015.5. b Values have been corrected for the −0.30 μas offset as reported by Lindegren et al. (2018).

References are (1) Gaia Collaboration et al. (2018), (2) Høg et al. (2000), (3) Stassun et al. (2018), (4) Cutri et al. (2003), (5) Zacharias et al. (2017).

Download table as:  ASCIITypeset image

2.4. Ground-based Photometry from the TESS Follow-up Observing Program Working Group

To refine the ephemerides and transit parameters for each system while ruling out false-positive scenarios, we obtained photometric follow-up observations on all five systems from the TESS Follow-up Observing Program Working Group 75 subgroup 1 (SG1) for seeing-limited photometry. Specifically, the follow-up comes from the Las Cumbres Observatory (LCO) telescope network (Brown et al. 2013), Whitin Observatory at Wellesley College, KeplerCam on the 1.2 m telescope at Fred Lawrence Whipple Observatory (FLWO), Brierfield Observatory, PEST Observatory, C. R. Chambliss Astronomical Observatory (CRCAO) at Kutztown University, Adams Observatory at Austin College, and Suto Observatory. To schedule the photometric transit follow-up observations, we used the TAPIR software package (Jensen 2013). The data reduction and aperture photometry extraction was performed using AstroImageJ (Collins et al. 2017) for all follow-up data other than the PEST observations, which were done using a custom software package PEST Pipeline. 76 The TFOP follow-up photometry for these five systems is shown in Figure 2, and a list of each telescope's information and details on each follow-up transit can be seen in Table 2. The follow-up transits presented in Figure 2 are available as machine-readable tables with this paper.

Figure 2.

Figure 2. TESS (orange) and TFOP SG1 follow-up transits of TOI-628 b (top left), TOI-640 b (top right), TOI-1333 b (middle left), TOI-1478 b (middle right), and TOI-1601 b (bottom). The EXOFASTv2 model for each transit observation is shown by the red solid line. The TFOP photometry is available as the Data behind the Figure.(The data used to create this figure are available.)

Standard image High-resolution image

Table 2. Photometric Follow-up Observations of These Systems Used in the Global Fits and the Detrending Parameters

TargetObservatoryDate (UT)Size (m)FilterFOVPixel ScaleExp (s)Additive Detrending
TOI-628 bFLWO/KeplerCam2019 Dec 61.2 ${z}^{{\prime} }$ 231 × 23106729None
TOI-628 bLCO TFN2019 Dec 160.4 ${z}^{{\prime} }$ 19' × 29'05718air mass
TOI-628 bWhitin2020 Feb 80.7 ${r}^{{\prime} }$ 24' × 24'06710air mass
TOI-640 bBrierfield2019 Dec 270.36 I 494 × 49414790X(FITS), Y(FITS)
TOI-640 bPEST2020 Mar 10.3048 R 31' × 21'1260none
TOI-640 bLCO SSO2020 Aug 231.0 ${z}^{{\prime} }$ 27' × 27'03960Y(FITS)
TOI-640 bLCO SSO2020 Nov 61.0 ${z}^{{\prime} }$ 27' × 27'03960air mass
TOI-1333 bCRCAO2020 Jul 290.6096 I 268 × 26803945air mass
TOI-1333 bLCO McDonald2020 Jul 290.4 ${z}^{{\prime} }$ 19' × 29'05730total counts
TOI-1333 bLCO McDonald2020 Aug 120.4 ${z}^{{\prime} }$ 19' × 29'05730Y(FITS)
TOI-1333 bCRCAO2020 Sep 190.6096 ${z}^{{\prime} }$ 268 × 26803980air mass
TOI-1478 bFLWO/KeplerCam2019 Dec 141.2 ${i}^{{\prime} }$ 231 × 23106727none
TOI-1478 bPEST2020 Jan 30.3048 R 31' × 21'1230air mass
TOI-1601 bGMU2020 Aug 300.8 R 23' × 23'03430air mass, sky/pixels
TOI-1601 bGMU2020 Sep 150.8 R 23' × 23'03430air mass, sky/pixels, X(FITS)
TOI-1601 bCRCAO2020 Oct 10.6096 ${z}^{{\prime} }$ 268 × 26803990air mass
TOI-1601 bAdams2020 Oct 170.61 I 26' × 26'03860air mass, total counts
TOI-1601 bLCO McDonald2020 Oct 171.0 ${z}^{{\prime} }$ 27' × 27'03960air mass, sky/pixels

Note. All of the follow-up photometry presented in this paper is available as the Data behind Figure 2 in machine-readable form in the online journal. See Section D in the Appendix of Collins et al. (2017) for a description of each detrending parameter.

Download table as:  ASCIITypeset image

2.5. Spectroscopy

To rule out false-positive scenarios and measure the mass and orbital eccentricity of each system, we obtained time-series spectroscopy coordinated through the TFOP WGs. A sample of one radial velocity (RV) point per target per instrument is shown in Table 3, with the full table available in machine-readable form in the online journal. The RVs and best-fit models from our EXOFASTv2 analysis are shown in Figure 3 (see Section 3). Following the methodology in Zhou et al. (2018), we measure the $v\sin {I}_{* }$ and macroturbulent broadening for all five systems from the Tillinghast Reflector Echelle Spectrograph except TOI-640, where the CHIRON observations were used (see Table 1).

Figure 3.

Figure 3. RV observations of TOI-628 (top left), TOI-640 (top middle), TOI-1333 (top right), TOI-1478 (bottom left), and TOI-1601 (bottom right). In each case, the top figure shows the RVs vs. time, and the bottom panel is phased to the best-fit ephemeris from our global fit. The EXOFASTv2 model is shown in red, and the residuals to the best fit are shown below each plot. We see no periodicity in the residuals from our fit.

Standard image High-resolution image

Table 3. First RV Point from Each Spectrograph for All Five Systems

BJDTDB RV (m s−1) σRV a (m s−1)TargetInstrument
2458604.64716193.043.6TOI-628TRES
2458742.2527019775.217.5TOI-628MINERVA3
2458742.2527019700.814.3TOI-628MINERVA4
2458742.2527019795.741.0TOI-628MINERVA6
2458748.8948239168.029.4TOI-640CHIRON
2458777.69583−50.648.0TOI-1333TRES
2458852.8017219374.421.9TOI-1478CHIRON
2458875.6099420738.522.0TOI-1478CORALIE
2458910.7353720910.37.5TOI-1478FEROS
2458829.8652567.427.6TOI-1478TRES
2459184.8362919502.024.0TOI-1478CHIRON2
2458847.77575−184.132.0TOI-1601TRES

Note.

a The internal RV error for the observation shown.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

2.5.1. TCES Spectroscopy

Reconnaissance spectroscopic observations of TOI-628 were carried out with the Tautenburg Coudé Echelle Spectrograph (TCES) mounted at the 2 m "Alfred Jensch" Telescope of the Thuringian State Observatory (TLS) in Tautenburg, Germany. The spectra cover the 470–740 nm wavelength range and have a resolution R = 57,000. A 40 minute exposure was taken at BJD = 2458777.6053 (orbital phase ϕ ∼ 0.5), and a 3 × 20 minute exposure at BJD = 2458855.3720 (ϕ ∼ 0.0). We measured for the two single-lined spectra at ΔRV ≲ 1 km s−1, ruling out an eclipsing binary as the cause of the event detected by TESS. These velocities were not included in the global fit for TOI-628.

2.5.2. TRES Spectroscopy

To confirm targets from TESS in the Northern hemisphere, we observed TOI-628, TOI-1333, TOI-1478, and TOI-1601 with the Tillinghast Reflector Echelle Spectrograph (TRES; Fűrész 2008) 77 on the 1.5 m Tillinghast Reflector. The telescope is located at the Fred L. Whipple Observatory on Mt. Hopkins, AZ, and the spectrograph has a resolving power of R = 44,000. See Buchhave et al. (2010) and Quinn et al. (2012) for a detailed description of the reduction and RV extraction pipeline. The only difference in our analysis is that we created the template spectra for the RV extraction by aligning and median-combining all of the out-of-transit spectra. We removed cosmic rays and cross-correlated the median-combined spectra against all of the observed spectra. Bisector spans for the TRES spectra were calculated following the technique described in Torres et al. (2007). There was no correlation between the bisector spans and the RVs. We also used the TRES spectra to provide constraints on the Teff and $\left[\mathrm{Fe}/{\rm{H}}\right]$ for our global analysis. We analyzed the TRES spectra with the Stellar Parameter Classification (SPC) package (Buchhave et al. 2012) to determine the metallicity and rotational velocity for all five host stars (see Tables 1 and 4). We also used SPC to determine a constraint on the Teff of 6250 ± 100 K for TOI-1333, which is used in Section 2.6.4 to constrain the dilution from nearby companions and the radius of TOI-1333.

2.5.3. CHIRON Spectroscopy

We obtained a series of spectroscopic observations with the 1.5 m SMARTS/CHIRON facility (see Table 3; Tokovinin et al. 2013) for TOI-640 and TOI-1478 to measure the host star parameters and constrain their masses and eccentricities. The 1.5 m SMARTS facility is located at Cerro Tololo Inter-American Observatory (CTIO), Chile. CHIRON is a high-resolution echelle spectrograph fed via an image slicer through a single multimode fiber, with a spectral resolving power of R ∼ 80,000 over the wavelength region from 410 to 870 nm. For the case of TOI-1478, we treat the pre- and postshutdown RVs as separate instruments within the fit (see Section 3).

To obtain the stellar atmospheric parameters, we matched the CHIRON spectra against an interpolated library of ∼10,000 observed spectra classified by SPC (Buchhave et al. 2012). The metallicity from this analysis was used as a prior for the global fit of TOI-640 (see Table 4). RVs were derived via the least-squares deconvolution (Donati et al. 1997; Zhou et al. 2020) of the spectra against nonrotating synthetic templates matching the spectral parameters of each host star, generated using the ATLAS9 model atmospheres (Kurucz 1992). The RVs were then measured by fitting a least-squares deconvolution line profile with a rotational broadening kernel as prescribed by Gray (2005). The velocities for each system are presented in Table 3.

