Brought to you by:

Articles

HAT-P-20b–HAT-P-23b: FOUR MASSIVE TRANSITING EXTRASOLAR PLANETS*

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

Published 2011 November 15 © 2011. The American Astronomical Society. All rights reserved.
, , Citation G. Á. Bakos et al 2011 ApJ 742 116 DOI 10.1088/0004-637X/742/2/116

0004-637X/742/2/116

ABSTRACT

We report the discovery of four relatively massive (2–7 MJ) transiting extrasolar planets. HAT-P-20b orbits the moderately bright V = 11.339 K3 dwarf star GSC 1910-00239 on a circular orbit, with a period P = 2.875317  ±  0.000004 days, transit epoch Tc = 2455080.92661  ±  0.00021 (BJDUTC), and transit duration 0.0770  ±  0.0008 days. The host star has a mass of 0.76  ±  0.03 M, radius of 0.69  ±  0.02 R, effective temperature 4595  ±  80 K, and metallicity [Fe/H] = +0.35  ±  0.08. The planetary companion has a mass of 7.246  ±  0.187 MJ and a radius of 0.867  ±  0.033 RJ yielding a mean density of 13.78  ±  1.50 g cm−3. HAT-P-21b orbits the V = 11.685 G3 dwarf star GSC 3013-01229 on an eccentric (e = 0.228  ±  0.016) orbit, with a period P = 4.124481  ±  0.000007 days, transit epoch Tc = 2454996.41312  ±  0.00069, and transit duration 0.1530  ±  0.0027 days. The host star has a mass of 0.95  ±  0.04 M, radius of 1.10  ±  0.08 R, effective temperature 5588  ±  80 K, and metallicity [Fe/H] = +0.01  ±  0.08. The planetary companion has a mass of 4.063  ±  0.161 MJ and a radius of 1.024  ±  0.092 RJ yielding a mean density of 4.68+1.59− 0.99 g cm−3. HAT-P-21b is a borderline object between the pM and pL class planets, and the transits occur near apastron. HAT-P-22b orbits the bright V = 9.732 G5 dwarf star HD 233731 on a circular orbit, with a period P = 3.212220  ±  0.000009 days, transit epoch Tc = 2454930.22001  ±  0.00025, and transit duration 0.1196  ±  0.0014 days. The host star has a mass of 0.92  ±  0.03 M, radius of 1.04  ±  0.04 R, effective temperature 5302  ±  80 K, and metallicity [Fe/H] = +0.24  ±  0.08. The planet has a mass of 2.147  ±  0.061 MJ and a compact radius of 1.080  ±  0.058 RJ yielding a mean density of 2.11+0.40− 0.29 g cm−3. The host star also harbors an M-dwarf companion at a wide separation. Finally, HAT-P-23b orbits the V = 12.432 G0 dwarf star GSC 1632-01396 on a close to circular orbit, with a period P = 1.212884  ±  0.000002 days, transit epoch Tc = 2454852.26464  ±  0.00018, and transit duration 0.0908  ±  0.0007 days. The host star has a mass of 1.13  ±  0.04 M, radius of 1.20  ±  0.07 R, effective temperature 5905  ±  80 K, and metallicity [Fe/H] = +0.15  ±  0.04. The planetary companion has a mass of 2.090  ±  0.111 MJ and a radius of 1.368  ±  0.090 RJ yielding a mean density of 1.01  ±  0.18 g cm−3. HAT-P-23b is an inflated and massive hot Jupiter on a very short period orbit, and has one of the shortest characteristic infall times (7.5+2.9− 1.8 Myr) before it gets engulfed by the star.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The majority of the ∼130 confirmed transiting extrasolar planets (TEPs) have been found to lie in the 0.5–2.0 MJ mass range. The apparent drop in their mass distribution at ∼2 MJ has been noted by, e.g., Southworth et al. (2009), and by Torres et al. (2010). In the currently known sample, 75% of the TEPs have planetary mass Mp < 2.0 MJ, and there appears to be a minor peak in their occurrence rate at Mp ≈ 2 MJ, which then sharply falls off toward higher masses.

Are there any biases present against discovering massive planets? Such planets tend to be less inflated, and theory dictates that their radii shrink as their mass increases toward the brown dwarf regime. According to Baraffe et al. (2010), this reversal of the MpRp relation happens around Mp ≈ 2–3 MJ, and falls off as RpM−1/8p (see, e.g., Fortney et al. 2010). The smaller radii for massive planets yield a minor bias against discovering them via the transit method, since they produce shallower transits. Very massive planets can perhaps induce stellar variability of their host stars (Shkolnik et al. 2009), somewhat decreasing the efficiency of detecting their shallow transits via simple algorithms that expect constant out-of-transit light curves. Also, the host stars of massive planets are typically more rapid rotators: the average vsin i for host stars with planets Mp < 2 MJ is 3.7 m s−1 (with 4.1 m s−1 standard deviation around the mean), whereas the same values for the massive planet host stars are 10.1 m s−1(with 10.8 m s−1 standard deviation around the mean).13 Six of the seven fastest rotators all harbor planets more massive than 2 MJ. This presents a bias against discovering them either via radial velocity (RV) searches, which are more efficient around quiet non-rotating dwarfs, or via transit searches, where the targets may be discarded during the confirmation phase. Along the same lines, the large RV amplitude of the host star, as caused by the planetary companion, may even lead to erroneous rejection during the reconnaissance phase of candidate confirmation, since such systems resemble eclipsing binaries. Finally, there is a tendency that massive planets are more likely to be eccentric14 (Southworth et al. 2009), meaning that they require more RV observations for proper mapping of their orbits, and thus leading to a slower announcement rate. On the other hand, a strong bias for detecting such planets—compensating for most of the effects above—is the fact that the large RV amplitudes of the host stars are easier to detect, since they do not require internal precisions at the meters per second level (see HAT-P-2, where valuable data were contributed to the RV fit by modest precision instruments yielding ∼1 km s−1 precision; Bakos et al. 2007). Altogether, while there are minor biases against detecting massive transiting planets, their overall effect is likely mitigated by the strong bias in favor of their detection, and the drop in frequency at ≳ 2 MJ seems to be real.

Massive planets are important for many reasons. They provide very strong constraints on formation and migration theories, which need to explain the observed distribution of planetary system parameters in a wide range (Baraffe et al. 2008, 2010), from 0.01 MJ (Corot-7b; Queloz et al. 2009) to 26.4 MJ (Corot-3b; Deleuil et al. 2008). Heavy mass objects necessitate the inclusion of other physical mechanisms for formation and migration, such as planet–planet scattering (Chatterjee et al. 2008; Ford & Rasio 2008), and the Kozai mechanism (Fabrycky & Tremaine 2007). They are borderline objects between planets and brown dwarfs, and help us understand how these populations differ and overlap (see Leconte et al. 2009 for a review). For example, a traditional definition of planets is that they have no deuterium burning, where the deuterium-burning limit is thought to be around 13 MJ (Spiegel et al. 2011). However, there are large uncertainties on this limit due to the numerous model parameters and solutions, and the fact that deuterium may be able to burn in the H/He layers above the core (Baraffe et al. 2008). Another possible definition of planets is based on their formation scenario, i.e., they are formed by accretion in a protoplanetary disk around their young host star, as opposed to the gravitational collapse of a molecular cloud (brown dwarfs).15

Perhaps related to the formation and migration mechanisms, a number of interesting correlations involving massive planets have been pointed out. Udry et al. (2002) noted that short-period massive planets are predominantly found in binary stellar systems. Southworth et al. (2009) noted that only 8.6% of the low-mass planets show significantly eccentric orbits, whereas 77% of the massive planets have eccentric orbits (although low-mass systems have lower signal-to-noise ratio (S/N) RV curves, rendering the detection of eccentric orbits more difficult). Curiously, there appears to be a lack of correlation between planetary mass and host star metallicity, while one would naively think that the formation of high-mass planets (via core accretion) would require higher metal content. Until this work, there was a hint of a correlation between planetary and stellar mass (e.g., Deleuil et al. 2008), in the sense that the most massive planets orbited M ≳ 1.2 M stars, and there was a (biased) tendency that lower mass planets orbit less massive stars.

All of these observations suffer from small-number statistics and heavy biases. One way of improving our knowledge is to expand the sample of well-characterized planets. In this work we report on four new massive transiting planets around bright stars. This extends the currently known sample of bright (V < 13.5) and massive (Mp > 2 MJ) transiting planets by 30% (from 13 to 17).16 These discoveries were made by the Hungarian-made Automated Telescope Network (HATNet; Bakos et al. 2004) survey. HATNet has been one of the main contributors to the discovery of TEPs, among others such as the ground-based SuperWASP (Pollacco et al. 2006), TrES (Alonso et al. 2004), and XO projects (McCullough et al. 2005), and space-borne searches such as CoRoT (Baglin et al. 2006) and Kepler (Borucki et al. 2010). In operation since 2003, HATNet has now covered approximately 14% of the sky, searching for TEPs around bright stars (8 ≲ I ≲ 14). We operate six wide-field instruments: four at the Fred Lawrence Whipple Observatory (FLWO) in Arizona and two on the roof of the hangar servicing the Smithsonian Astrophysical Observatory's Submillimeter Array, in Hawaii.

The layout of this paper is as follows. In Section 2 we report the detections of the photometric signals and the follow-up spectroscopic and photometric observations for each of the planets. In Section 3 we describe the analysis of the data, beginning with the determination of the stellar parameters, continuing with a discussion of the methods used to rule out non-planetary, false positive scenarios which could mimic the photometric and spectroscopic observations, and finishing with a description of our global modeling of the photometry and RVs. Our findings are discussed in Section 4.

