Articles

HAT-P-39b–HAT-P-41b: THREE HIGHLY INFLATED TRANSITING HOT JUPITERS*

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

Published 2012 October 9 © 2012. The American Astronomical Society. All rights reserved.
, , Citation J. D. Hartman et al 2012 AJ 144 139 DOI 10.1088/0004-6256/144/5/139

1538-3881/144/5/139

ABSTRACT

We report the discovery of three new transiting extrasolar planets orbiting moderately bright (V = 11.1, 11.7, and 12.4) F stars. The planets HAT-P-39b through HAT-P-41b have periods of P = 3.5439 days, 4.4572 days, and 2.6940 days, masses of 0.60 MJ, 0.62 MJ, and 0.80 MJ, and radii of 1.57 RJ, 1.73 RJ, and 1.68 RJ, respectively. They orbit stars with masses of 1.40 M, 1.51 M, and 1.51 M, respectively. The three planets are members of an emerging population of highly inflated Jupiters with 0.4 MJ < M < 1.5 MJ and R > 1.5 RJ.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Transiting exoplanets (TEPs) are key objects for the study of planets outside the solar system. The geometry of these planetary systems enables measurements of several important physical parameters, such as planetary masses and radii, or the sky-projected angle between the orbital axis of a planet and the spin axis of its host star (e.g., Queloz et al. 2000). The vast majority of TEPs have been discovered by dedicated photometric surveys, including the Wide Angle Search for Planets (WASP; Pollacco et al. 2006), the Hungarian-made Automated Telescope Network (HATNet; Bakos et al. 2004) and its southern extension (HATSouth; Bakos et al. 2012), Kepler (Borucki et al. 2010), CoRoT (Barge et al. 2008), OGLE (Udalski et al. 2002), TrES (Alonso et al. 2004), XO (McCullough et al. 2006), the Qatar Exoplanet Survey (QES; Alsubai et al. 2011), the Kilodegree Extremely Little Telescope survey (KELT; Siverd et al. 2012), and MEarth (Charbonneau et al. 2009).

Significant among these are the ground-based, wide-field surveys using small aperture telescopes, including WASP, HATNet, HATSouth, TrES, XO, QES, and KELT. While these surveys are heavily biased toward discovering large planets on short-period orbits compared to the Kepler and CoRoT space-based surveys, the planets discovered by ground-based surveys tend to orbit stars that are brighter than those discovered by the space-based surveys, making such planets more amenable to detailed characterization and follow-up studies (this is true as well for the few, but valuable, transiting planets discovered by radial velocity (RV) searches, which are found around even brighter stars than those discovered by photometric surveys). Additionally the extreme environments in which these planets are discovered, while perhaps not representative of most planetary systems, create a natural experiment for testing theories of planet structure and formation. For example, a number of gas-giant planets have been discovered with radii that are substantially larger than theoretically expected (e.g., Mandushev et al. 2007; Collier Cameron et al. 2007; Snellen et al. 2009; Hebb et al. 2009; Latham et al. 2010; Fortney et al. 2011; Anderson et al. 2010, 2011; Enoch et al. 2011; Hartman et al. 2011; Smalley et al. 2012). These have been used to empirically determine the factors affecting the radii of planets (e.g., Enoch et al. 2012), which in turn informs theoretical work on the subject.

In this paper, we present the discovery and characterization of three new transiting planets around the relatively bright stars GSC 1364-01424, GSC 3607-01028, and GSC 0488-02442, by the HATNet survey. As members of the growing sample of highly inflated planets, these objects will provide valuable leverage for understanding the physics that determines the structure of planets.

In Section 2, we summarize the detection of the photometric transit signal and the subsequent spectroscopic and photometric observations of each star to confirm the planets. In Section 3, we analyze the data to rule out false positive scenarios, and to determine the stellar and planetary parameters. Our findings are briefly discussed in Section 4.

2. OBSERVATIONS

The observational procedure employed by HATNet to discover TEPs has been described in detail in several previous discovery papers (e.g., Bakos et al. 2010; Latham et al. 2009). In the following subsections, we highlight specific details of the procedure that are relevant to the discoveries of HAT-P-39b through HAT-P-41b.