Table 4. Median Values and 68% Confidence Intervals for the Global Models

Priors:TOI-628 bTOI-640 bTOI-1333 b a TOI-1478 b b TOI-1601 b
Gaussian c Gaia parallax (mas)5.601 ± 0.1032.925 ± 0.0336.542 ± 0.0472.974 ± 0.080
Gaussian[Fe/H] Metallicity (dex)0.24 ± 0.080.02 ± 0.10.11 ± 0.080.07 ± 0.080.31 ± 0.08
Upper Limit AV V-band extinction (mag)2.9770.2920.1200.14694
Gaussian R Stellar radius ( R)1.963 ± 0.064
Gaussian Teff Stellar effective temperature (K)6250 ± 100
Gaussian TC d Time of conjunction (HJDTDB)2454822.00318 ± 0.004112455696.36710 ± 0.004922454186.65253 ± 0.01283
Gaussian e DT Dilution in TESS0.00000 ± 0.003440.00000 ± 0.003600.00000 ± 0.038170.0 ± 0.00196
Gaussian f DI Dilution in I 0.3609 ± 0.0180
Gaussian f DR Dilution in R 0.3499 ± 0.0175
Gaussian f ${D}_{{z}^{{\prime} }}$ Dilution in ${z}^{{\prime} }$ 0.0664 ± 0.0033
ParameterUnitsValues  
Stellar Parameters:    
Probability 100%100%100%100%68.4%31.6%
M* Mass ( M) ${1.311}_{-0.075}^{+0.066}$ ${1.536}_{-0.076}^{+0.069}$ ${1.464}_{-0.079}^{+0.076}$ ${0.947}_{-0.041}^{+0.059}$ ${1.517}_{-0.049}^{+0.053}$ ${1.340}_{-0.045}^{+0.042}$
R* Radius ( R) ${1.345}_{-0.040}^{+0.046}$ ${2.082}_{-0.058}^{+0.064}$ ${1.925}_{-0.063}^{+0.064}$ ${1.048}_{-0.029}^{+0.030}$ ${2.186}_{-0.063}^{+0.074}$ ${2.185}_{-0.080}^{+0.086}$
L* Luminosity ( L) ${2.48}_{-0.28}^{+0.41}$ ${6.83}_{-0.54}^{+0.41}$ ${5.17}_{-0.45}^{+0.49}$ ${0.971}_{-0.035}^{+0.036}$ ${5.40}_{-0.33}^{+0.35}$ ${5.25}_{-0.33}^{+0.35}$
FBol Bolometric flux × 10−9 (cgs) ${2.52}_{-0.27}^{+0.39}$ ${1.87}_{-0.14}^{+0.10}$ 1.32 ± 0.45 ${1.505}_{-0.61}^{+0.52}$ ${1.485}_{-0.61}^{+0.59}$
ρ* Density (cgs) ${0.762}_{-0.078}^{+0.067}$ ${0.240}_{-0.025}^{+0.024}$ ${0.289}_{-0.028}^{+0.031}$ ${1.16}_{-0.11}^{+0.14}$ ${0.205}_{-0.018}^{+0.016}$ ${0.181}_{-0.019}^{+0.020}$
$\mathrm{log}g$ Surface gravity (cgs) ${4.300}_{-0.036}^{+0.026}$ ${3.987}_{-0.036}^{+0.030}$ ${4.034}_{-0.033}^{+0.032}$ ${4.374}_{-0.032}^{+0.039}$ ${3.940}_{-0.025}^{+0.022}$ ${3.885}_{-0.031}^{+0.029}$
Teff Effective temperature (K) ${6250}_{-190}^{+220}$ ${6460}_{-150}^{+130}$ 6274 ± 97 ${5597}_{-82}^{+83}$ ${5948}_{-89}^{+87}$ ${5910}_{-99}^{+96}$
[Fe/H]Metallicity (dex) ${0.258}_{-0.080}^{+0.078}$ ${0.072}_{-0.076}^{+0.085}$ ${0.119}_{-0.077}^{+0.078}$ ${0.078}_{-0.066}^{+0.072}$ ${0.329}_{-0.075}^{+0.073}$ 0.299 ± 0.076
[Fe/H]0 g Initial metallicity ${0.271}_{-0.068}^{+0.070}$ ${0.174}_{-0.080}^{+0.088}$ ${0.212}_{-0.084}^{+0.083}$ ${0.110}_{-0.064}^{+0.069}$ 0.347 ± 0.069 ${0.296}_{-0.069}^{+0.070}$
AgeAge (Gyr) ${1.28}_{-0.91}^{+1.6}$ ${1.99}_{-0.40}^{+0.55}$ ${2.33}_{-0.56}^{+0.71}$ ${9.1}_{-3.9}^{+3.1}$ ${2.64}_{-0.39}^{+0.38}$ ${4.27}_{-0.42}^{+0.53}$
EEP h Equal evolutionary phase ${331}_{-37}^{+31}$ ${380}_{-17}^{+19}$ ${381}_{-21}^{+20}$ ${400}_{-37}^{+16}$ ${403.1}_{-11}^{+7.9}$ ${452.9}_{-5.2}^{+4.3}$
AV V-band extinction (mag) ${0.22}_{-0.14}^{+0.18}$ ${0.217}_{-0.095}^{+0.055}$ ${0.062}_{-0.041}^{+0.039}$ ${0.103}_{-0.053}^{+0.032}$ ${0.086}_{-0.053}^{+0.043}$
σSED SED photometry error scaling ${3.27}_{-0.73}^{+1.2}$ ${0.84}_{-0.22}^{+0.35}$ ${0.70}_{-0.17}^{+0.29}$ ${1.17}_{-0.27}^{+0.42}$ ${1.20}_{-0.27}^{+0.44}$
ϖ Parallax (mas)5.64 ± 0.102.925 ± 0.0336.536 ± 0.047 ${2.948}_{-0.075}^{+0.074}$ ${2.973}_{-0.076}^{+0.077}$
d Distance (pc) ${177.4}_{-3.1}^{+3.2}$ ${341.8}_{-3.8}^{+3.9}$ 153.0 ± 1.1 ${339.2}_{-8.3}^{+8.8}$ ${336.4}_{-8.5}^{+8.9}$
Planetary Parameters:    
P Period (days) ${3.4095675}_{-0.0000069}^{+0.0000070}$ 5.0037775 ± 0.00000484.720219 ± 0.00001110.180249 ± 0.0000155.331751 ± 0.0000115.331751 ± 0.000011
RP Radius ( RJ) ${1.060}_{-0.034}^{+0.041}$ ${1.771}_{-0.056}^{+0.060}$ ${1.396}_{-0.054}^{+0.056}$ ${1.060}_{-0.039}^{+0.040}$ ${1.239}_{-0.039}^{+0.046}$ ${1.241}_{-0.049}^{+0.052}$
MP Mass ( MJ) ${6.33}_{-0.31}^{+0.29}$ 0.88 ± 0.162.37 ± 0.24 ${0.851}_{-0.047}^{+0.052}$ 0.99 ± 0.11 ${0.912}_{-0.10}^{+0.095}$
T0 i Optimal conjunction time (BJDTDB)2458629.47972 ± 0.00039 ${2458459.73877}_{-0.00083}^{+0.00071}$ 2458913.37033 ± 0.00045 ${2458607.90338}_{-0.00049}^{+0.00052}$ ${2458990.55302}_{-0.00079}^{+0.00081}$ ${2458990.55306}_{-0.00082}^{+0.00081}$
a Semimajor axis (au) ${0.04860}_{-0.00094}^{+0.00080}$ ${0.06608}_{-0.0011}^{+0.00098}$ ${0.0626}_{-0.0012}^{+0.0011}$ ${0.0903}_{-0.0013}^{+0.0018}$ ${0.06864}_{-0.00075}^{+0.00079}$ ${0.06586}_{-0.00075}^{+0.00069}$
i Inclination (degrees) ${88.41}_{-0.93}^{+1.0}$ ${82.54}_{-0.59}^{+0.42}$ ${85.70}_{-0.65}^{+1.3}$ ${88.51}_{-0.22}^{+0.29}$ ${88.84}_{-1.1}^{+0.81}$ ${88.4}_{-1.3}^{+1.1}$
e Eccentricity ${0.072}_{-0.023}^{+0.021}$ ${0.050}_{-0.035}^{+0.054}$ ${0.073}_{-0.052}^{+0.092}$ ${0.024}_{-0.017}^{+0.032}$ ${0.036}_{-0.026}^{+0.044}$ ${0.057}_{-0.039}^{+0.050}$
τcirc Tidal circularization timescale (Gyr) ${3.32}_{-0.61}^{+0.53}$ ${0.206}_{-0.057}^{+0.060}$ ${1.29}_{-0.33}^{+0.37}$ ${42.5}_{-8.1}^{+11}$ 1.85 ± 0.38 ${1.52}_{-0.38}^{+0.42}$
ω* Argument of periastron (degrees) $-{74}_{-15}^{+13}$ ${159}_{-99}^{+86}$ ${105}_{-40}^{+64}$ $-{110}_{-130}^{+120}$ $-{180}_{-100}^{+110}$ ${120}_{-70}^{+57}$
Teq Equilibrium temperature (K) ${1586}_{-40}^{+52}$ ${1749}_{-30}^{+26}$ ${1679}_{-34}^{+35}$ 918 ± 11 ${1619}_{-23}^{+24}$ ${1642}_{-24}^{+25}$
K RV semiamplitude (m s−1) ${713}_{-22}^{+20}$ 78 ± 14 ${223}_{-21}^{+22}$ ${82.5}_{-3.8}^{+4.0}$ ${87.3}_{-9.9}^{+9.2}$ ${87.6}_{-9.6}^{+8.9}$
RP /R* Radius of planet in stellar radii ${0.08108}_{-0.00071}^{+0.00075}$ ${0.08738}_{-0.00086}^{+0.00091}$ ${0.0745}_{-0.0015}^{+0.0014}$ 0.1040 ± 0.0015 ${0.05827}_{-0.00070}^{+0.00071}$ ${0.05834}_{-0.00072}^{+0.00074}$
a/R* Semimajor axis in stellar radii ${7.78}_{-0.27}^{+0.22}$ ${6.82}_{-0.24}^{+0.22}$ ${6.98}_{-0.23}^{+0.24}$ ${18.54}_{-0.60}^{+0.70}$ ${6.76}_{-0.20}^{+0.17}$ 6.48 ± 0.23
DepthFlux decrement at midtransit0.00657 ± 0.00012 ${0.00764}_{-0.00015}^{+0.00016}$ 0.00555 ± 0.000210.01083 ± 0.00032 ${0.003395}_{-0.000081}^{+0.000083}$ ${0.003403}_{-0.000083}^{+0.000087}$
τ Ingress/egress transit duration (days) ${0.01247}_{-0.00061}^{+0.0012}$ ${0.0454}_{-0.0026}^{+0.0028}$ ${0.0177}_{-0.0033}^{+0.0027}$ 0.0208 ± − 0.0022 ${0.01483}_{-0.00034}^{+0.00084}$ ${0.01503}_{-0.00049}^{+0.0013}$
T14 Total transit duration (days) ${0.1576}_{-0.0013}^{+0.0015}$ 0.1502 ± 0.0017 ${0.1934}_{-0.0029}^{+0.0025}$ 0.1736 ± 0.00230.2627 ± 0.0020 ${0.2631}_{-0.0021}^{+0.0022}$
b Transit impact parameter ${0.23}_{-0.15}^{+0.13}$ ${0.8763}_{-0.0067}^{+0.0063}$ ${0.504}_{-0.19}^{+0.085}$ ${0.481}_{-0.076}^{+0.050}$ ${0.136}_{-0.095}^{+0.13}$ ${0.18}_{-0.12}^{+0.14}$
TS,14 Total eclipse duration (days) ${0.1387}_{-0.0055}^{+0.0063}$ ${0.1481}_{-0.019}^{+0.0070}$ ${0.208}_{-0.013}^{+0.041}$ ${0.1736}_{-0.0066}^{+0.0075}$ ${0.263}_{-0.013}^{+0.014}$ ${0.279}_{-0.017}^{+0.024}$
ρP Density (cgs) ${6.58}_{-0.75}^{+0.70}$ ${0.195}_{-0.040}^{+0.042}$ ${1.08}_{-0.15}^{+0.18}$ ${0.88}_{-0.11}^{+0.13}$ ${0.640}_{-0.093}^{+0.097}$ ${0.589}_{-0.091}^{+0.098}$
$\mathrm{log}{g}_{P}$ Surface gravity ${4.145}_{-0.038}^{+0.031}$ ${2.841}_{-0.095}^{+0.078}$ ${3.479}_{-0.056}^{0.054}$ ${3.273}_{-0.043}^{+0.045}$ ${3.201}_{-0.059}^{+0.052}$ ${3.165}_{-0.061}^{+0.055}$
TS Time of eclipse (BJDTDB) ${2458467.566}_{-0.038}^{+0.031}$ ${2454824.46}_{-0.20}^{+0.12}$ ${2458717.445}_{-0.13}^{+0.097}$ ${2455701.431}_{-0.12}^{+0.092}$ ${2454189.289}_{-0.17}^{+0.095}$ ${2454189.27}_{-0.19}^{+0.12}$
$e\cos {\omega }_{* }$   ${0.019}_{-0.018}^{+0.014}$ $-{0.012}_{-0.062}^{+0.038}$ $-{0.012}_{-0.042}^{+0.032}$ $-{0.002}_{-0.018}^{+0.014}$ $-{0.007}_{-0.049}^{+0.028}$ $-{0.013}_{-0.057}^{+0.035}$
$e\sin {\omega }_{* }$   $-{0.068}_{-0.022}^{+0.024}$ ${0.007}_{-0.029}^{+0.042}$ ${0.053}_{-0.056}^{+0.10}$ ${0.000}_{-0.029}^{+0.030}$ ${0.001}_{-0.027}^{+0.028}$ ${0.031}_{-0.033}^{+0.044}$
d/R* Separation at midtransit ${8.29}_{-0.40}^{+0.36}$ ${6.76}_{-0.51}^{+0.38}$ ${6.59}_{-0.77}^{+0.54}$ ${18.58}_{-0.88}^{+0.97}$ ${6.74}_{-0.36}^{+0.32}$ ${6.27}_{-0.47}^{+0.41}$