2. OBSERVATIONS

2.1. Photometric Detection

Table 1 summarizes the HATNet discovery observations of each new planetary system. The calibration of the HATNet frames was carried out using standard procedures correcting for the CCD bias, dark current, and flat-field structure. The calibrated images were then subjected to star detection and astrometry, as described in Pál & Bakos (2006). Aperture photometry was performed on each image at the stellar centroids derived from the Two Micron All Sky Survey (2MASS; Skrutskie et al. 2006) catalog and the individual astrometric solutions. For certain data sets (HAT-P-20, HAT-P-22, HAT-P-23) we also carried out an image subtraction (Alard 2000) based photometric reduction using discrete kernels (Bramich 2008), as described in Pál (2009b). The resulting light curves were decorrelated (cleaned of trends) using the External Parameter Decorrelation (EPD; see Bakos et al. 2010) technique in "constant" mode and the Trend Filtering Algorithm (TFA; see Kovács et al. 2005). The light curves were searched for periodic box-shaped signals using the Box Least-squares (BLS; see Kovács et al. 2002) method. Candidates were passed through a number of automatic filters to reject false alarms and obvious eclipsing binaries or other variable stars (two examples include checking for differences between the depths of even and odd transits when the period is doubled, and checking for evidence of a secondary eclipse at the period of the transit), and then passed a by-eye selection procedure which primarily rejects obvious blends with nearby eclipsing binaries or cases where the candidate selection is due to an artifact which is not captured by one of the automated filters. We detected significant signals in the light curves of the stars as summarized below.

  • 1.  
    HAT-P-20—GSC 1910-00239 (also known as 2MASS 07273995+2420118; $\alpha = 07^{\mathrm h}27^{\mathrm m}39\mbox{$.\!\!^{\mathrm s}$}96$, δ = +24°20'11farcs9; J2000; V = 11.339). A signal was detected for this star with an apparent depth of ∼10.8 mmag, and a period of P = 2.8753 days (see Figure 1). Note that the apparent HATNet depth stated here is attenuated by the presence of the fainter neighbor star that is not resolved on the coarse resolution (9farcs5 pixel−1) HATNet pixels. Also, the depth by fitting a trapezoid instead of the correct Mandel & Agol (2002) model is somewhat shallower than the maximum depth in the Mandel & Agol (2002) model fit (which was 19.6 mmag; see later in Section 3.3). Note that the dilution in the HATNet light curve is accounted for by the dilution factor parameter included in the global fitting procedure described in Section 3.3. The follow-up light curves discussed in Section 2.4 are obtained with instruments that have much higher spatial resolution, and when necessary are reduced using the image subtraction technique, so that they are not affected by dilution from any known neighbors. The drop in brightness had a first-to-last-contact duration, relative to the total period, of q = 0.0268  ±  0.0003, corresponding to a total duration of Pq = 1.848  ±  0.019 hr. HAT-P-20 has a red companion (2MASS 07273963+2420171, JK = 0.92) at 6farcs86 separation that is fainter than HAT-P-20 by ΔR = 1.36 mag.
  • 2.  
    HAT-P-21—GSC 3013-01229 (also known as 2MASS 11250598+4101406; $\alpha = 11^{\mathrm h}25^{\mathrm m}05\mbox{$.\!\!^{\mathrm s}$}88$, δ = +41°01'40farcs6; J2000; V = 11.685). A signal was detected for this star with an apparent depth of ∼8.0 mmag, and a period of P = 4.1245 days (see Figure 2). The drop in brightness had a first-to-last-contact duration, relative to the total period, of q = 0.0371  ±  0.0007, corresponding to a total duration of Pq = 3.672  ±  0.065 hr.
  • 3.  
    HAT-P-22—HD 233731 (also known as GSC 03441-00925 and 2MASS 10224361+5007420; $\alpha = 10^{\mathrm h}22^{\mathrm m}43\mbox{$.\!\!^{\mathrm s}$}68$, δ = +50°07'42farcs0; J2000; V = 9.732). A signal was detected for this star with an apparent depth of ∼9.7 mmag, and a period of P = 3.2122 days (see Figure 3). The drop in brightness had a first-to-last-contact duration, relative to the total period, of q = 0.0372  ±  0.0004, corresponding to a total duration of Pq = 2.869  ±  0.033 hr. HAT-P-22 has a close red companion star (2MASS 10224397+5007504, JK = 0.86) at 9farcs1 separation and Δi = 2.58 mag fainter.
  • 4.  
    HAT-P-23—GSC 1632-01396 (also known as 2MASS 20242972+1645437; $\alpha = 20^{\mathrm h}24^{\mathrm m}29\mbox{$.\!\!^{\mathrm s}$}88$, δ = +16°45'43farcs7; J2000; V = 12.432). A signal was detected for this star with an apparent depth of ∼11.5 mmag, and a period of P = 1.2129 days (see Figure 4). Similarly to HAT-P-20, the depth was attenuated by close-by faint neighbors. The drop in brightness had a first-to-last-contact duration, relative to the total period, of q = 0.0748  ±  0.0006, corresponding to a total duration of Pq = 2.178  ±  0.017 hr.
Figure 1.

Figure 1. Unbinned light curve of HAT-P-20 including all 2600 instrumental R-band 5.5 minute cadence measurements obtained with the HAT-7 and HAT-8 telescopes of HATNet (see Table 1 for details), and folded with the period P = 2.8753172 days resulting from the global fit described in Section 3. The solid line shows the "P1P3" transit model fit to the light curve (Section 3.3). The bold points in the lower panel show the light curve binned in phase with a bin size of 0.002.

Standard image High-resolution image
Figure 2.

Figure 2. Unbinned light curve of HAT-P-21 including all 28,000 instrumental I-band and R-band 5.5 minute cadence measurements obtained with the HAT-5, HAT-6, HAT-8, and HAT-9 telescopes of HATNet (see Table 1 for details), and folded with the period P = 4.1244808 days resulting from the global fit described in Section 3. The solid line shows the "P1P3" transit model fit to the light curve (Section 3.3). The bold points in the lower panel show the light curve binned in phase with a bin size of 0.002.

Standard image High-resolution image
Figure 3.

Figure 3. Unbinned light curve of HAT-P-22 including all 4200 instrumental R-band 5.5 minute cadence measurements obtained with the HAT-5 telescope of HATNet (see the text for details), and folded with the period P = 3.2122198 days resulting from the global fit described in Section 3. The solid line shows the "P1P3" transit model fit to the light curve (Section 3.3). The bold points in the lower panel show the light curve binned in phase with a bin size of 0.002.

Standard image High-resolution image
Figure 4.

Figure 4. Unbinned light curve of HAT-P-23 including all 3500 instrumental R-band 5.5 minute cadence measurements obtained with the HAT-6 and HAT-9 telescopes of HATNet (see Table 1 for details), and folded with the period P = 1.2128841 days resulting from the global fit described in Section 3. The solid line shows the "P1P3" transit model fit to the light curve (Section 3.3). The bold points in the lower panel show the light curve binned in phase with a bin size of 0.002.

Standard image High-resolution image

Table 1. Summary of Photometric Observations

Instrument/Field Date Number Median FWHM Cadence Filter
  (s) of Images of PSF ('') (s)  
HAT-P-20          
  HAT-7/G267 2007 Dec–2008 May 802 32.8 330 R
  HAT-8/G267 2007 Oct–2008 May 1850 25.3 330 R
  FLWO12/KeplerCam 2009 Mar 11 268 3.5 43 Sloan i
  FLWO12/KeplerCam 2009 Oct 21 343 3.5 32 Sloan i
HAT-P-21          
  HAT-6/G183 2006 Dec–2007 May 4528 37.1 330 I
  HAT-9/G183 2006 Nov–2007 Jun 4586 40.5 330 I
  HAT-5/G184 2006 Dec–2007 Jun 4040 36.0 330 I
  HAT-8/G184 2006 Dec–2007 Jun 5606 40.5 330 I
  HAT-6/G141 2008 Jan–2008 Jun 5142 25.9 330 R
  HAT-9/G141 2008 Jan–2008 Jun 3964 27.5 330 R
  FLWO12/KeplerCam 2009 Apr 20 243 3.9 53 Sloan i
  FLWO12/KeplerCam 2010 Feb 15 412 4.8 43 Sloan i
  LCOGT/FTNa 2010 Feb 19 511 3.0 31 Sloan i
HAT-P-22          
  HAT-5/G139 2007 Dec–2008 May 4288 27.7 330 R
  FLWO12/KeplerCam 2009 Feb 28 532 3.4 28 Sloan z
  FLWO12/KeplerCam 2009 Apr 30 353 3.8 33 Sloan g
HAT-P-23          
  HAT-6/G341 2007 Sep–2007 Dec 1178 24.7 330 R
  HAT-9/G341 2007 Sep–2007 Nov 2351 27.7 330 R
  FLWO12/KeplerCam 2008 Jun 14 147 3.0 73 Sloan i
  FLWO12/KeplerCam 2008 Sep 8 246 3.6 73 Sloan i
  FLWO12/KeplerCam 2008 Sep 13 265 3.2 73 Sloan i
  FLWO12/KeplerCam 2008 Nov 3 117 3.3 89 Sloan i
  FLWO12/KeplerCam 2009 Apr 19 46 4.6 150 Sloan g
  FLWO12/KeplerCam 2009 Jul 13 150 4.1 73 Sloan i

Note. aObservations were performed without guiding due to a technical problem with the guiding system, and resulted in decreased data quality.

Download table as:  ASCIITypeset image

2.2. Reconnaissance Spectroscopy

As is routine in the HATNet project, all candidates are subjected to careful scrutiny before investing valuable time on large telescopes. This includes spectroscopic observations at relatively modest facilities to establish whether the transit-like feature in the light curve of a candidate might be due to astrophysical phenomena other than a planet transiting a star. Many of these false positives are associated with large RV variations in the star (tens of kilometers per second) that are easily recognized. The reconnaissance spectroscopic observations and results for each system are summarized in Table 2; below we provide a brief description of the instruments used, the data reduction, and the analysis procedure.