Notes. See Table 3 in Eastman et al. (2019) for a detailed description of all derived and fitted parameters.

a The SED was not included within the global fit for TOI-1333. b No TESS dilution prior was used for TOI-1478 b because initial fits showed a fitted dilution past the 10% prior we used in the other systems. We fit for a dilution term within the fit for the TESS bandpass but with no prior. c The tidal quality factor (QS ) is assumed to be 106. d TC prior comes from analysis of the WASP photometry (see Section 2.2). We note that this time is in HJDTDB while all data files and results here are BJDTDB. The difference between these two time systems is on the order of seconds while the precision on TC used as a prior is on order of minutes, and therefore has no influence on the results. e We assume the TESS correction for blending is much better than 10%. We use a prior of 10% of the determined blending from TICv8 (Stassun et al. 2018). f Dilution prior for TOI-1333 comes from our three-component SED analysis (see Section 2.6.4). g The initial metallicity is the metallicity of the star when it was formed. h The equal evolutionary point (EEP) corresponds to static points in a star's evolutionary history when using the MIST isochrones and can be a proxy for age. See Section 2 in Dotter (2016) for a more detailed description of EEP. i Optimal time of conjunction minimizes the covariance between TC and period. This is the transit midpoint.

Download table as:  ASCIITypeset images: 1 2

2.5.4. FEROS Spectroscopy

Using the FEROS spectrograph (Kaufer et al. 1999) mounted on the 2.2 m MPG telescope at La Silla observatory in Chile, we obtained eight observations of TOI-1478. FEROS has R = 48,000, and the observations were between UT 2020 March 2 and UT 2020 March 22. A ThAr+Ne lamp was used to illuminate the fiber simultaneously with the observations to determine the instrumental offset. We reduced the spectra, derived the RVs, and produced the bisector spans using the CERES suite for echelle pipelines (Brahm et al. 2017).

2.5.5. CORALIE Spectroscopy

TOI-1478 was observed with the CORALIE high-resolution spectrograph (R = 60,000) on the Swiss 1.2 m Euler telescope at La Silla Observatories, Chile (Queloz et al. 2001). A total of 14 spectra were obtained between UT 2020 January 26 and March 16, covering several orbits of TOI-1478 b. CORALIE is fed by a 2'' science fiber and a secondary fiber with simultaneous Fabry–Perot for wavelength calibration. RVs were computed with the standard CORALIE data-reduction pipeline via cross-correlation with a binary G2 mask. Activity indices, bisector spans (BIS), and other line profile diagnostics were extracted as well. We find no correlation between the RVs and BIS, nor activity indicators. Our exposure times varied between 1200 and 1800 s depending on the site conditions and observing schedule.

2.5.6. MINERVA-Australis Spectroscopy

Minerva-Australis is an array of four PlaneWave CDK700 telescopes located in Queensland, Australia, fully dedicated to the precise RV follow-up of TESS candidates. The four telescopes can be simultaneously fiber-fed to a single KiwiSpec R4-100 high-resolution (R = 80,000) spectrograph (Barnes et al. 2012; Addison et al. 2019, 2021). TOI-628 was monitored by Minerva-Australis using two or three telescopes in the array (Minerva3, Minerva4, Minerva6) between UT 2019 September 15 and December 3. Each epoch consists of two 30 minute exposures, and the resulting RVs are binned to a single point, as shown by the example in Table 3. Telescopes 1 and 3 obtained seven RV epochs, while Telescope 4 obtained four epochs. RVs for the observations are derived for each telescope by cross-correlation, where the template being matched is the mean spectrum of each telescope. The instrumental variations are corrected by using the RVs computed from different Minerva telescopes as originating from independent instruments within our global model.

2.6. High-resolution Imaging

As part of our standard process for validating transiting exoplanets to assess possible contamination of bound or unbound companions on the derived planetary radii (Ciardi et al. 2015), we obtained high-spatial-resolution imaging observations of all five systems.

2.6.1. Speckle Imaging

We searched for close companions to TOI-628, TOI-640, and TOI-1478 with speckle imaging in the I band on the 4.1 m Southern Astrophysical Research (SOAR) telescope (Tokovinin 2018). The speckle imaging was conducted using HRCam (field of view of 15'') and had a 001575 pixel scale. TOI-628 was observed on UT 2019 November 11 with a sensitivity of ΔMag = 7.2 at 1''. Speckle observations of TOI-640 were taken on UT 2019 May 18 and had an estimated contrast of Δmag = 6.6 at 1''. Observations of TOI-1478 were taken on UT 2020 January 7 and had an estimated contrast of Δmag = 6.8 at 1''. See Ziegler et al. (2020) for a description of the general observing strategy for TESS targets. No nearby companion was observed for any of the three targets out to 3''.

We also obtained two sets of high-resolution speckle images of TOI-1478. One was collected on UT 2020 January 14 using the Zorro instrument mounted on the 8 m Gemini South telescope located on the summit of Cerro Pachon in Chile and the other on UT 2020 February 18 using the Alopeke instrument mounted on the 8 m Gemini-North telescope located on the summit of Maunakea in Hawaii. These twin instruments simultaneously observe in two bands, $\tfrac{\lambda }{{\rm{\Delta }}\lambda }$ = 832/40 nm and 562/54 nm, obtaining diffraction limited images with inner working angles 0026 and 0017, respectively. Each observation consisted of 6 minutes of total integration time at each telescope taken as sets of 1000 × 0.06 second images. All of the images were combined and subjected to Fourier analysis, leading to the production of final data products including speckle reconstructed imagery (see Howell et al. 2011). Both speckle imaging results showed similar contrast limits and revealed that TOI-1478 is a single star to contrast limits of 5–9 mag (out to 117), ruling out most main-sequence companions to TOI-1478 within the spatial limits of ∼4–180 au (for d = 153 pc, as determined from the Gaia DR2 parallax; Gaia Collaboration et al. 2018).

2.6.2. Adaptive Optics Imaging

We observed TOI-628 on UT 2019 November 11 using the ShARCS adaptive optics system on the 3 m Shane Telescope at Lick Observatory. ShARCS has a field of view of 20'' × 20'' and a pixel scale of 0033 pixel−1. We conducted our observations using a square four-point dither pattern with a separation of 4'' between each dither position. Our observations were taken in natural guide-star mode with high winds. We obtained one sequence of observations in the Ks band and a second sequence in the J band for a total integration time of 510 s in the Ks band and 225 s in the J band. See Savel et al. (2020) for a detailed description of the observing strategy. Neither set of observations revealed any companions for TOI-628.

We observed TOI-1333 (Br–γ and H − cont) and TOI-1478 with infrared high-resolution adaptive optics (AO) imaging at Palomar Observatory. The Palomar Observatory observations were made with the PHARO instrument (Hayward et al. 2001) behind the natural guide-star AO system P3 K (Dekany et al. 2013). The observations were made on 2019 November 10 UT in a standard five-point quincunx dither pattern with steps of 5''. Each dither position was observed three times, offset in position from each other by 05 for a total of 15 frames. The camera was in the narrow-angle mode with a full field of view of ∼25'' and a pixel scale of approximately 0025 per pixel. Observations were made in the narrow-band Br–γ filter (λo = 2.1686; Δλ = 0.0326 μm) for TOI-1333 and TOI-1478, and in the H − cont filter (λo = 1.668; Δλ = 0.018 μm) for TOI-1333. The observations get down to ΔMag = 6.54 (Br–γ) and 7.52 (H − cont) for TOI-1333 and ΔMag = 6.8 (Br–γ) for TOI-1478 (all at ∼05).

TOI-1333 was also observed using NIRI on Gemini-North (Hodapp et al. 2003) on UT 2019 November 14 in the Br–γ filter. NIRI has a 22'' × 22'' field of view with a 0022 pixel scale. Our sequence consisted of nine images, each with exposure time 4.4 s, and we dithered the telescope between each exposure. A sky background was constructed from the dithered frames and subtracted from each science image. We also performed bad pixel removal and flat-fielding, and then aligned and coadded frames. NIRI got down to ΔMag = 6.7 at 0472.

We also observed TOI-1601 using the Near Infrared Camera 2 (NIRC2) AO setup on the W. M. Keck Observatory in the Br–γ filter and in the J − cont filter on UT 2020 September 9. The NIRC2 detector has a 9.971 mas pixel−1 using a 1024 × 1024 CCD (field of view = 10'' × 10''; Service et al. 2016). Unfortunately, the lower left quadrant of the CCD is known to have higher than typical noise levels in comparison to the others. To avoid this issue, a three-point dither pattern technique was used. The images were aligned and stacked after normal flat-field and sky background corrections. No nearby companions were seen down to Δmag = 6.680 (J − cont) and 6.402 (Br–γ) for TOI-1601 at 05.

While the observing strategy differed, all of the AO data were processed and analyzed with a custom set of IDL tools. The science frames were flat-fielded and sky-subtracted. The flat fields were generated from a median combination of the dark-subtracted flats taken on sky. The flats were normalized such that the median value of the flats is unity. The sky frames were generated from the median average of the 15 dithered science frames; each science image was then sky-subtracted and flat-fielded. The reduced science frames were combined into a single combined image using an intrapixel interpolation that conserves flux, shifts the individual dithered frames by the appropriate fractional pixels, and median-coadds the frames (see Figure 4). The final resolution of the combined dither was determined from the FWHM of the point-spread function; the resolutions of the Br–γ and H − cont images are 0092 and 0075, respectively (Figure 4).

Figure 4.

Figure 4. Palomar PHARO (left) H-band and (middle) Brγ-band 4σ contrast curve for TOI-1333 with the AO image embedded in the plot. The (right) Gemini NIRI Brγ-band AO 5σ contrast curve for TOI-1333. The NIRI AO image is embedded in the plot. The second star in the image is TIC 2010985858, and we properly account for its blending in our fit (see Section 3). The colored swath represents the uncertainty on the 5σ contrast curve (see Section 2.6.2).

Standard image High-resolution image

The sensitivities of the final combined AO images were determined by injecting simulated sources azimuthally around the primary target every 20° at separations of integer multiples of the central source's FWHM (Furlan et al. 2017; M. Lund et al. 2021, in preparation). The brightness of each injected source was scaled until standard aperture photometry detected it with 5σ significance. The resulting brightness of the injected sources relative to the target set the contrast limits at that injection location. The final 5σ limit at each separation was determined from the average of all of the determined limits at that separation, and the uncertainty on the limit was set by the rms dispersion of the azimuthal slices at a given radial distance. The sensitivity curves for TOI-1333 are shown in Figure 4 along with an inset image zoomed in to the primary target showing no other companion stars.

2.6.3. TOI-1333 Companions

In the case of TOI-1333, two additional sources were detected in the PHARO and GEMINI AO imaging (Figure 4; only the 3'' companion is shown). The first source is 743 (Gaia Collaboration et al. 2018) to the east of the primary target and is separately resolved by the Two Micron All Sky Survey (2MASS J21400422+4824221; TIC 395171213). The second source is only 281 (Gaia Collaboration et al. 2018) away and was not separately detected by 2MASS, although it was detected by Gaia and hence is in the TESS Input Catalog (TIC 2010985858). Based on the differential magnitudes measured by Palomar, the deblended infrared magnitudes for the primary star and nearby companion are K1 = 8.355 ± 0.024 mag, H1 = 8.514 ± 0.043 mag (H1K1 = 0.159 ± 0.049) and K2 = 10.614 ± 0.026 mag, H2 = 10.871 ± 0.044 mag (H2K2 = 0.257 ± 0.051), respectively. The 3'' star has a Gaia magnitude of G = 12.6221 ± 0.0016.

The primary target has a Gaia distance of 200.5 ± 1.2 pc, whereas the 7'' companion has a Gaia distance of 1030 ± 35 pc, indicating that this companion is not bound to the primary and is simply a chance alignment near the line of sight to TOI-1333. However, the 3'' companion has a Gaia distance of 196.8 ± 1.6 pc and proper motions that are nearly identical to that of the primary star (μ1α = − 9.81 ± 0.05 versus μ2α = − 9.06 ± 0.08 and μ1δ = − 10.50 ± 0.04 versus μ2δ = − 9.24 ± 0.12 mas/yr); it is, therefore, highly probable that TIC 2010985858 is physically bound to the primary star with a projected separation of ∼590 au. We account for the blended light from these two companions in our global analysis (see Section 3 for details on how).

We use the LOFTI software package (Pearce et al. 2020) to derive orbital parameters of the visual binary system formed by TOI-1333 and its companion. LOFTI uses the relative proper motions of the two stars from the Gaia catalog to sample probable orbits for a binary star system. To derive the mass of the companion—required to fit the orbit using LOFTI—we use the isochrones package (Morton 2015). We perform a spectral energy distribution (SED) fit on photometry from TOI 1333's binary companion using the Gaia G, BP, and RP magnitudes, along with its Gaia parallax. Using the MESA Isochrones and Stellar Tracks (MIST) isochrone (Dotter 2016) as the base isochrone, we derive a mass for the companion of ${0.808}_{-0.042}^{+0.043}$ M. We use the astrometric parameters for the two systems from Gaia EDR3.

The LOFTI fit reveals that the semimajor axis of the binary orbit is ${570}_{-170}^{+590}$ au and the orbital inclination is ${125}_{-10}^{+18}$ degrees, ruling out an edge-on orbit for the binary at high confidence. The orbital eccentricity is weakly constrained to be less than 0.69 with 95% confidence (with a slight preference for values between 0.5 and 0.7, but consistent with zero). An independent analysis of TOI-1333 and its companion were presented in Mugrauer & Michel (2020).

2.6.4. TOI-1333 Spectral Energy Distribution

The presence of the two stellar companions within a few arcseconds of TOI-1333 implies that the TESS and SG1 light curves of the TOI-1333 planet transit are likely to be diluted to some extent by the light from these other stars. Although the QLP corrects the TESS light curve for the blended contributions of known targets in the TESS input catalog (TIC), we need to correct the follow-up photometry from TFOP for different amounts of dilution. To quantify this flux dilution, we performed a multicomponent SED fit with Kurucz model atmospheres following the procedures described in Stassun & Torres (2016), utilizing the resolved broadband measurements from Gaia, 2MASS, and our AO observations (see Section 2.6.2).

We adopted the spectroscopic Teff (6250 ± 100 K) from TRES for TOI-1333 and the Teff from the TICv8 (Stassun et al. 2019) and from the Gaia DR2 catalog for the companions, with AV being left as a free parameter but limited to the maximum line-of-sight value from the dust maps of Schlegel et al. (1998). The resulting fits are shown in Figure 5, with best-fit AV = 0.06 ± 0.06 for the TOI-1333 planet host. Integrating the SED models over the TESS bandpass gives the total flux dilution (F2 + F3)/F1 = 0.55 with values of 0.56 in the I band, 0.54 in the R band, and 0.58 in the Sloan ${z}^{{\prime} }$. We also used the SED fits to constrain the contribution from only the 2'' companion since some of our follow-up photometry only resolved the 7'' companion. The flux dilution of F2/F1 is 0.07 in the I band, 0.06 in the R band, and 0.07 in the Sloan ${z}^{{\prime} }$. By combining the deblended SED of the primary star TOI-1333 with the known Gaia DR2 parallax, we measure its radius to be R = 1.963 ± 0.064 R. We use this as a prior on the global fit for TOI-1333 (see Section 3).

Figure 5.

Figure 5. Three-component SED fit for TOI-1333. The blue points are the predicted integrated fluxes for the primary star. The red and black points are the observed values at the corresponding passbands for each star. The widths of the passbands are the horizontal error bars, and the vertical errors represent the 1σ uncertainties. The final model fit is shown by the solid line for TOI-1333 (black) and the 7'' (cyan) and bound 3'' (red) companions.

Standard image High-resolution image

2.7. Location in the Galaxy, UVW Space Motion, and Galactic Population

For each of the TOIs analyzed here, we used their parallaxes, proper motions, and radial velocities and associated uncertainties from the Gaia DR2 catalog (Gaia Collaboration et al. 2018) to determine their location, kinematics, orbits, and associations with known stellar populations. 78 We corrected the native DR2 parallaxes and uncertainties using the prescription given in Lindegren et al. (2018). From these we computed the prior (DR2-based) estimates of the distances to the systems. 79 We used these to compute the heliocentric UVW space motions of the host stars, and then corrected for the Sun's motion (UVW) with respect to the local standard of rest (LSR) as determined by Coşkunoǧlu et al. (2011). These resulting (UVW) values are shown in Table 1. We note that we adopt a right-handed coordinate system, such that a positive U is toward the Galactic center.

We used the Galactic latitudes and distances to the systems to estimate their Z height relative to the Sun, and then corrected for the Z ≃ 30 pc offset of the Sun from the Galactic plane as determined by Bovy (2017) from the analysis of local giants. We use the UVW velocities relative to the LSR to determine the likelihood that the star belongs to a thin disk, thick disk, halo, or Hercules stream, 80 using the categorization scheme of Bensby et al. (2014). We also report the estimates of the parameters of the Galactic orbits of the systems as determined by Mackereth & Bovy (2018) using Gaia DR2 astrometry and radial velocities. 81 We estimated the spectral type of each TOI using its Teff as determined from the global fit and given in Table 4 and using the Teff–spectral type relations of Pecaut & Mamajek (2013). We then compared the position and orbits of the systems to the scale height hZ of stars of similar spectral type as determined by Bovy (2017).