Table 2. Summary of Reconnaissance Spectroscopy Observations

Instrument Date(s) Number of Spectra Teff⋆ log g vsin i γRVa
      (K) (cgs) (km s−1) (km s−1)
HAT-P-20            
  DS 2009 Feb 11–2009 Feb 15 3 4500 ± 125 4.0 ± 0.25 0 ± 40 −18.81 ± 0.68
  FIES 2009 Oct 7 1 4500 4.0 4 −16.76 ± 0.1
HAT-P-21            
  DS 2009 Mar 8–2009 Apr 5 3 5750 ± 125 4.5 ± 0.25 3.0 ± 3.0 −53.19 ± 0.09
HAT-P-22            
  DS 2009 Feb 11–2009 Feb 16 4 5250 ± 125 4.5 ± 0.25 2.0 ± 2.0 +12.49 ± 0.28
HAT-P-23            
  DS 2008 May 19–2008 Sep 14 5 6000 ± 125 4.5 ± 0.25 9.0 ± 1.0 −15.10 ± 0.30

Note. aThe mean heliocentric RV of the target (in the IAU system). For the DS observations, the error gives the rms of the individual velocity measures for the target. For the FIES observation this is the estimated uncertainty on the individual measurement.

Download table as:  ASCIITypeset image

One of the tools we have used for this purpose is the Harvard-Smithsonian Center for Astrophysics (CfA) Digital Speedometer (DS; Latham 1992), an echelle spectrograph mounted on the FLWO 1.5 m telescope. This instrument delivers high-resolution spectra (λ/Δλ ≈ 35, 000) over a single order centered on the Mg i b triplet (∼5187 Å), with typically low S/Ns that are nevertheless sufficient to derive RVs with moderate precisions of 0.5–1.0 km s−1 for slowly rotating stars. The same spectra can be used to estimate the effective temperature, surface gravity, and projected rotational velocity of the host star, as described by Torres et al. (2002). With this facility we are able to reject many types of false positives, such as F dwarfs orbited by M dwarfs, grazing eclipsing binaries, and many triple or quadruple star systems. Additional tests are performed with other spectroscopic material described in the next section.

Another of the tools we have used for this purpose is the Fiber-fed Échelle Spectrograph (FIES) at the 2.5 m Nordic Optical Telescope (NOT) at La Palma, Spain (Djupvik & Andersen 2010). We used the medium-resolution fiber which produces spectra at a resolution of λ/Δλ ≈ 46, 000 and a wavelength coverage of ∼3600–7400 Å to observe HAT-P-20. The spectrum was extracted and analyzed to measure the RV, effective temperature, surface gravity, and projected rotation velocity of the host star, following the procedures described by Buchhave et al. (2010).

Based on the observations summarized in Table 2, we find that HAT-P-21, HAT-P-22, and HAT-P-23 have rms residuals consistent with no detectable RV variation within the precision of the measurements. The reconnaissance observations of HAT-P-20 hinted at a K ∼ 1 km s−1 orbital variation roughly in phase with the photometric ephemeris (these velocities are consistent with the K ∼ 1.2 km s−1 orbit of HAT-P-20 measured subsequently with Keck/HIRES given the ∼0.5 km s−1 precision of the DS). All spectra were single-lined, i.e., there is no evidence that any of these targets consist of more than one star. The gravities for all of the stars indicate that they are dwarfs.

2.3. High-resolution, High-S/N Spectroscopy

We proceeded with the follow-up of each candidate by obtaining high-resolution, high-S/N spectra to characterize the RV variations, and to refine the determination of the stellar parameters. These observations are summarized in Table 3. The RV measurements and uncertainties are given in Tables 47 for HAT-P-20–HAT-P-23, respectively. The period-folded data, along with our best fits described below in Section 3, are displayed in Figures 69 for HAT-P-20–HAT-P-23. Below we briefly describe the instruments used, the data reduction, and the analysis procedure.

Table 3. Summary of High-resolution/High-S/N Spectroscopic Observations

Instrument Date(s) Number of
    RV Obs.
HAT-P-20    
  Keck/HIRES 2009 Apr–2009 Dec 10
HAT-P-21    
  Keck/HIRES 2009 May–2010 Feb 15
HAT-P-22    
  Keck/HIRES 2009 Apr–2009 Dec 12
HAT-P-23    
  Keck/HIRES 2008 Jun–2009 Dec 13

Download table as:  ASCIITypeset image

Table 4. Relative Radial Velocities, Bisector Spans, and Activity Index Measurements of HAT-P-20

BJDUTCa RVb σRVc BS σBS Sd σS
(2,454,000+) (m s−1) (m s−1) (m s−1) (m s−1)    
934.84149 ... −1139.51 2.94 ... ... ... ...
954.82485 ... ... ... −8.00 3.22 2.051 0.075
954.83175 ... −942.88 2.24 −18.65 8.02 2.900 0.035
956.80481 ... 1092.32 2.28 −42.21 10.81 3.305 0.018
983.73851 ... −1151.02 2.39 18.53 4.51 2.853 0.016
984.74064 ... 328.09 2.14 6.85 6.11 2.879 0.016
986.74296 ... −1225.77 2.34 15.28 7.78 2.588 0.029
1107.15226. −870.02 4.89 10.00 17.65 1.352 0.043
1109.05752. 1233.55 3.76 −1.23 7.66 2.783 0.018
1191.10274. −1147.72 2.07 14.55 6.16 3.412 0.016
1192.11754. 1100.72 2.12 4.88 2.83 3.346 0.015

Notes. Note that for the iodine-free template exposures we do not measure the RV but do measure the BS and S index. Such template exposures can be distinguished by the missing RV value. aBarycentric Julian dates throughout the paper are calculated from Coordinated Universal Time (UTC). bThe zero-point of these velocities is arbitrary. An overall offset γrel fitted to these velocities in Section 3.3 has not been subtracted. cInternal errors excluding the component of astrophysical jitter considered in Section 3.3. dRelative chromospheric activity index, not calibrated to the scale of Vaughan et al. (1978).

Download table as:  ASCIITypeset image

Table 5. Relative Radial Velocities, Bisector Spans, and Activity Index Measurements of HAT-P-21

BJDUTC RV σRV BS σBS S σS
(2,454,000+) (m s−1) (m s−1) (m s−1) (m s−1)    
954.91548 ... ... ... 3.67 5.10 0.819 0.002
954.92307 ... 221.51 2.88 −9.60 13.62 0.829 0.003
955.86532 ... −274.01 2.48 −8.08 5.81 0.823 0.003
955.98849 ... −314.29 2.76 7.51 12.62 0.855 0.005
957.00006 ... −378.18 3.00 2.88 8.36 0.866 0.009
963.95573 ... −197.70 2.68 −12.46 5.48 0.789 0.003
983.85638 ... 212.70 2.72 −4.61 3.33 0.815 0.003
985.87988 ... −406.23 2.46 5.88 4.95 0.809 0.003
986.85645 ... 509.76 3.01 −0.70 7.24 0.803 0.005
987.86480 ... 270.66 2.92 −6.96 4.20 0.810 0.004
1174.14589. −145.85 3.13 0.01 6.54 0.833 0.005
1191.14119. −391.67 2.75 5.76 5.16 0.832 0.005
1192.12302. −457.57 2.67 7.04 3.15 0.809 0.005
1193.11001. 496.07 2.58 2.50 3.85 0.794 0.005
1197.15057. 459.61 2.59 9.69 4.91 0.777 0.005
1229.12322. −469.62 2.81 −2.53 5.81 0.821 0.005

Note. Same as those of Table 4.

Download table as:  ASCIITypeset image

Table 6. Relative Radial Velocities, Bisector Spans, and Activity Index Measurements of HAT-P-22

BJDUTC RV σRV BS σBS S σS
(2,454,000+) (m s−1) (m s−1) (m s−1) (m s−1)    
928.94934 ... 187.55 1.12 13.71 4.73 0.557 0.002
954.88137 ... 284.38 1.12 7.07 3.69 0.556 0.002
954.88605 ... ... ... −2.17 3.34 0.554 0.001
955.79962 ... 77.44 1.46 6.00 13.88 0.533 0.008
956.96383 ... −273.80 1.30 4.41 10.92 0.531 0.005
963.92157 ... −14.63 1.33 −0.17 6.57 0.570 0.002
983.82710 ... 282.58 0.98 2.20 3.79 0.572 0.003
985.81064 ... −298.55 1.14 −2.81 2.96 0.560 0.002
986.84124 ... 239.74 1.39 −7.48 6.58 0.553 0.003
988.83025 ... −300.15 1.31 −16.80 8.24 0.555 0.002
1191.14866. −326.24 1.32 −0.33 5.00 0.843 0.003
1193.10170. 255.62 1.30 −8.71 6.79 0.835 0.003
1193.93435. −192.43 1.38 5.08 5.27 0.843 0.004

Note. Same as those of Table 4.

Download table as:  ASCIITypeset image

Table 7. Relative Radial Velocities, Bisector Spans, and Activity Index Measurements of HAT-P-23

BJDUTC RV σRV BS σBS S σS
(2,454,000+) (m s−1) (m s−1) (m s−1) (m s−1)    
638.09243 ... ... ... −4.89 4.59 1.170 0.020
638.10601 ... −137.90 4.49 −0.45 5.22 0.797 0.006
674.91404 ... 374.64 5.15 −10.71 3.99 0.794 0.006
723.78719 ... −209.16 5.55 −1.68 13.38 0.738 0.006
725.85309 ... 329.57 6.18 12.86 6.55 0.749 0.006
726.87976 ... 277.98 4.20 −26.82 6.17 0.770 0.005
727.76247 ... −210.01 4.81 −35.11 5.18 0.811 0.006
727.89619 ... −19.91 4.77 −30.29 5.05 0.795 0.005
777.82189 ... 214.73 4.54 −12.64 4.43 0.754 0.005
810.72960 ... 290.22 5.63 −13.57 5.59 0.726 0.006
955.04676 ... 388.82 4.74 −15.76 3.57 0.761 0.006
985.99494 ... −369.93 4.42 −19.69 5.06 0.718 0.006
1192.70812. 348.01 4.40 −14.40 4.53 0.726 0.005
1193.71080. 83.68 4.57 −20.67 4.23 0.825 0.008

Notes. Same as those of Table 4.