Finally, we also consider whether the systems may belong to any of the known nearby (≲150 pc), young (≲Gyr) associations using the BANYAN Σ (Bayesian Analysis for Nearby Young AssociatioNs Σ) estimator (Gagné et al. 2018). Not surprisingly, none of the systems had any significant (≳1%) probability of being associated with these young associations, and the BANYAN Σ estimator assigned all five systems as belonging to the "field" with a high probability ≳99%. We now discuss the results for each of the systems individually.

TOI-628: We find a distance from the Sun of d = 178 ± 3 pc, consistent with the posterior value listed in Table 4, and ZZ ≃ − 23 pc. We derive velocities relative to the LSR of (U, V, W) = (−8.7 ± 0.4, 3.2 ± 0.2, 4.2 ± 0.1) km s−1. According to the categorization scheme of Bensby et al. (2014), the system has a >99% probability of belonging to the thin disk. The Galactic orbit as estimated by Mackereth & Bovy (2018) has a perigalacticon of Rp = 7.67 kpc and apogalacticon of Ra = 8.19 kpc, an eccentricity of e = 0.03, and a maximum Z excursion from the Galactic plane of ${Z}_{\max }=63\,\mathrm{pc}$. This orbit is consistent with both the current location of the system and the scale height of 97 pc for stars of similar spectral type (F7V). Indeed, TOI-628 is relatively dynamically "cold" for its spectral type. In other words, it has an orbit that is fairly close to that of the local LSR.

TOI-640: We find a distance from the Sun of d = 340 ± 4 pc, consistent with the posterior value listed in Table 4, and ZZ ≃ − 76 pc. We derive velocities relative to the LSR of (U, V, W) = (−16.8 ± 0.2, −16.7 ± 0.4, −8.7 ± 0.2) km s−1. According to the categorization scheme of Bensby et al. (2014), the system has an ∼99% probability of belonging to the thin disk. The Galactic orbit as estimated by Mackereth & Bovy (2018) has a perigalacticon of Rp = 6.28 kpc and apogalacticon of Ra = 8.16 kpc, an eccentricity of e = 0.13, and a maximum Z excursion from the Galactic plane of ${Z}_{\max }=150\,\mathrm{pc}$. This orbit is both consistent with the current location of the system and suggests that the system is nearing its maximum excursion above the plane. It is also consistent with the scale height of 85 pc for stars of similar spectral type (F5.5V).

TOI-1333: We find a distance from the Sun of d = 200 ± 2 pc and ZZ ≃ 19 pc. We derive velocities relative to the LSR of (U, V, W) = (23.0 ± 0.1, −1.00 ± 0.3, −6.0 ± 0.1) km s−1. According to the categorization scheme of Bensby et al. (2014), the system has an ∼99% probability of belonging to the thin disk. The Galactic orbit as estimated by Mackereth & Bovy (2018) has a perigalacticon of Rp = 7.25 kpc and apogalacticon of Ra = 8.32 kpc, an eccentricity of e = 0.07, and a maximum Z excursion from the Galactic plane of ${Z}_{\max }=91\,\mathrm{pc}$. This orbit is consistent with the current location of the system. It is also consistent with the scale height of 97 pc for stars of similar spectral type (F7V).

TOI-1478: We find a distance from the Sun of d = 153 ± 1 pc, and ZZ ≃ 67 pc. We derive velocities relative to the LSR of (U, V, W) = (−37.0 ± 0.3, 26.4 ± 0.4, 32.5 ± 0.2) km s−1. According to the categorization scheme of Bensby et al. (2014), the system has an ∼88% probability of belonging to the thin disk and an ∼12% probability of belonging to the thick disk (and negligible probabilities of belonging to the halo or Hercules stream). The Galactic orbit as estimated by Mackereth & Bovy (2018) has a perigalacticon of Rp = 7.71 kpc and apogalacticon of Ra = 10.34 kpc, an eccentricity of e = 0.14, and a maximum Z excursion from the Galactic plane of ${Z}_{\max }=650\,\mathrm{pc}$. Unfortunately, Bovy (2017) was unable to determine the scale height of stars of similar spectral type (G6V), due to incompleteness. Nevertheless, it would appear that TOI-1478's orbit has a maximum Z excursion that exceeds the expected scale height for stars of similar spectral type as estimated by extrapolating from the results of Bovy (2017) from earlier spectral types. Surprisingly, its current distance above the plane is only a small fraction of its predicted maximum excursion. In summary, the weight of evidence suggests that TOI-1478 may well be a thick disk star that we happen to be observing when it is near the Galactic plane. Detailed chemical abundance measurements (e.g., [α/Fe]) may provide corroborating evidence for or against this hypothesis.

TOI-1601: We find a distance from the Sun of d = 336 ± 9 pc, consistent with the posterior value listed in Table 4, and ZZ ≃ − 73 pc. We derive velocities relative to the LSR of (U, V, W) = (−8.1 ± 0.7, −14.5 ± 0.7, 20.9 ± 0.4) km s−1. According to the categorization scheme of Bensby et al. (2014), the system has an ∼98% probability of belonging to the thin disk. The Galactic orbit as estimated by Mackereth & Bovy (2018) has a perigalacticon of Rp = 6.50 kpc and apogalacticon of Ra = 8.32 kpc, an eccentricity of e = 0.12, and a maximum Z excursion from the Galactic plane of ${Z}_{\max }=351\,\mathrm{pc}$. This orbit is consistent with the current location of the system. The maximum Z excursion is a factor of ∼3.3 times larger than the scale height of 103 pc for stars of similar spectral type (G0V). The probability that a star in a population with a given scale height hz = 108 pc has a maximum excursion of ${z}_{\max }=351$ pc is nonnegligible. Thus, we expect that TOI-1601 is a thin disk star that is simply in the tail of the distribution of ${z}_{\max }$. Again, detailed abundances could corroborate or refute this conclusion.

3. EXOFASTv2 Global Fits

We use the publicly available exoplanet fitting suite, EXOFASTv2 (Eastman et al. 2013, 2019), to globally fit the available photometry and RVs to determine the host star and planetary parameters for TOI-628 b, TOI-640 b, TOI-1333 b, TOI-1478 b, and TOI-1601 b. We fit the TESS and SG1 transits (see Section 2.4), accounting for the 30 minute smearing from the FFIs. Within the fit, the SG1 light curves were detrended (additive) against the corresponding parameters shown in Table 1. See Section D in the appendix of Collins et al. (2017) for a description of each detrending parameter. We use the MIST stellar evolution models (Paxton et al. 2011, 2013, 2015; Choi et al. 2016; Dotter 2016) and the SED within the fit to determine the host star parameters for all systems but TOI-1333 b. The SED fit within the global fit puts systematic floors on the broadband photometry errors (see Table 1; Stassun & Torres 2016). We also note that EXOFASTv2 defaults to a lower limit on the systematic error on the bolometric flux (Fbol ∼ 3%) given the spread seen from various techniques to calculate it (Zinn et al. 2019). We place a Gaussian prior on the metallicity from our analysis of the host star's spectra from TRES, or CHIRON in the case of TOI-640 (see Sections 2.5.3 and 2.5.2). We also place a Gaussian prior on the parallax from Gaia (Gaia Collaboration et al. 2016, 2018), correcting for the 30 μas offset reported by Lindegren et al. (2018), and an upper limit on the line-of-sight extinction from Schlegel et al. (1998) and Schlafly & Finkbeiner (2011). We also fit for a dilution term on the TESS band. Since the QLP corrects the TESS light curves for all known blended companions, we place a Gaussian prior of 0% ± 10% of the contamination ratio reported by the TIC (Stassun et al. 2018). We assume that the light curve has been corrected to a precision better than 10% (and test this with preliminary EXOFASTv2, showing the dilution to be consistent with zero), but this flexibility also provides an independent check on the correction applied and allows us to propagate the uncertainty in the correction. We do not find any significant additional dilution in any of the systems (within this prior and consistent with zero dilution) other than TOI-1478 b, where our fit suggests an additional 13% dilution (${0.126}_{-0.021}^{+0.020}$) is needed for the TESS light curve to be consistent with the TFOP photometry. For this fit, we remove this prior, essentially allowing the TFOP observations to constrain the depth of the transit. The cause of this additional dilution is not clear since we see no evidence for any unknown companions in our high-spatial-resolution imaging. We note that TESS only observed three transits in one sector for TOI-1478 b, the longest-period planet in our sample, and the TFOP light curves were both at higher spatial resolution. We also ran a fit where we allowed for a slope in the RVs, but we found no significant trends for any system (we do not fit for a slope in the final fits). A list of the priors for each target is shown in Table 4. Table 3 of Eastman et al. (2019) shows a list and description of the fitted and derived parameters, including the bounds that EXOFASTv2 adopts for each fitted parameter. We note that eccentricity, a key parameter for this study, is bound as such, $0\leqslant e\leqslant 1-\tfrac{a+{R}_{p}}{{R}_{* }}$, in order to ensure that the periastron values of the planet orbits are larger than the sum of the stellar and planetary radius. We deem a fit to be fully converged by following the recommended statistical threshold of a Gelman–Rubin statistic (<1.01) and the independent draw (>1000) that is recommended by Eastman et al. (2019). The results from our EXOFASTv2 fits are shown in Tables 4 and 5, and the models are shown for the transits and RVs in Figures 2 and 3.

Table 5. Median Values and 68% Confidence Intervals for the Global Models

TOI-628  
Wavelength Parameters: B r' z'TESS
u1 Linear limb-darkening coeff ${0.646}_{-0.061}^{+0.058}$ 0.329 ± 0.051 ${0.213}_{-0.052}^{+0.053}$ ${0.236}_{-0.046}^{+0.045}$
u2 Quadratic limb-darkening coeff0.286 ± 0.0570.306 ± 0.0490.313 ± 0.050 ${0.292}_{-0.048}^{+0.047}$
AD Dilution from neighboring stars0.0012 ± 0.0036
Telescope Parameters:MINERVAF3MINERVAF4MINERVAF6TRES
γrel Relative RV offset (m s−1) ${19990}_{-28}^{+26}$ 19969 ± 35 ${20032}_{-25}^{+26}$ $-{633}_{-18}^{+20}$
σJ RV Jitter (m s−1) ${60}_{-18}^{+29}$ ${102}_{-24}^{+36}$ ${83}_{-20}^{+26}$ ${54}_{-20}^{+28}$
${\sigma }_{J}^{2}$ RV Jitter variance ${3700}_{-1800}^{+4400}$ ${10400}_{-4300}^{+8600}$ ${7000}_{-2900}^{+4900}$ ${2900}_{-1700}^{+3800}$
Transit Parameters:TESS KeplerCam UT 2019-12-06 (z')LCO SSO UT 2019-12-16 (B)Whitin UT 2020-02-08 (r')
σ2 Added variance ${0.000000119}_{-0.000000020}^{+0.000000024}$ ${0.0000366}_{-0.0000018}^{+0.0000020}$ ${0.00001138}_{-0.00000091}^{+0.0000010}$ ${0.00000765}_{-0.00000044}^{+0.00000047}$
F0 Baseline flux ${0.999998}_{-0.000038}^{+0.000039}$ 0.99958 ± 0.000231.00071 ± 0.000211.00125 ± 0.00012
C0 Additive detrending coeff ${0.00806}_{-0.00057}^{+0.00058}$ −0.00277 ± 0.00031
TOI-640  
Wavelength Parameters: I z'TESS V
u1 Linear limb-darkening coeff0.191 ± 0.0510.170 ± 0.0360.216 ± 0.0460.376 ± 0.050
u2 Quadratic limb-darkening coeff0.311 ± 0.0490.313 ± 0.0350.314 ± 0.0460.329 ± 0.049
AD Dilution from neighboring stars0.0000 ± 0.0022
Telescope Parameters:CHIRON
γrel Relative RV offset (m s−1) ${39091}_{-11}^{+10.}$
σJ RV Jitter (m s−1) ${32.3}_{-8.3}^{+12}$
${\sigma }_{J}^{2}$ RV Jitter variance ${1040}_{-470}^{+900}$
Transit Parameters:TESS Brierfield UT 2019-12-27 (I)PEST UT 2020-03-01 (V)LCO SSO UT 2020-08-23 (z')LCO SSO UT 2020-11-06 (z')
σ2 Added variance $-{0.0000000009}_{-0.0000000081}^{+0.0000000093}$ ${0.00000259}_{-0.00000051}^{+0.00000059}$ ${0.00000220}_{-0.00000045}^{+0.00000052}$ ${0.00000187}_{-0.00000041}^{+0.00000053}$ ${0.000000056}_{-0.000000080}^{+0.000000098}$
F0 Baseline flux1.000016 ± 0.0000220.99995 ± 0.000180.99984 ± 0.000190.99956 ± 0.000231.00034 ± 0.00012
C0 Additive detrending coeff−0.0183 ± 0.0033 ${0.00080}_{-0.00049}^{+0.00050}$ −0.00052 ± 0.00025
C1 Additive detrending coeff−0.0233 ± 0.0033
TOI-1333  
Wavelength Parameters: I R z'TESS
u1 Linear limb-darkening coeff0.228 ± 0.0490.258 ± 0.0460.183 ± 0.035 ${0.226}_{-0.044}^{+0.043}$
u2 Quadratic limb-darkening coeff0.323 ± 0.0490.297 ± 0.0480.303 ± 0.0350.306 ± 0.047
AD Dilution from neighboring stars0.345 ± 0.0160.344 ± 0.0170.0671 ± 0.0033−0.016 ± 0.028
Telescope Parameters:TRES
γrel Relative RV offset (m s−1) ${122}_{-14}^{+15}$
σJ RV Jitter (m s−1) ${41}_{-18}^{+21}$
${\sigma }_{J}^{2}$ RV Jitter variance ${1700}_{-1200}^{+2100}$
Transit Parameters:TESS CRCAO UT 2020-07-29 (I)LCO McD UT 2020-07-29 (z')LCO McD UT 2020-08-12 (z')CRCAO UT 2020-09-19 (R)
σ2 Added variance ${0.000000175}_{-0.000000023}^{+0.000000026}$ ${0.00000078}_{-0.00000012}^{+0.00000014}$ ${0.00000732}_{-0.00000097}^{+0.0000011}$ ${0.00000353}_{-0.00000046}^{+0.00000052}$ ${0.000000883}_{-0.000000084}^{+0.000000092}$
F0 Baseline flux0.999984 ± 0.0000360.99985 ± 0.00012 ${0.99927}_{-0.00025}^{+0.00024}$ 1.00245 ± 0.000181.00015 ± 0.00012
C0 Additive detrending coeff0.00033 ± 0.00023−0.00060 ± 0.000650.01022 ± , 0.00068 $-{0.00017}_{-0.00022}^{+0.00021}$
TOI-1478  
Wavelength Parameters: R i'TESS
u1 Linear limb-darkening coeff0.408 ± 0.0480.349 ± 0.0440.410 ± 0.045
u2 Quadratic limb-darkening coeff0.258 ± 0.0490.264 ± 0.047 ${0.320}_{-0.047}^{+0.048}$
AD Dilution from neighboring stars ${0.126}_{-0.021}^{+0.020}$
Telescope Parameters:CHIRON1CHIRON2CORALIEFEROSTRES
γrel Relative RV offset (m s−1)19384.9 ± 7.019419 ± 11 ${20833.5}_{-5.4}^{+5.2}$ 20821.5 ± 4.5−30 ± 12
σJ RV Jitter (m s−1) ${32.3}_{-5.5}^{+6.6}$ ${1.5}_{-1.5}^{+37}$ ${0.00}_{-0.00}^{+13}$ ${10.4}_{-4.5}^{+4.2}$ ${20.1}_{-3.3}^{+1.7}$
${\sigma }_{J}^{2}$ RV Jitter variance ${1040}_{-320}^{+470}$ ${0}_{-300}^{+1500}$ $-{7}_{-91}^{+180}$ ${108}_{-74}^{+110}$ ${405}_{-120}^{+69}$
Transit Parameters:TESS KeplerCam UT 2019-12-14 (i')PEST UT 2020-01-03 (R)
σ2 Added variance ${0.000000083}_{-0.000000039}^{+0.000000049}$ ${0.00000537}_{-0.00000036}^{+0.00000039}$ ${0.0000136}_{-0.0000011}^{+0.0000012}$
F0 Baseline flux1.000054 ± 0.0000611.00003 ± 0.000161.00026 ± 0.00022
C0 Additive detrending coeff0.00146 ± 0.00035
TOI-1601  
Wavelength Parameters: I R z'TESS
u1 Linear limb-darkening coeff0.281 ± 0.0500.285 ± 0.0500.215 ± 0.0470.305 ± 0.045
u2 Quadratic limb-darkening coeff ${0.305}_{-0.049}^{+0.050}$ ${0.260}_{-0.050}^{+0.049}$ ${0.281}_{-0.048}^{+0.049}$ 0.297 ± 0.048
AD Dilution from neighboring stars−0.0005 ± 0.0047
Telescope Parameters:TRES
γrel Relative RV offset (m s−1) $-{100.8}_{-7.3}^{+7.0}$
σJ RV Jitter (m s−1) ${0.00}_{-0.00}^{+21}$
${\sigma }_{J}^{2}$ RV Jitter variance $-{110}_{-300}^{+540}$
Transit Parameters:TESS GMU UT 2020-08-30 (R)GMU UT 2020-09-15 (R)CRCAO UT 2020-10-01 (R)Adams UT 2020-10-17 (I)LCO McD UT 2020-10-17 (z')
σ2 Added variance $-{0.000000044}_{-0.000000020}^{+0.000000024}$ ${0.00001324}_{-0.00000097}^{+0.0000010}$ ${0.0000105}_{-0.0000017}^{+0.0000020}$ ${0.00000134}_{-0.00000021}^{+0.00000025}$ ${0.00000066}_{-0.00000015}^{+0.00000017}$ ${0.00000069}_{-0.00000014}^{+0.00000016}$
F0 Baseline flux0.999969 ± 0.0000420.99318 ± 0.000181.00037 ± 0.000321.00043 ± 0.000121.00036 ± 0.00011 ${0.998577}_{-0.000092}^{+0.000091}$
C0 Additive detrending coeff−0.01148 ± 0.00042−0.015 ± 0.010−0.00039 ± 0.00033−0.00232 ± 0.00034−0.00039 ± 0.00024
C1 Additive detrending coeff0.00226 ± 0.000670.018 ± 0.010−0.00638 ± 0.00063−0.0015 ± 0.0013
C2 Additive detrending coeff−0.00160 ± 0.00053

Download table as:  ASCIITypeset images: 1 2 3

In the case of TOI-1333 b, we deviate slightly from the methodology in the previous paragraph because there are two nearby bright companions, both detected by high-resolution imaging (see Section 4). The 2'' nearby companion and 7'' star were blended in the TESS and CRCAO photometry (see Section 2.4), but only the 2'' companion was blended in the LCO observations. While the TESS light curve has already been deblended as part of the reduction pipeline (see Section 2.1), the SG1 observations were not. Our three-component SED analysis (see Section 2.6.4) determined that the nearby companion 2'' from TOI-1333 accounts for 6.6% in the ${z}^{{\prime} }$band, where LCO did not resolve the close companion. The combined flux contribution from both stars is 36.1% in the I band and 35.0% in the R band, where both companions were unresolved by CRCAO. We use these values with a 5% Gaussian prior EXOFASTv2 global fit (also placing the prior on the TESS dilution as discussed in the previous paragraph). We place a Gaussian prior on the host star's radius from the SED analysis of R = 1.963 ± 0.064 R. Preliminary SED fits of TOI-1333 using EXOFASTv2 and independent $\mathrm{log}g$ constraints from the SPC analysis of the TRES spectra suggested that TOI-1333 is a slightly evolved star. Given that the SED would normally constrain the Teff of the host star within the fit but was excluded for TOI-1333, we also place a prior on the Teff of 6250 ± 100 K from the SPC analysis of the TRES spectroscopy.

For TOI-1478 global fit, five RVs were acquired in the summer and winter of 2020, after a multimonth shutdown due to the COVID-19 pandemic. When included as part of the CHIRON RVs in the fit, we see a statistically significant slope measured of 0.176 m s−1 day−1. However, the RV baseline for CHIRON observations has appeared to shift slightly when pre- and postshutdown RVs were compared for standard stars, consistent with the shift we measured when including these postpandemic RVs (T. Henry 2021, private communication). Since we saw no evidence of a slope in fits without these RVs and we know there is a shift observed in the RV baseline for CHIRON, we treat these observations as a separate instrument (labeled "CHIRON2"). When fitting the CHIRON RVs separately within the EXOFASTv2 fit, we see a difference of 65 m s−1 in the fitted RV zero-points. We note that the planet and host star parameters are consistent to <1σ whether or not this slope is included in the fit.