Download table as:  ASCIITypeset image

Observations were made of all four planet host stars with the HIRES instrument (Vogt et al. 1994) on the Keck I telescope located on Mauna Kea, Hawaii. The width of the spectrometer slit was 0farcs86, resulting in a resolving power of λ/Δλ ≈ 55, 000, with a wavelength coverage of ∼3800–8000 Å. We typically used the B5 decker yielding a 3farcs5(H) × 0farcs861(W) slit, and for the last few observations on each target we used the C2 decker that enables a better sky subtraction due to the longer slit 14farcs0(H) × 0farcs861(W). The slit height was oriented with altitude (vertical), except for rare cases, when the slit would have run through the faint companion to HAT-P-20 or HAT-P-22. A Keck/HIRES snapshot for each planet host star is shown in Figure 5. Spectra were obtained through an iodine gas absorption cell, which was used to superimpose a dense forest of I2 lines on the stellar spectrum and establish an accurate wavelength fiducial (see Marcy & Butler 1992). For each target an additional exposure was taken without the iodine cell, for use as a template in the reductions. Relative RVs in the solar system barycentric frame were derived as described by Butler et al. (1996), incorporating full modeling of the spatial and temporal variations of the instrumental profile.

Figure 5.

Figure 5. Keck/HIRES guider camera snapshots of HAT-P-20–HAT-P-23 (labeled). North is up and east is to the left. The snapshots cover an area of approximately 30'' × 20''. The slit is also visible, as positioned on the planet host stars.

Standard image High-resolution image

In each of Figures 69 we show also the relative S index, which is a measure of the chromospheric activity of the star derived from the flux in the cores of the Ca ii H and K lines. This index was computed following the prescription given by Vaughan et al. (1978), and as described in Hartman et al. (2009). Note that our relative S index has not been calibrated to the scale of Vaughan et al. (1978). We do not detect any significant variation of the index correlated with orbital phase; such a correlation might have indicated that the RV variations could be due to stellar activity, casting doubt on the planetary nature of the candidate.

Figure 6.

Figure 6. Top panel: Keck/HIRES RV measurements for HAT-P-20 shown as a function of orbital phase, along with our best-fit eccentric model (see Table 14). Zero phase corresponds to the time of mid-transit. The center-of-mass velocity has been subtracted. The rms around the best orbital fit is 16.2 m s−1. Second panel: velocity OC residuals from the best fit. The error bars for both the top and second panels include a component from the jitter (15.7 m s−1) added in quadrature to the formal errors (see Section 3.3). Third panel: bisector spans (BS), with the mean value subtracted. The measurement from the template spectrum is included (see Section 3.2). Bottom panel: relative chromospheric activity index S measured from the Keck spectra.

Standard image High-resolution image
Figure 7.

Figure 7. Keck/HIRES observations of HAT-P-21. The panels are as in Figure 6. The parameters used in the best-fit model are given in Table 14, the RV jitter was 26.4 m s−1, and the fit rms was 26.6 m s−1. Observations shown twice are represented with open symbols.

Standard image High-resolution image
Figure 8.

Figure 8. Keck/HIRES observations of HAT-P-22. The panels are as in Figure 6. The parameters used in the best-fit model are given in Table 14, the RV jitter was 9.9 m s−1, and the fit rms was 10.0 m s−1. Observations shown twice are represented with open symbols.

Standard image High-resolution image
Figure 9.

Figure 9. Keck/HIRES observations of HAT-P-23. The panels are as in Figure 6. The parameters used in the best-fit model are given in Table 14, and the RV jitter was 34.7 m s−1. Observations shown twice are represented with open symbols.

Standard image High-resolution image

2.4. Photometric Follow-up Observations

In order to permit a more accurate modeling of the light curves, we conducted additional photometric observations with the KeplerCam CCD camera on the FLWO 1.2 m telescope for each star, and with the Faulkes North Telescope (FTN) of the Las Cumbres Observatory Global Network (LCOGT) at Hawaii for HAT-P-21 only. The observations for each target are summarized in Table 1.

The reduction of these images, including basic calibration, astrometry, and aperture photometry, was performed as described by Bakos et al. (2010). In each case we obtained light curves for all stars in the 2MASS catalog within the field of view of the observations and confirm that the primary target is the source of the transit rather than one of the known neighboring stars. We found that the aperture photometry for HAT-P-20 was significantly affected by the close-by neighbor star 2MASS 07273995+2420118 with Δi = 1.1 mag difference at 6farcs86 separation (Figure 5). Thus, we performed image subtraction on the FLWO 1.2 m images with the same toolset used for the HATNet reductions, but applied a discrete kernel with half-size of 5 pixels and no spatial variations. Indeed, for this stellar configuration, the image subtraction results proved to be superior to the aperture photometry. For all of the follow-up light curves, we performed EPD and TFA to remove trends simultaneously with the light curve modeling (for more details, see Section 3, and Bakos et al. 2010). The final time series, together with our best-fit transit light curve model, are shown in the top portion of Figures 1013; the individual measurements are reported in Tables 811, for HAT-P-20 through HAT-P-23, respectively.

Figure 10.

Figure 10. Unbinned transit light curves for HAT-P-20, acquired with KeplerCam at the FLWO 1.2 m telescope. The light curves have been EPD and TFA processed, as described in Section 3.3. The dates of the events are indicated. Curves after the first are displaced vertically for clarity. Our best fit from the global modeling described in Section 3.3 is shown by the solid lines. Residuals from the fits are displayed at the bottom, in the same order as the top curves. The error bars represent the photon and background shot noise, plus the readout noise.

Standard image High-resolution image
Figure 11.

Figure 11. Similar to Figure 10; here we show the follow-up light curves for HAT-P-21.

Standard image High-resolution image
Figure 12.

Figure 12. Similar to Figure 10; here we show the follow-up light curves for HAT-P-22.

Standard image High-resolution image
Figure 13.

Figure 13. Similar to Figure 10; here we show the follow-up light curves for HAT-P-23.

Standard image High-resolution image

Table 8. High-precision Differential Photometry of HAT-P-20

BJDUTC Maga σMag Mag(orig)b Filter
(2,400,000+)        
54902.59749 −0.00169 0.00075 10.66286 i
54902.59799 −0.00139 0.00075 10.66200 i
54902.59848 −0.00469 0.00075 10.65893 i
54902.59919 0.00281 0.00075 10.66596 i
54902.59970 −0.00108 0.00075 10.66316 i
54902.60020 0.00141 0.00075 10.66524 i
54902.60089 0.00440 0.00075 10.66726 i
54902.60139 0.00040 0.00075 10.66424 i
54902.60190 −0.00064 0.00075 10.66234 i
54902.60259 −0.00009 0.00075 10.66388 i

Notes. aThe out-of-transit level has been subtracted. These magnitudes have been subjected to the EPD and TFA procedures, carried out simultaneously with the transit fit. bRaw magnitude values without application of the EPD and TFA procedures.

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

Table 9. High-precision Differential Photometry of HAT-P-21

BJDUTC Maga σMag Mag(orig)b Filter
(2,400,000+)        
54942.76369 0.01133 0.00088 10.43700 i
54942.76432 0.01370 0.00088 10.43950 i
54942.76495 0.00989 0.00088 10.43550 i
54942.76574 0.00891 0.00088 10.43490 i
54942.76637 0.00974 0.00088 10.43570 i
54942.76698 0.01073 0.00088 10.43650 i
54942.76779 0.00961 0.00088 10.43480 i
54942.76839 0.00938 0.00088 10.43450 i
54942.76902 0.01249 0.00088 10.43800 i
54942.76980 0.01011 0.00088 10.43520 i

Note. Same as those of Table 8.

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

Table 10. High-precision Differential Photometry of HAT-P-22

BJDUTC Maga σMag Mag(orig)b Filter
(2,400,000+)        
54891.60570 0.00028 0.00069 8.78803 z
54891.60603 −0.00100 0.00069 8.78658 z
54891.60636 −0.00189 0.00069 8.78541 z
54891.60667 −0.00138 0.00069 8.78514 z
54891.60702 −0.00237 0.00069 8.78534 z
54891.60735 0.00030 0.00069 8.78923 z
54891.60767 0.00404 0.00069 8.79234 z
54891.60801 −0.00170 0.00069 8.78696 z
54891.60831 0.00257 0.00069 8.79009 z
54891.60865 −0.00145 0.00069 8.78653 z

Notes. Same as those of Table 8.

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

Table 11. High-precision Differential Photometry of HAT-P-23

BJDUTC Maga σMag Mag(orig)b Filter
(2,400,000+)        
54632.73784 0.02143 0.00111 10.98360 i
54632.73870 0.01830 0.00116 10.98170 i
54632.73954 0.01503 0.00115 10.97810 i
54632.74039 0.01859 0.00115 10.98340 i
54632.74125 0.01123 0.00111 10.97530 i
54632.74208 0.02006 0.00109 10.98390 i
54632.74295 0.01831 0.00111 10.98220 i
54632.74381 0.01766 0.00114 10.97840 i
54632.74637 0.02025 0.00117 10.99540 i
54632.74722 0.01556 0.00112 10.97990 i

Notes. Same as those of Table 8.