3.1. TOI-1601 Bimodality

After each EXOFASTv2 fit, we inspect the posteriors of each fitted and derived parameter, visually inspecting for any anomalies such as multimodal distributions. In all cases but TOI-1601 b, no issues were noted. For TOI-1601 b, we see a clear bimodal distribution in the mass and age of the host star (see Figure 6). We find two peaks in the mass distribution at 1.340 and 1.517 M, which correspond to the two peaks seen in the age distribution at 4.27 and 2.63 Gyr. There is no optimal way to represent the bimodal solution, so we split the mass of the host star at the minimum value of 1.415 M and extract two solutions for both peaks identified. We adopt the high-mass solution for the discussion since it has a higher probability (66.7%) of being correct from our analysis, but we present both solutions in Table 4 for future analysis. We note that we observed no significant change in any systematic parameters, suggesting the bimodality is due to our limited precision that is not sufficient to completely separate similar solutions, due to the host star being slightly evolved.

Figure 6.

Figure 6. The (left) Mstar and (right) age probability distribution function for TOI-1601 from our global fit. We split this panel at the valley of M* = 1.415 M and extract two separate solutions, one for each of the peaks in the posteriors (see Table 4). The red line shows the median value for each parameter from the higher-mass solution with a probability of 68.4% (see Section 3).

Standard image High-resolution image

4. Discussion

These five newly discovered hot Jupiters from NASA's TESS mission significantly increase (>5%) the number of well-characterized (measured masses and orbital eccentricities) giant planets that reside in orbits with periods between 5 and 15 days, a regime where planets experience weaker tidal forces than those experienced by planets closer to their host stars and, as a result, are not likely to have had enough time to circularize their orbits. Additionally, these longer-period systems enable us to explore the "reinflation" scenario, an area where TESS should make a huge impact given its high photometric precision. By continuing to discover and characterize new giant planets at longer periods, we can look for evidence (through their eccentricity distribution) of the dominant migration mechanism. Additionally, TESS will provide a complete magnitude-limited sample of hot Jupiters (P < 10 days, Zhou et al. 2019), allowing us to test whether multiple populations exist within the distribution of key parameters (mass, radius, eccentricity), where some tentative trends have been suggested (Nelson et al. 2017; Ikwut-Ukwa et al. 2021). Here we provide a short overview of our global fit results on each of the five new systems. We note that all three of the planets found to be orbiting subgiant host stars (TOI-640, TOI-1333, and TOI-1601) reside in circular orbits.

4.1. TOI-628 b

TOI-628 (TIC 281408474) is a V = 10.18 late-F star with a mass of M* = ${1.311}_{-0.075}^{+0.066}$ M, radius of R* = ${1.345}_{-0.040}^{+0.046}$ R, and an age of ${1.28}_{-0.91}^{+1.6}$ Gyr. Its planetary companion (TOI-628 b) has a radius of RP = ${1.060}_{-0.034}^{+0.041}$ RJ and a mass of MP = ${6.33}_{-0.31}^{+0.29}$ MJ and is on a 3.4096 day period orbit. Our global analysis measures a nonzero orbital eccentricity of e = ${0.072}_{-0.023}^{+0.021}$. Within our global fit, we derive a circularization timescale of τcirc = ${3.32}_{-0.61}^{+0.53}$ Gyr (for this system, assuming equilibrium tides and tidal quality factors for the planet and star of Qp 106 and Q* 106; Adams & Laughlin 2006) for this system, which is longer than our estimated age from MIST of ${1.28}_{-0.91}^{+1.6}$ Gyr. Thus the small but nonzero eccentricity is likely a vestige of the initially high eccentricity that the planet obtained during some process that initiated high-eccentricity migration, a high eccentricity that was subsequently damped to the eccentricity we see today. Without a tighter constraint on the system's age, this is not conclusive. Also of interest is the high mass of TOI-628 b, which makes it one of only a few dozen known hot Jupiters with a mass >6 MJ, and the most massive hot Jupiter found from TESS to date (see Figure 7).

Figure 7.

Figure 7. Left: eccentricity and log of the orbital period of all known giant planets with a mass greater than 0.4 MJ with period between 0.8 and 16 days. The TESS-discovered systems are colored by the host star's effective temperature. The systems with a measured eccentricity from the NASA Exoplanet Archive (NEA) are shown as black circles with errors. Systems where the eccentricity was assumed to be zero are shown with gray crosses. Right: radius and log of the orbital period of all known transiting giant planets. The systems known prior to TESS are in black, while the systems discovered by TESS, including those presented in this paper and Ikwut-Ukwa et al. (2021), are shown as circles colored by their planet's mass.

Standard image High-resolution image

4.2. TOI-640 b

The host star TOI-640 (TIC 147977348) is an F-star with a mass of ${1.526}_{-0.079}^{+0.072}$ M and a radius of ${2.082}_{-0.058}^{+0.064}$ R. The host star appears to be just transitioning off the main sequence onto the subgiant branch, as suggested by our measured $\mathrm{log}g$ = ${3.987}_{-0.036}^{+0.030}$ dex (cm s−2) and corresponding tight age constraint within our global fit from the MIST evolutionary tracks of ${1.99}_{-0.40}^{+0.55}$ Gyr. It hosts a planetary companion, TOI-640 b, which is a highly inflated (RP = ${1.771}_{-0.056}^{+0.060}$ RJ), Jupiter-mass (MP = 0.88 ± 0.16 MJ) planet with a near-integer orbital period of 5.0037775 ± 0.0000048 days. The orbit of the planet is consistent with circular, e = ${0.050}_{-0.035}^{+0.054}$. It is only the third hot Jupiter known with a highly inflated radius (RP > 1.7) and on a period >5 days, joining KELT-12 b (Stevens et al. 2017) and Kepler-435 b (Almenara et al. 2015). Interestingly, TOI-640 b is almost a twin of KELT-12 b, in that they are highly inflated Jupiter-mass planets on ∼5 day orbits around similar subgiant host stars. All three host stars in this regime are evolved, possibly suggesting that the inflation is a result of the host star's recent evolution (Assef et al. 2009; Spiegel & Madhusudhan 2012; Hartman & Bakos 2016; Lopez & Fortney 2016). Similar to KELT-12b and Kepler-435 b, we see no evidence of any significant eccentricity.

4.3. TOI-1333 b

TOI-1333 (TIC 395171208) is a bright (V = 9.49), evolved F-star with a mass of M* = ${1.464}_{-0.079}^{+0.076}$ M and radius of R* = ${1.925}_{-0.063}^{+0.064}$ R. The star appears to be slightly evolved, as suggested by its $\mathrm{log}g$ of ${4.034}_{-0.033}^{+0.032}$ dex (cm s−2). As a result of its evolutionary stage, we estimate a relatively tight age constraint from the MIST evolutionary tracks of ${2.33}_{-0.56}^{+0.71}$ Gyr. Orbiting on a 4.720219 ± 0.000011 day period, the planetary companion TOI-1333 b has a radius of ${1.396}_{-0.054}^{+0.056}$ RJ, a mass of MP = 2.37 ± 0.24 MJ, and an eccentricity that is consistent with circular (e = ${0.073}_{-0.052}^{+0.092}$). This is not surprising given that our derived circularization timescale, ${1.29}_{-0.33}^{+0.37}$ Gyr, is similar to the age of the system.

In the case of TOI-1333, we have measured a periodicity of ∼5.3 days from the ground-based and TESS photometry (we note that WASP identified a period 3× this). We have also measured a $v\sin {I}_{* }$ of 16.5 ± 0.5 km s−1. If the periodicity identified in the photometry indeed is the average rotation period of the host star, then we can estimate the inclination of the host star's rotation axis and compare it to the derived inclination of TOI-1333 b's orbit following the methodology presented in Masuda & Winn (2020). Using the EXOFASTv2 implementation of a Markov Chain Monte Carlo, we run a simple fit of the host star's rotational velocity and its projection onto our line of sight ($v\sin {I}_{* }$) using the values from our global fit for Rstar (${1.925}_{-0.063}^{+0.064}$ R), the derived rotational period of TOI-1333 from TESS and KELT (5.3 days), and the $v\sin {I}_{* }$ from the TRES spectroscopy (14.2 ± 0.5 km s−1) to calculate the inclination of TOI-1333's rotation axis (relative to our line of sight). The latitudes on the Sun that show starspots have a differential rotation on the surface of a few percent. Therefore, we place a 3% error on the rotational velocity for this analysis. We require the same Gelman–Rubin statistic (<1.01) and independent draw (>1000) for convergence as the default for EXOFASTv2. We derive the inclination of the rotation axis to be $51\buildrel{\circ}\over{.} {3}_{-3\buildrel{\circ}\over{.} 3}^{+3\buildrel{\circ}\over{.} 5}$. From our global fit, TOI-1333 b has an inclination of $85\buildrel{\circ}\over{.} {70}_{-0\buildrel{\circ}\over{.} 65}^{+1\buildrel{\circ}\over{.} 3}$, suggesting that the rotation axis of the star and the orbital plane are misaligned. TOI-1333 b is an excellent candidate to confirm this result through spin–orbit alignment (λ) measurements using the Rossiter–McLauglin or Doppler tomography techniques. The planet's orbit is also misaligned with the orbit of the wide binary companion TOI-1333 B, for which we measured an inclination of ${125}_{-10}^{+18}$ degrees from our LOFTI analysis. Interestingly, we do not detect a significant orbital eccentricity from our global fit for TOI-1333 b (though a small eccentricity is still possible), but this suggested misalignment might be a remnant left over from high-eccentricity migration. The likely bound companion at 470 au (see Section 2.6.2) could be responsible for Kozai–Lidov migration of the planet.

4.4. TOI-1478 b

TOI-1478 (TIC 409794137, V = 10.81) is a Sun-like G-dwarf with radius of R* = ${1.048}_{-0.029}^{+0.030}$ R, mass of M* = ${0.946}_{-0.041}^{+0.059}$ M, and an age of ${9.2}_{-3.9}^{+3.1}$ Gyr. Orbiting TOI-1478 is a warm Jupiter with a period of 10.180249 ± 0.000015 days, a radius of RP = ${1.060}_{-0.039}^{+0.040}$ RJ, and a mass of MP = ${0.851}_{-0.047}^{+0.052}$ MJ, and it resides in a circular orbit (e = ${0.024}_{-0.017}^{+0.032}$). TOI-1478 b is the longest-period planet in our sample, and the planet and its host star (other than their orbital distances) resemble the Sun and Jupiter in mass and radius, possibly an example of an alternate outcome of our own solar system. As a result of the long orbital period, the tidal forces on TOI-1478 b are too weak to have circularized the orbit. Therefore, the lack of a significant eccentricity could suggest a more dynamically quiescent migration history.

4.5. TOI-1601 b

In the case of TOI-1601 b, our global model showed a clear bimodality in the posterior distribution of the host star's mass and age (see Section 3.1). This is likely due to the host star's evolutionary status, because the star sits on the Hertzprung–Russell diagram where isochrones cross, so the evolutionary state is ambiguous given the precision of our observations. To account for this, we extract two separate solutions, one for each peak in our posteriors. The higher host star mass solution, M* = ${1.517}_{-0.049}^{+0.053}$ M, has a higher probability of being correct at 66.7%, so we adopt this solution for the discussion, but both results are available in Table 4. TOI-1601 (TIC 139375960, V = 10.71) is an evolved subgiant ($\mathrm{log}g$ of ${3.940}_{-0.025}^{+0.022}$ dex (cm s−2)) with a radius of R* = ${2.186}_{-0.063}^{+0.074}$ R. We estimate the age of the system within our fit to be ${2.64}_{-0.39}^{+0.38}$ Gyr. TOI-1601 b is a Jupiter-mass planet (0.99 ± 0.11 MJ) that shows some inflation (RP = ${1.159}_{-0.059}^{+0.062}$ RJ) and a circular orbit (e = ${0.037}_{-0.026}^{+0.045}$), and a 5.331752 ± 0.000011 day orbit. The spectroscopic analysis of the TRES spectra of TOI-1601 shows some metal enhancement ([Fe/H] = ${0.316}_{-0.074}^{+0.072}$).

4.6. TESS's Impact on Giant Planets

While the primary goal of NASA's TESS mission is to discover and measure the masses of small planets (Ricker et al. 2015), TESS has already provided some valuable discoveries in the field of giant planets (see, e.g., Armstrong et al. 2020; Huang et al. 2020a; Vanderburg et al. 2020). Given the minimum ∼27 day baseline for any target and the completeness in the sensitivity of space-based photometry to detect a hot-Jupiter transit, TESS provides the opportunity to obtain a near-complete sample of hot Jupiters (Zhou et al. 2019). To date, TESS has discovered 26 giant planets (MP > 0.4 MJ), 16 of which have an orbital period >5.0 days (these numbers include the five systems presented in this paper and two additional systems from Ikwut-Ukwa et al. 2021). For comparison, 36 hot Jupiters have been discovered with orbital periods >5.0 days from ground-based transit surveys (NASA Exoplanet Archive, Akeson et al. 2013).

If giant planets predominantly migrate through dynamical interactions, we may find evidence of this evolutionary history in the eccentricity distribution of hot Jupiters, specifically those that are dynamically young (where the circularization timescale by tidal forces is longer than the age of the system). Figure 7 shows the current distribution of giant planet eccentricities as a function of orbital period out to 16 days. Those 26 systems discovered by TESS are shown colored by their host star's effective temperature. Although this is not a homogeneous sample, since a variety of different analysis methods and assumptions were made within this population, there is a wider distribution of eccentricities for those systems with an orbital period >5 days, where tidal circularization timescales are longer (Adams & Laughlin 2006). Interestingly, of the five systems presented here, only TOI-628 b has a statistically significant measured eccentricity (e = ${0.072}_{-0.023}^{+0.021}$) and is consistent with dynamically driven migration since its estimated age is less than the circularization timescale of the orbit. Although the other systems show some nonzero eccentricities from our global fits, they are not statistically significant (>3σ) and could be a result of the Lucy–Sweeney bias (Lucy & Sweeney 1971). We also note that there is one very massive hot Jupiter in our sample, TOI-628 b (MP ${6.33}_{-0.31}^{+0.29}$ MJ), and it is the most massive hot Jupiter discovered to date by TESS (we note that TESS has discovered a few transiting brown dwarfs; Jackman et al. 2019; Carmichael et al. 2020, 2021; Šubjak et al. 2020; and WD 1856+534, which has a mass limit <13.8 MJ, Vanderburg et al. 2020). These massive Jupiters provide a great laboratory for studying the effect of high gravity on the atmosphere of a gas giant, while studying the transition point between giant planets and brown dwarfs.

5. Conclusion

We present the discovery and characterization of five new giant planets (TOI-628, TOI-640 b, TOI-1333 b, TOI-1478 b, and TOI-1601 b) from NASA's TESS mission. These planets were discovered in the primary mission using the 30 minute cadence, full-frame images. Of the systems, TOI-640 b, TOI-1333 b, and TOI-1601 b all orbit stars that appear to have just evolved off the main sequence and are entering the subgiant phase, as suggested by their estimated $\mathrm{log}g$ being under 4.1 dex (cm s−2). None of the planets orbiting these subgiants appear to reside in significantly eccentric orbits. TOI-628 b is the most massive hot Jupiter discovered by TESS (MP = ${6.33}_{-0.31}^{+0.29}$ MJ) and resides in an eccentric orbit that is consistent with dynamically driven migration. Another planet from this work, TOI-640 b, is one of the only highly inflated (>1.7 RJ) hot Jupiters with an orbital period greater than 5 days. TOI-1478 b is the only planet in this sample with an orbital period >10 days, and it and its star are similar in size and mass to Jupiter and the Sun. All five planets orbit bright (V < 10.7) host stars and significantly increase the sample of well-characterized, long-period (>5 day), hot Jupiters, an area where NASA's TESS mission should continue to provide a wealth of discoveries.

C.Z. is supported by a Dunlap Fellowship at the Dunlap Institute for Astronomy & Astrophysics, funded through an endowment established by the Dunlap family and the University of Toronto. T.H. acknowledges support from the European Research Council under the Horizon 2020 Framework Program via the ERC Advanced Grant Origins 83 24 28. J.V.S. acknowledges funding from the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation program (project Four Aces; grant agreement No. 724427). P.R. acknowledges support from NSF grant No. 1952545. R.B. acknowledges support from FONDECYT Project 11200751 and from CORFO project No. 14ENI2-26865. A.J., R.B., and M.H. acknowledge support from project IC120009 "Millennium Institute of Astrophysics (MAS)" of the Millennium Science Initiative, Chilean Ministry of Economy. D.J.S. acknowledges funding support from the Eberly Research Fellowship from The Pennsylvania State University Eberly College of Science. The Center for Exoplanets and Habitable Worlds is supported by the Pennsylvania State University, the Eberly College of Science, and the Pennsylvania Space Grant Consortium. K.K.M. gratefully acknowledges support from the New York Community Trust's Fund for Astrophysical Research. L.G. and A.G. are supported by NASA Massachusetts Space Grant Fellowships. E.W.G., M.E., and P.C. acknowledge support by Deutsche Forschungsgemeinschaft (DFG) grant HA 3279/12-1 within the DFG Schwerpunkt SPP1992, Exploring the Diversity of Extrasolar Planets. B.S.G. was partially supported by the Thomas Jefferson Chair for Space Exploration at the Ohio State University. C.D. acknowledges support from the Hellman Fellows Fund and NASA XRP via grant 80NSSC20K0250.

We thank the CHIRON team members, including Todd Henry, Leonardo Paredes, Hodari James, Azmain Nisak, Rodrigo Hinojosa, Roberto Aviles, Wei-Chun Jao, and CTIO staffs, for their work in acquiring RVs with CHIRON at CTIO. This research has made use of SAO/NASA's Astrophysics Data System Bibliographic Services. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France. This work has made use of data from the European Space Agency (ESA) mission Gaia (https://www.cosmos.esa.int/gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC; https://www.cosmos.esa.int/web/gaia/dpac/consortium). Funding for the DPAC has been provided by national institutions, in particular the institutions participating in the Gaia Multilateral Agreement. This work makes use of observations from the LCO network. This work is based in part on observations obtained at the Southern Astrophysical Research (SOAR) telescope, which is a joint project of the Ministério da Ciência, Tecnologia e Inovações (MCTI/LNA) do Brasil, the U.S. National Science Foundation's NOIRLab, the University of North Carolina at Chapel Hill (UNC), and Michigan State University (MSU).

Funding for the TESS mission is provided by NASA's Science Mission directorate. We acknowledge the use of public TESS Alert data from pipelines at the TESS Science Office and at the TESS Science Processing Operations Center. This research has made use of the NASA Exoplanet Archive and the Exoplanet Follow-up Observation Program website, which are operated by the California Institute of Technology, under contract with the National Aeronautics and Space Administration under the Exoplanet Exploration Program. This paper includes data collected by the TESS mission, which are publicly available from the Mikulski Archive for Space Telescopes (MAST). This paper includes observations obtained under Gemini program GN-2018B-LP-101. 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. This publication makes use of The Data & Analysis Center for Exoplanets (DACE), which is a facility based at the University of Geneva (CH) dedicated to extrasolar planet data visualization, exchange, and analysis. DACE is a platform of the Swiss National Centre of Competence in Research (NCCR) PlanetS, federating the Swiss expertise in exoplanet research. The DACE platform is available at https://dace.unige.ch.

Some of the data presented herein were obtained at the W. M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California, and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W. M. Keck Foundation. The authors wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Maunakea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain.

Minerva-Australis is supported by Australian Research Council LIEF Grant LE160100001, Discovery Grant DP180100972, Mount Cuba Astronomical Foundation, and institutional partners University of Southern Queensland, UNSW Sydney, MIT, Nanjing University, George Mason University, University of Louisville, University of California Riverside, University of Florida, and The University of Texas at Austin. We respectfully acknowledge the traditional custodians of all lands throughout Australia, and recognize their continued cultural and spiritual connection to the land, waterways, cosmos, and community. We pay our deepest respects to all Elders, ancestors and descendants of the Giabal, Jarowair, and Kambuwal nations, upon whose lands the Minerva-Australis facility at Mt. Kent is situated.

Facilities: TESS, FLWO 1.5 m (Tillinghast Reflector Echelle Spectrograph), 4.1 m Southern Astrophysical Research (SOAR), LCO 0.4 m, LCO 1.0 m, 2.2 m telescope La Silla (Fiber-fed Extended Range Optical Spectrograph), KECK (NIRC2), PALOMAR (PHARO), TESS, KELT, WASP, CTIO 1.5 m (CHIRON), Minerva-Australis, GEMINI (NIRI).

Software: EXOFASTv2 (Eastman et al. 2013, 2019), AstroImageJ (Collins et al. 2017), TAPIR (Jensen 2013), PEST Pipeline (http://pestobservatory.com/the-pest-pipeline/), LOFTI (Pearce et al. 2020), Isochrones package (Morton 2015), QLP Pipeline (Huang et al. 2020b), CETES (Brahm et al. 2017).

Footnotes

Please wait… references are loading.
10.3847/1538-3881/abe38a