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

3. ANALYSIS

3.1. Properties of the Parent Stars

Fundamental parameters for each of the host stars, including the mass (M) and radius (R), which are needed to infer the planetary properties, depend strongly on other stellar quantities that can be derived spectroscopically. For this we have relied on our template spectra obtained with the Keck/HIRES instrument, and the analysis package known as Spectroscopy Made Easy (SME; Valenti & Piskunov 1996), along with the atomic line database of Valenti & Fischer (2005). For each star, SME yielded the following initial values and uncertainties (which we have conservatively doubled for the temperature and metallicity to include our estimates of the systematic errors).

  • 1.  
    HAT-P-20—effective temperature Teff⋆ = 4626  ±  104 K, stellar surface gravity log g = 4.80  ±  0.1 (cgs), metallicity [Fe/H] = +0.31  ±  0.08 dex, and projected rotational velocity vsin i = 3.1  ±  0.5 km s−1.
  • 2.  
    HAT-P-21—effective temperature Teff⋆ = 5701  ±  144 K, stellar surface gravity log g = 4.48  ±  0.14 (cgs), metallicity [Fe/H] = +0.04  ±  0.1 dex, and projected rotational velocity vsin i = 4.1  ±  0.5 km s−1.
  • 3.  
    HAT-P-22—effective temperature Teff⋆ = 5338  ±  88 K, stellar surface gravity log g = 4.52  ±  0.14 (cgs), metallicity [Fe/H] = +0.26  ±  0.08 dex, and projected rotational velocity vsin i = 1.5  ±  0.5 km s−1.
  • 4.  
    HAT-P-23—effective temperature Teff⋆ = 5987  ±  120 K, stellar surface gravity log g = 4.48  ±  0.12 (cgs), metallicity [Fe/H] = +0.18  ±  0.1 dex, and projected rotational velocity vsin i = 8.0  ±  0.5 km s−1.

The effective temperature and metallicity, along with the surface gravity taken as a luminosity indicator, can be used as constraints to infer the stellar mass and radius by comparison with stellar evolution models. However, the effect of log g on the spectral line shapes is rather subtle, and as a result it is typically difficult to determine accurately, so that it is a rather poor luminosity indicator in practice. Unfortunately, a trigonometric parallax is not available for any of the host stars, since they were not included among the targets of the Hipparcos mission (Perryman et al. 1997). For planetary transits, another constraint is provided by the a/R normalized semi-major axis, which is closely related to ρ, the mean stellar density. The quantity a/R can be derived directly from the combination of the transit light curves (Sozzetti et al. 2007) and the RV data (required for eccentric cases; see Section 3.3). This, in turn, allows us to improve on the determination of the spectroscopic parameters by supplying an indirect constraint on the weakly determined spectroscopic value of log g that removes degeneracies. We take this approach here, as described below.

For each system, our initial values of Teff⋆, log g, and [Fe/H] were used to determine auxiliary quantities needed in the global modeling of the follow-up photometry and RVs (specifically, the limb-darkening coefficients). This modeling, the details of which are described in Section 3.3, uses a Monte Carlo approach to deliver the numerical probability distribution of a/R and other fitted variables. When combining a/R (used as a proxy for luminosity) with assumed Gaussian distributions for Teff⋆ and [Fe/H] based on the SME determinations, a comparison with stellar evolution models allows the probability distributions of other stellar properties to be inferred, including log g. Here we use the stellar evolution calculations from the Yonsei–Yale group (YY; Yi et al. 2001) for all planets presented in this work. The comparison against the model isochrones was carried out for each of 10,000 Monte Carlo trial sets for HAT-P-21, HAT-P-22, and HAT-P-23, and for 20,000 Monte Carlo trial sets for HAT-P-20 (see Section 3.3). Parameter combinations corresponding to unphysical locations in the H-R diagram (26% of the trials for HAT-P-20, and less than 1% of the trials for the other objects) were ignored, and replaced with another randomly drawn parameter set. For each system we carried out a second SME iteration in which we adopted the value of log g so determined and held it fixed in a new SME analysis (coupled with a new global modeling of the RV and light curves), adjusting only Teff⋆, [Fe/H], and vsin i. This gave the following:

  • 1.  
    HAT-P-20: log g = 4.64  ±  0.06, Teff⋆ = 4595  ±  80 K, [Fe/H] = +0.35  ±  0.08, and vsin i = 2.1  ±  0.5 km s−1.
  • 2.  
    HAT-P-21: log g = 4.31  ±  0.06, Teff⋆ = 5588  ±  80 K, [Fe/H] = +0.01  ±  0.08, and vsin i = 3.5  ±  0.5 km s−1.
  • 3.  
    HAT-P-22: log g = 4.37  ±  0.06, Teff⋆ = 5302  ±  80 K, [Fe/H] = +0.24  ±  0.08, and vsin i = 0.5  ±  0.5 km s−1.
  • 4.  
    HAT-P-23: log g = 4.33  ±  0.06, Teff⋆ = 5905  ±  80 K, [Fe/H] = +0.15  ±  0.04, and vsin i = 8.1  ±  0.5 km s−1.

In each case the conservative uncertainties for Teff⋆ and [Fe/H] have been increased by a factor of two over their formal values, as before. For each system, a further iteration did not change log g significantly, so we adopted the values stated above as the final atmospheric properties of the stars. They are collected in Table 12.

Table 12. Stellar Parameters for HAT-P-20–HAT-P-23

Parameter HAT-P-20 HAT-P-21 HAT-P-22 HAT-P-23 Source
Spectroscopic properties          
Teff⋆ (K)... 4595 ± 80 5588 ± 80 5302 ± 80 5905 ± 80 SMEa
[Fe/H]... +0.35 ± 0.08 +0.01 ± 0.08 +0.24 ± 0.08 +0.15 ± 0.04 SME
vsin i (km s−1)... 2.1 ± 0.5 3.5 ± 0.5 0.5 ± 0.5 8.1 ± 0.5 SME
vmac (km s−1)... 2.21 3.74 3.30 4.22 SME
vmic (km s−1)... 0.85 0.85 0.85 0.85 SME
γRV (km s−1)... −18.81 ± 0.68 −53.19 ± 0.09 +12.49 ± 0.28 −15.10 ± 0.30 DS
Photometric properties          
V (mag)... 11.339 11.685 9.732 12.432 TASSc
VIC (mag)... 1.50 ± 0.12 0.654 ± 0.097 0.992 ± 0.066 0.68 ± 0.12 TASS
J (mag)... 9.276 ± 0.022 10.503 ± 0.022 8.293 ± 0.023 11.103 ± 0.022 2MASS
H (mag)... 8.743 ± 0.021 10.154 ± 0.019 7.935 ± 0.029 10.846 ± 0.022 2MASS
Ks (mag)... 8.601 ± 0.019 10.111 ± 0.018 7.837 ± 0.021 10.791 ± 0.020 2MASS
JK (mag,ESO)... 0.715 ± 0.033 0.419 ± 0.031 0.486 ± 0.034 0.335 ± 0.032 2MASS
Derived properties          
M (M)... 0.756 ± 0.028 0.947 ± 0.042 0.916 ± 0.035 1.130 ± 0.035 YY+a/R+SMEb
R (R)... 0.694 ± 0.021 1.105 ± 0.083 1.040 ± 0.044 1.203 ± 0.074 YY+a/R+SME
log g (cgs)... 4.63 ± 0.02 4.33 ± 0.06 4.36 ± 0.04 4.33 ± 0.05 YY+a/R+SME
L (L)... 0.19 ± 0.02 1.06+0.20− 0.16 0.77 ± 0.09 1.58 ± 0.23 YY+a/R+SME
MV (mag)... 7.07 ± 0.17 4.80 ± 0.19 5.22 ± 0.14 4.31 ± 0.16 YY+a/R+SME
MK (mag,ESO)... 4.42 ± 0.09 3.12 ± 0.16 3.30 ± 0.10 2.86 ± 0.14 YY+a/R+SME
JK (mag,ESO)... 0.66 ± 0.02 0.44 ± 0.02 0.50 ± 0.02 0.37 ± 0.01 YY+a/R+SME
Age (Gyr)... 6.7+5.7− 3.8 10.2 ± 2.5 12.4 ± 2.6 4.0 ± 1.0 YY+a/R+SME
Distance (pc)... 70 ± 3 254 ± 19 82 ± 3 393 ± 25 YY+a/R+SME

Notes. aSME = "Spectroscopy Made Easy" package for the analysis of high-resolution spectra (Valenti & Piskunov 1996). These parameters rely primarily on SME, but have a small dependence also on the iterative analysis incorporating the isochrone search and global modeling of the data, as described in the text. bYY+a/R+SME = Based on the YY isochrones (Yi et al. 2001), a/R as a luminosity indicator, and the SME results. cFrom Droege et al. (2006).

Download table as:  ASCIITypeset image

With the adopted spectroscopic parameters the model isochrones yield the stellar mass and radius, and other properties. These are listed for each of the systems in Table 12. According to these models HAT-P-20 is a dwarf star with an estimated age of 6.7+5.7− 3.8 Gyr, HAT-P-21 is a slightly evolved star with an estimated age of 10.2  ±  2.5 Gyr, HAT-P-22 is a slightly evolved star with an estimated age of 12.4  ±  2.6 Gyr, and HAT-P-23 is a slightly evolved star with an estimated age of 4.0  ±  1.0 Gyr. The inferred location of each star in a diagram of a/R versus Teff⋆, analogous to the classical H-R diagram, is shown in Figure 14. In all cases the stellar properties and their 1σ and 2σ confidence ellipsoides are displayed against the backdrop of model isochrones for a range of ages, and the appropriate stellar metallicity. For comparison, the locations implied by the initial SME results are also shown (in each case with a triangle).

Figure 14.

Figure 14. Upper left: model isochrones from Yi et al. (2001) for the measured metallicity of HAT-P-20, [Fe/H] = +0.35, and ages of 3.0, 4.0, 5.0, 5.5, 6.0, 6.5, 7.0, 7.5, 8.0, 8.5, 9.0, 10.0, and 11.0 Gyr (left to right). The adopted values of Teff⋆ and a/R are shown together with their 1σ and 2σ confidence ellipsoids. The initial values of Teff⋆ and a/R from the first SME and light curve analyses are represented with a triangle. Upper right: same as upper left, here we show the results for HAT-P-21, with [Fe/H] = +0.01, and ages of 5.0, 6.0, 7.0, 8.0, 9.0, 10.0, 11.0, 12.0, and 13.0 Gyr (left to right). Lower left: same as upper left, here we show the results for HAT-P-22, with [Fe/H] = +0.24, and ages of 5.0, 6.0, 7.0, 8.0, 9.0, 10.0, 11.0, 12.0, 13.0, and 14.0 Gyr (left to right). Lower right: same as upper left, here we show the results for HAT-P-23, with [Fe/H] = +0.15, and ages of 0.2, 0.5, 1.0, 1.5, 2.0, 2.5, 3.0, 3.5, 4.0, 4.5, and 5.0 Gyr (left to right).

Standard image High-resolution image

The stellar evolution modeling provides color indices (see Table 12) that may be compared against the measured values as a sanity check. For each star, the best available measurements are the near-infrared magnitudes from the 2MASS Catalog (Skrutskie et al. 2006), which are given in Table 12. These are converted to the photometric system of the models (ESO system) using the transformations by Carpenter (2001). The resulting color indices are also shown in Table 12 for HAT-P-20 through HAT-P-23, respectively. Indeed, the colors from the stellar evolution models and from the observations agree for all of the host stars within 2σ. The distance to each object was computed from the absolute K magnitude from the models and the 2MASS Ks magnitudes, which has the advantage of being less affected by extinction than optical magnitudes. The results are given in Table 12, where in all cases the uncertainty excludes possible systematics in the model isochrones that are difficult to quantify.

3.2. Excluding Blend Scenarios

Our initial spectroscopic analyses discussed in Sections 2.2 and 2.3 rule out the most obvious astrophysical false positive scenarios. However, more subtle phenomena such as blends (contamination by an unresolved eclipsing binary, whether in the background or associated with the target) can still mimic both the photometric and spectroscopic signatures we see. In the following section, we investigate whether such scenarios may have caused the observed photometric and spectroscopic features.

3.2.1. Spectral Line-bisector Analysis

Following Torres et al. (2007), we explored the possibility that the measured RVs are not real, but are instead caused by distortions in the spectral line profiles due to contamination from a nearby unresolved eclipsing binary. A bisector span (BS) analysis for each system based on the Keck spectra was done as described in Section 5 of Bakos et al. (2007). In general, none of the Keck/HIRES spectra suffer significant sky contamination. Nevertheless, we calculated the sky contamination factors (SCF) as described in Hartman et al. (2009), and corrected for the minor correlation between SCF and BS. The results are exhibited in Figure 15, where we show the SCF–BS and RV–BSSCF (BS after SCF correction) plots for each planetary system. We also calculated the Spearman rank-order correlation coefficients (denoted as Rs for the RV versus BS quantities and the false alarm probabilities (see Table 13). There is no correlation for HAT-P-21, HAT-P-22, and HAT-P-23, and thus the interpretation of these systems as transiting planets is clear. There is an anti-correlation present for HAT-P-20, which is strengthened when the SCF correction is applied. A plausible explanation for this is that the neighboring star at 6'' separation (see Figure 5) is bleeding into the slit, even though we were careful during the observations to keep the slit centered on the main target, and adjusted the slit orientation to be perpendicular to the direction to the neighbor. We simulated this scenario, and calculated the expected BS as a function of RV due to the neighbor, assuming that the two stars have the same systemic velocity and the seeing is 1''. Indeed, we get a slight anti-correlation from this simulation, and the range of magnitude in the BS variation is consistent with the observations. It is also possible that some of the anti-correlation is due to the fact that the slit was not in vertical angle for many of the observations. The non-vertical slit mode may result in wavelength-dependent slit losses due to atmospheric dispersion, and this could bring in a correlation with the sky background, and change the shape of the spectral lines.

Figure 15.

Figure 15. Panels on the left show the bisector spans (BS) as a function of sky contamination factor (SCF). Panels on the right exhibit the SCF-corrected BS as a function of the RVs. The individual planets are labeled.

Standard image High-resolution image

Table 13. Summary of RV versus BS Correlations

Name Rs1a FAP1b Rs2c FAP2d
HAT-P-20 −0.73 2.46% −0.92 0.05%
HAT-P-21 −0.24   39% −0.20   47%
HAT-P-22 0.33   30% 0.27   37%
HAT-P-23 0.27   37% 0.27   37%

Notes. aThe Spearman correlation coefficient between the bisector (BS) variations and the radial velocities (RV). bFalse alarm probability for Rs1. cThe Spearman correlation coefficient between BS corrected for the sky contamination factor (SCF) and the RVs. dFalse alarm probability for Rs2.

Download table as:  ASCIITypeset image

3.3. Global Modeling of the Data

This section describes the procedure we followed for each system to model the HATNet photometry, the follow-up photometry, and the RVs simultaneously. Our model for the follow-up light curves used analytic formulae based on Mandel & Agol (2002) for the eclipse of a star by a planet, with limb darkening being prescribed by a quadratic law. The limb-darkening coefficients for the Sloan g-band, Sloan i-band, and Sloan z-band were interpolated from the tables by Claret (2004) for the spectroscopic parameters of each star as determined from the SME analysis (Section 3.1). The transit shape was parameterized by the normalized planetary radius pRp/R, the square of the impact parameter b2, and the reciprocal of the half-duration of the transit ζ/R. We chose these parameters because of their simple geometric meanings and the fact that these show negligible correlations (see Bakos et al. 2010). The relation between ζ/R and the quantity a/R, used in Section 3.1, is given by

Equation (1)

(see, e.g., Tingley & Sackett 2005; Kipping 2010). Note the subtle dependency of a/R on the kecos ω and hesin ω Lagrangian orbital parameters that are typically derived from the RV data (ω is the longitude of periastron). This dependency is often ignored in the literature, and a/R is quoted as a "pure" light curve parameter. Of course, if high-quality secondary eclipse observations are available that determine both the location and duration of the occultation, then k and h can be determined without RV data. Our model for the HATNet data was the simplified "P1P3" version of the Mandel & Agol (2002) analytic functions (an expansion in terms of Legendre polynomials), for the reasons described in Bakos et al. (2010). Following the formalism presented by Pál (2009a), the RVs were fitted with an eccentric Keplerian model parameterized by the semi-amplitude K and Lagrangian elements k and h. Note that we allowed for an eccentric orbit for all planets, even if the results were consistent with a circular orbit. There are several reasons for this: (1) many of the close-in hot Jupiters show eccentric orbits, thus the assumption of fixing e = 0 has no physical justification (while this has been customary in early discoveries relying on very few data points), (2) the error bars on various other derived quantities (including a/R) are more realistic with the inclusion of eccentricity, and (3) non-zero eccentricities can be very important in proper interpretation of these systems.

We assumed that there is a strict periodicity in the individual transit times. For each system we assigned the transit number Ntr = 0 to a complete follow-up light curve. For HAT-P-20b this was the light curve gathered on 2009 October 21, for HAT-P-21b: 2010 February 19, HAT-P-22b: 2009 February 28, and HAT-P-23b: 2008 September 13. The adjustable parameters in the fit that determine the ephemeris were chosen to be the time of the first transit center observed with HATNet (Tc, −252, Tc, −286, Tc, −135, and Tc, −312 for HAT-P-20b through HAT-P-23b, respectively) and that of the last transit center observed with the FLWO 1.2 m telescope (Tc, 0, Tc, 1, Tc, 19, and Tc, 250 for HAT-P-20b through HAT-P-23b, respectively). We used these as opposed to period and reference epoch in order to minimize correlations between parameters (see Pál et al. 2008). Times of mid-transit for intermediate events were interpolated using these two epochs and the corresponding transit number of each event, Ntr. The eight main parameters describing the physical model for each system were thus the first and last transit center times, Rp/R, b2, ζ/R, K, kecos ω, and hesin ω. For HAT-P-20b, HAT-P-22b, and HAT-P-23b, three additional parameters were included (for each system) that have to do with the instrumental configuration (dilution factor, out-of-transit magnitudes, gamma velocities; see later). For HAT-P-21b seven additional parameters were included, because it was observed in three different HATNet fields. These are the HATNet dilution factor Binst (one for each HATNet field), which accounts for dilution of the transit in the HATNet light curve from background stars due to the broad point spread function (PSF; 25'' FWHM) and for reduction of the HATNet transit depth due to the TFA procedure, the HATNet out-of-transit magnitude M0, HATNet (one for each HATNet field), and the relative zero-point γrel of the Keck RVs.

The physical model was extended with an instrumental model that describes brightness variations caused by systematic errors in the measurements. This was done in a similar fashion to the analysis presented by Bakos et al. (2010), i.e., using the "ELTG" (EPD local, TFA global) method for correcting the follow-up light curves. In all cases the total number of fitted parameters (43, 36, 33, and 67 for HAT-P-20b through HAT-P-23b) was much smaller than the number of data points (755, 1172, 892, and 953, counting only RV measurements and follow-up photometry measurements). The joint fit was performed as described in Bakos et al. (2010) using a Markov Chain Monte Carlo approach (MCMC; see Ford 2006).

The a priori distributions of the parameters for these chains were chosen to be Gaussian, with eigenvalues and eigenvectors derived from the Fisher covariance matrix for the best-fit solution. The MCMC analysis provided the full a posteriori probability distributions of all adjusted variables.

Following this procedure we obtained the a posteriori distributions for all other quantities of interest such as a/R. As described in Section 3.1, a/R was used together with stellar evolution models to infer a value for log g that is significantly more accurate than the spectroscopic value. The improved estimate was in turn applied to a second iteration of the SME analysis, as explained previously, in order to obtain better estimates of Teff⋆ and [Fe/H]. The global modeling was then repeated with updated limb-darkening coefficients based on those new spectroscopic determinations. The resulting geometric parameters pertaining to the light curves and velocity curves for each system are listed in Table 14.

Table 14. Orbital and Planetary Parameters for HAT-P-20b–HAT-P-23b

Parameter HAT-P-20b HAT-P-21b HAT-P-22b HAT-P-23b
Light curve parameters        
  P (days) ... 2.875317 ± 0.000004 4.124481 ± 0.000007 3.212220 ± 0.000009 1.212884 ± 0.000002
  Tc (BJDUTC)a ... 2455080.92661 ± 0.00021 2454996.41312 ± 0.00069 2454930.22001 ± 0.00025 2454852.26464 ± 0.00018
  T14 (days)a ... 0.0770 ± 0.0008 0.1530 ± 0.0027 0.1196 ± 0.0014 0.0908 ± 0.0007
  T12 = T34 (days)a ... 0.0137 ± 0.0009 0.0184 ± 0.0029 0.0144 ± 0.0013 0.0105 ± 0.0007
  a/R ... 11.17 ± 0.29 9.60 ± 0.71 8.55 ± 0.35 4.14 ± 0.23
  ζ/R ... 31.32 ± 0.22 14.81 ± 0.15 18.97 ± 0.09 24.90 ± 0.12
  Rp/R ... 0.1284 ± 0.0016 0.0950 ± 0.0022 0.1065 ± 0.0017 0.1169 ± 0.0012
  b2 ... 0.398+0.032− 0.034 0.298+0.089− 0.118 0.217+0.052− 0.065 0.105+0.053− 0.049
  bacos i/R ... 0.631+0.025− 0.028 0.546+0.074− 0.139 0.466+0.052− 0.083 0.324+0.070− 0.101
  i (deg) ... 86.8 ± 0.2 87.2 ± 0.7 86.9+0.6− 0.5 85.1 ± 1.5
Limb-darkening coefficientsb        
  ai (linear term) ... 0.4719 0.2976 0.3587 0.2524
  bi (quadratic term) ... 0.2174 0.3131 0.2831 0.3426
  ag ... ⋅⋅⋅ ⋅⋅⋅ 0.7080 0.5311
  bg ... ⋅⋅⋅ ⋅⋅⋅ 0.1165 0.2539
RV parameters        
  K (m s−1) ... 1246.0 ± 8.1 548.3 ± 14.2 313.3 ± 4.2 368.5 ± 17.6
  kRVc ... 0.012 ± 0.004 0.147 ± 0.011 −0.008 ± 0.008 −0.048 ± 0.023
  hRVc ... −0.007 ± 0.008 −0.175 ± 0.019 0.004 ± 0.016 0.090 ± 0.052
  e ... 0.015 ± 0.005 0.228 ± 0.016 0.016 ± 0.009 0.106 ± 0.044
  ω (deg) ... 317 ± 130 309 ± 3 156 ± 66 118 ± 25
  RV jitter (m s−1) ... 15.7 26.4 9.9 34.7
Secondary eclipse parameters (derived)        
  Ts (BJDUTC) ... 2455082.385 ± 0.007 2454998.865 ± 0.029 2454931.809 ± 0.016 2454852.834 ± 0.018
  Ts, 14 ... 0.0764 ± 0.0010 0.1163 ± 0.0068 0.1204 ± 0.0032 0.1064 ± 0.0095
  Ts, 12 ... 0.0134 ± 0.0009 0.0117 ± 0.0015 0.0145 ± 0.0014 0.0129 ± 0.0020
Planetary parameters        
  Mp (MJ) ... 7.246 ± 0.187 4.063 ± 0.161 2.147 ± 0.061 2.090 ± 0.111
  Rp (RJ) ... 0.867 ± 0.033 1.024 ± 0.092 1.080 ± 0.058 1.368 ± 0.090
  C(Mp, Rp)d ... 0.50 0.28 0.36 0.56
  ρp (g cm−3) ... 13.78 ± 1.50 4.68+1.59− 0.99 2.11+0.40− 0.29 1.01 ± 0.18
  log gp (cgs) ... 4.38 ± 0.03 3.98 ± 0.08 3.66 ± 0.05 3.44 ± 0.05
  a (AU) ... 0.0361 ± 0.0005 0.0494 ± 0.0007 0.0414 ± 0.0005 0.0232 ± 0.0002
  Teq (K) ... 970 ± 23 1283 ± 50 1283 ± 32 2056 ± 66
  Θe ... 0.794 ± 0.031 0.413 ± 0.038 0.179 ± 0.010 0.062 ± 0.004
  Fper (108 erg s−1 cm−2)f ... 2.06 ± 0.20 10.0 ± 1.5 6.33 ± 0.67 50.0 ± 11.4
  Fap (108 erg s−1 cm−2)f ... 1.94 ± 0.19 3.96 ± 0.66 5.91 ± 0.60 32.7 ± 2.7
  〈F〉 (108 erg s−1 cm−2)f ... 2.00 ± 0.19 6.12 ± 0.97 6.12 ± 0.62 40.3 ± 5.3

Notes. aTc: reference epoch of mid-transit that minimizes the correlation with the orbital period. It corresponds to Ntr = −16. BJD is calculated from UTC. T14: total transit duration, time between first to last contact; T12 = T34: ingress/egress time, time between first and second, or third and fourth contact. bValues for a quadratic law, adopted from the tabulations by Claret (2004) according to the spectroscopic (SME) parameters listed in Table 12. ck = ecos ω and h = esin ω. These orbital parameters are derived from the global modeling, and are primarily determined by the RV data. dCorrelation coefficient between the planetary mass Mp and radius Rp. eThe Safronov number is given by $\Theta = \frac{1}{2}(V_{\rm esc}/V_{\rm orb})^2 = (a/R_{p})(M_{p}/ M_\star)$ (see Hansen & Barman 2007). fIncoming flux per unit surface area. 〈F〉 is averaged over the orbit.

Download table as:  ASCIITypeset image

Included in each table is the RV "jitter," which is a noise term that we added in quadrature to the internal errors for the RVs in order to achieve χ2/dof = 1 from the RV data for the global fit. The jitter is a combination of assumed astrophysical noise intrinsic to the star, plus instrumental noise rising from uncorrected instrumental effects (such as a template spectrum taken under suboptimal conditions).

The planetary parameters and their uncertainties were derived by combining the a posteriori distributions for the stellar, light curve, and RV parameters. In this way we find masses and radii for each planet. These and other planetary parameters are listed at the bottom of Table 14, and further discussed in Section 4.

4. DISCUSSION

4.1. HAT-P-20b

HAT-P-20b is a very massive (Mp = 7.246  ±  0.187 MJ = 2302.9  ±  59.5 M) and very compact (Rp = 0.867  ±  0.033 RJ) hot Jupiter orbiting a K3 (Skiff 2009) star. With ρp = 13.78  ±  1.50 g cm−3, HAT-P-20b is among the most massive and most dense transiting planets known (see Figure 16). Modeling HAT-P-20b may be a challenge, as the oldest (4 Gyr, i.e., yielding the most compact planets) Fortney et al. (2007) models with Mp = 2154 M total mass and 100 M core mass predict a much bigger radius (1.04 RJ). The observed radius of 0.87 MJ would require a very high metal content. We note that the host star is one of the most metal-rich stars that have a transiting planet ([Fe/H] = +0.35  ±  0.08). Curiously, HAT-P-20b orbits a fairly late-type star (K3), as compared to most of the massive hot Jupiters that orbit ∼F5 dwarfs. The irradiation HAT-P-20b receives is one of the smallest, clearly making it a pL class exoplanet (Fortney et al. 2008): 〈F〉 = (2.00  ±  0.19) × 108 erg s−1 cm−2, comparable to the mean flux per orbit for another "heavy" planet HD 17156b on a 21 day period orbit. HAT-P-20 is an outlier in the MpM plane; it is a relatively small mass star harboring a very massive planet. Another outlier (albeit to a much lesser extent) with similar planetary radius and stellar mass is WASP-10b (Johnson et al. 2009; Christian et al. 2009), but this planet has less than half of the mass of HAT-P-20b (3.09 MJ). We also calculated the maximum mass of a stable moon for both the prograde and retrograde cases, and derived 0.128 M and 8.31 M, respectively, i.e., HAT-P-20b can harbor a fairly massive moon. An 8.31 M retrograde moon would cause ∼10 s variations in the transit times, which is marginally detectable from the ground.

Figure 16.

Figure 16. Mass–radius diagram of known TEPs (filled circles). HAT-P-20b–HAT-P-23b are shown as filled triangles. Overlaid are Fortney et al. (2007) planetary isochrones interpolated to the solar equivalent semi-major axis of HAT-P-20b for ages of 1.0 Gyr (upper, solid lines) and 4 Gyr (lower, dash-dotted lines) and core masses of 0 and 10 M (upper and lower lines, respectively), as well as isodensity lines for 0.1, 0.3, 0.9, 3.0, 9.0, 25.0, and 100.0 g cm−3 (dotted lines from top to bottom). Solar system planets are shown with filled squares.

Standard image High-resolution image

HAT-P-20 has a close-by faint and red companion star at ∼6farcs86 separation. Based on the Palomar sky survey archival plates, we confirm that they form a close common proper motion pair, thus it is very likely that the two stars are physically associated. The binary has appeared in the Washington Double Star compilation (WDS) as POU2795, and was discovered by Pourteau (1933). Furthermore, based on the summary of observations in the WDS, there is already a hint of orbital motion of the companion to HAT-P-20 over the last century. The position angle of the companion changed from P.A. = 323° to P.A. = 320° over the course of 89 years (between 1909 and 1998), and it seems to be retrograde on the sky (clockwise). Thus, HAT-P-20 is yet another example of a massive planet in a binary system (Udry et al. 2002). The binary companion makes this system ideal for high precision ground or space-based studies, as it provides a natural comparison source, even though it has a later spectral type.

4.2. HAT-P-21b

HAT-P-21b has a mass of Mp = 4.063  ±  0.161 MJ, radius of Rp = 1.024  ±  0.092 RJ, mean density of ρp = 4.68+1.59− 0.99 g cm−3, and orbits on a moderately eccentric orbit with e = 0.228  ±  0.016, ω = 309°  ±  3°. The transits occur near apastron. As noted by Buchhave et al. (2010), 4 MJ mass planets are quite rare in the sample of currently known transiting exoplanets, and the only siblings of HAT-P-21b are WASP-32b (3.6 MJ; Maxted et al. 2010), HD 80606b (4.08 MJ; Naef et al. 2001), and HAT-P-16b (4.19 MJ; Buchhave et al. 2010). WASP-32b and HAT-P-16b are similar to each other with P ∼ 2.7–2.8 day orbital periods and large radii (Rp > 1.15 RJ in both cases), but dissimilar from the longer period, smaller radius, and more eccentric planet HAT-P-21b. HD 80606b, on the other hand, has a similar radius than HAT-P-21b, and orbits on an extremely eccentric (e = 0.93) orbit with a period of 111 days. It appears that HAT-P-21b is thus an unusual, short-period, eccentric, massive, and compact planet.

The only models from Fortney et al. (2007) consistent with the observed radius are 4 Gyr models with 100 M core mass, yielding 1.05 RJ radius. HAT-P-21b has a very high mean density; similar to HD 80606b and WASP-14b.

The flux received by the planet varies between (10.0  ±  1.5) × 108 erg s−1 cm−2 and (3.96  ±  0.66) × 108 erg s−1 cm−2. Interestingly, this puts HAT-P-21b on the bordlerline between pL (low irradiation) and pM (high irradiation) planets. At the time of occultation, HAT-P-21b is just approaching its periastron, thus entering the irradiation level quoted for pM-type planets.

4.3. HAT-P-22b

HAT-P-22b has a mass of Mp  =  2.147  ±  0.061 MJ, radius of Rp  =  1.080  ±  0.058 RJ, and mean density of ρp = 2.11+0.40− 0.29 g cm−3. HAT-P-22b orbits a fairly metal-rich ([Fe/H]  =  + 0.24  ±  0.08), bright (V = 9.732), and close-by (82  ±  3 pc) star. Similar to HAT-P-20, the host star has a faint and red neighbor at 9'' separation that is co-moving with HAT-P-22 (based on the POSS plates and recent Keck/HIRES snapshots), thus they are likely to form a physical pair.

HAT-P-22b belongs to the moderately massive (∼2 MJ) and compact (Rp ≈ 1 RJ) hot Jupiters, such as HAT-P-15b (Mp = 1.95MJ, Rp = 1.07 RJ; Kovács et al. 2010), HAT-P-14b (Mp = 2.23MJ, Rp = 1.15 RJ; Torres et al. 2010), and WASP-8b (Mp = 2.25MJ, Rp = 1.05 RJ; Queloz et al. 2010).

HAT-P-22b is broadly consistent with the models of Fortney et al. (2008). For 300 Myr, 1 Gyr, and 4 Gyr models, it requires a 100 M, 50 M, and 25 M core, respectively, to have a radius of ∼1.08 RJ. The low incoming flux (see Table 14) means that HAT-P-22b is a pL class planet. HAT-P-22b can harbor a 0.96 M mass retrograde moon, which would cause transit timing variations (TTVs) of ∼2 s.

4.4. HAT-P-23b

HAT-P-23b has a mass of Mp = 2.090  ±  0.111 MJ, radius of Rp = 1.368  ±  0.090 RJ, and mean density of ρp = 1.01  ±  0.18 g cm−3. It belongs to the inflated group of 2 MJ planets. A few other planets which are similar to it include TrES-3b (Mp = 1.91MJ, Rp = 1.37 RJ; Sozzetti et al. 2009), Kepler-5b (Mp = 2.10MJ, Rp = 1.31 RJ; Kipping & Bakos 2011; Koch et al. 2010), WASP-46b (Mp = 2.10MJ, Rp = 1.33 RJ; Anderson et al. 2011), and CoRoT-11b (Mp = 2.33MJ, Rp = 1.43 RJ; Gandolfi et al. 2010). The orbit is nearly circular, with the eccentricity being marginally significant. The reason for the somewhat higher than usual errors in the RV parameters is the high jitter of the star (34.7 m s−1), which may be related to the moderately high vsin i = 8.1  ±  0.5 km s−1 and the very close-in orbit of HAT-P-23b. The Fortney et al. (2008) models cannot reproduce the observed radius of HAT-P-23b; even for the youngest, (300 Myr) coreless models, the theoretical radius for its mass is 1.25 RJ. HAT-P-23b orbits its host star on a very close-in orbit. The orbital period is only 1.2129 days; almost identical to that of OGLE-TR-56b (1.21192 days). The nominal planetary radius of the two objects is also the same within 1%, but OGLE-TR-56b is much less massive (1.39 MJ). The flux falling on HAT-P-23b from its host star is one of the highest (i.e., belongs to the pM class objects), and is similar to that of HAT-P-7b and OGLE-TR-56b. We also calculated the spiral infall timescale for each new discovery based on Levrard et al. (2009) and Dobbs-Dixon et al. (2004). By assuming that the stellar dissipation factor is Q = 106, the infall time for HAT-P-23b is τinfall = 7.5+2.9− 1.8 Myr, one of the shortest among exoplanets.

The Rossiter–McLaughlin (R-M) effect for HAT-P-23b should be quite significant, given the moderately high vsin i = 8.1  ±  0.5 km s−1 of the host star and the Δi = 17 mmag deep transit. The impact parameter is also "ideal" (b = 0.324+0.070− 0.101), i.e., it is not equatorial (b = 0), where there would be a strong degeneracy between the stellar rotational velocity vsin i and the sky-projected angle of the stellar spin axis and the orbital normal, λ, and is also far from grazing, where the transit is short, and other system parameters have lower accuracy. The effective temperature of the star (Teff⋆ = 5905  ±  80 K) is close to the critical temperature of 6250 K noted recently by Winn et al. (2010), which may be a borderline between systems where the stellar spin axes and planetary orbital normals are preferentially aligned (Teff⋆ < 6250 K) and those that are misaligned (Teff⋆ > 6250 K). An alternative hypothesis has been brought up by Schlaufman (2010), where misaligned stellar spin axes and orbital normals are related to the mass of the host star. The mass of HAT-P-23 (1.13  ±  0.04) is sufficiently close to the suggested dividing line of M = 1.2 M, and thus it will provide an excellent additional test for these ideas.17

4.5. Summary

We presented the discovery of four new massive transiting planets, and provided accurate characterization of the host star and planetary parameters. These four new systems are very diverse, and significantly expand the sample of massive (Mp ≳ 2 MJ) planets. Two of the new discoveries orbit stars that have fainter companions, which are probably physically associated. The new discoveries do not tend to enhance the mass–eccentricity correlation, since only one (HAT-P-21b) is significantly eccentric. Also, the tentative mass–vsin i correlation noted in Section 1 is weakened by the new discoveries. The heavier mass planets (HAT-P-20b and HAT-P-21b) are inconsistent with current theoretical models in that they are extremely dense, and would require a substantial core (or high metal content) to have such small radii. One planet (HAT-P-23b) is also inconsistent with the models (unless we assume that the planet is very young), but in the sense that it has a density that is lower than can be explained by the models. It has been noted by Winn et al. (2010) and Schlaufman (2010) that systems exhibiting stellar spin axis–planetary orbital normal misalignment are preferentially eccentric and heavy mass planets (in addition to the dependence noted by these authors of the spin–orbit angle on the effective temperature or mass of the host star). The four new planets presented in this work will provide additional important tests for checking these conjectures. The host stars are all bright (9.7 < V < 12.4), and thus enable in-depth future characterization of these systems.

HATNet operations have been funded by NASA grants NNG04GN74G, NNX08AF23G, and SAO IR&D grants. Work of G.Á.B. was supported by the Postdoctoral Fellowship of the NSF Astronomy and Astrophysics Program (AST-0702843 and AST-0702821, respectively). G.T. acknowledges partial support from NASA grant NNX09AF59G. We acknowledge partial support also from the Kepler Mission under NASA Cooperative Agreement NCC2-1390 (D.W.L., PI). G.K. thanks the Hungarian Scientific Research Foundation (OTKA) for support through grant K-81373. This research has made use of Keck telescope time granted through NOAO and NASA, and uses observations obtained with facilities of the Las Cumbres Observatory Global Telescope. The authors recognize and acknowledge the very significant cultural role and reverence that the summit of Mauna Kea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain.

Footnotes

  • Based in part on observations obtained at the W. M. Keck Observatory, which is operated by the University of California and the California Institute of Technology. Keck time has been granted by NOAO and NASA.

  • 13 

    This includes those four planets announced in this paper.

  • 14 

    See, e.g., http://exoplanets.org for statistics.

  • 15 

    Note that the recent definition adopted by the exoplanet encyclopedia is 25 MJ (Schneider et al. 2011).

  • 16 

    Since the submission of this paper, the sample of massive transiting planets has further increased to ∼30.

  • 17 

    Following the submission of this paper, Moutou et al. (2011) reported a detection of the R-M effect for HAT-P-23b. They find an amplitude of ∼50 m s−1 for this effect, and measure a projected spin–orbit angle of +15°  ±  22° for this system.

Please wait… references are loading.
10.1088/0004-637X/742/2/116