2.1. Photometric Detection

Table 1 summarizes the photometric observations of each new planetary system, including the discovery observations made with the HATNet system. The HATNet images were processed and reduced to trend-filtered light curves following the procedure described by Bakos et al. (2010). The light curves were searched for periodic box-shaped signals using the box least-squares (BLS; see Kovács et al. 2002) method. We detected significant signals in the light curves of the stars summarized below (see Figure 1).

  • 1.  
    HAT-P-39GSC 1364-01424 (also known as Two Micron All Sky Survey (2MASS) 07350197+1749482; $\alpha = 07^{\mathrm h}35^{\mathrm m}01\mbox{$.\!\!^{\mathrm s}$}97$, δ = +17°49'48farcs3; J2000; V = 12.422; Droege et al. 2006). A signal was detected for this star with an apparent depth of ∼10.9 mmag and a period of P = 3.5439 days.
  • 2.  
    HAT-P-40GSC 3607-01028 (also known as 2MASS 22220308+4527265; $\alpha = 22^{\mathrm h}22^{\mathrm m}03\mbox{$.\!\!^{\mathrm s}$}00$, δ = +45°27'26farcs6; J2000; V = 11.699; Droege et al. 2006). A signal was detected for this star with an apparent depth of ∼4.4 mmag and a period of P = 4.4572 days.
  • 3.  
    HAT-P-41GSC 0488-02442 (also known as 2MASS 19491743+0440207; $\alpha = 19^{\mathrm h}49^{\mathrm m}17\mbox{$.\!\!^{\mathrm s}$}40$, δ = +04°40'20farcs7; J2000; V = 11.087; Droege et al. 2006). A signal was detected for this star with an apparent depth of ∼8.4 mmag and a period of P = 2.6940 days.
Figure 1.

Figure 1. HATNet light curves of HAT-P-39 (top), HAT-P-40 (middle), and HAT-P-41 (bottom) phase-folded with the transit period. In each case we show two panels: the top shows the unbinned light curve, while the bottom shows the region zoomed-in on the transit, with dark filled circles for the light curve binned in-phase with a bin size of 0.002. The solid line shows the model fit to the light curve. A total of 14, 12, and 17 individual transits were observed for HAT-P-39b through HAT-P-41b, respectively.

Standard image High-resolution image

2.2. Reconnaissance Spectroscopy

High-resolution, low-signal-to-noise (S/N) "reconnaissance" spectra were obtained for HAT-P-39, HAT-P-40, and HAT-P-41 using the Harvard-Smithsonian Center for Astrophysics Digital Speedometer (DS; Latham 1992) until it was retired in 2009, and thereafter the Tillinghast Reflector Echelle Spectrograph (TRES; Fűrész 2008), both on the 1.5 m Tillinghast Reflector at the Fred Lawrence Whipple Observatory (FLWO) in AZ. The reconnaissance spectroscopic observations and results for each system are summarized in Table 2. The DS observations were reduced and analyzed following the procedure described by Torres et al. (2002), while the TRES observations were reduced and analyzed following the procedure described by Quinn et al. (2012) and Buchhave et al. (2010).

Table 1. Summary of Photometric Observations

Instrument/Field Date(s) Number of Images Mode Cadence Filter
      (minutes)  
HAT-P-39        
   HAT-7/G267 2007 Dec–2008 May 800        5.5 R band
   HAT-8/G267 2007 Oct–2008 May 1850        5.5 R band
   HAT-6/G315 2007 Dec–2008 May 4700        5.5 R band
   HAT-9/G315 2007 Oct–2008 May 2800        5.5 R band
   KeplerCam 2009 Mar 30 257        0.89 Sloan i band
   KeplerCam 2009 Apr 6 204        0.89 Sloan i band
   KeplerCam 2009 Dec 17 373        0.89 Sloan i band
   KeplerCam 2011 Feb 12 240        1.0 Sloan i band
HAT-P-40        
   HAT-5/G159 2004 Sep–2006 Jan 4403        5.5 I band
   HAT-8/G159 2005 Nov–2006 Jan 1591        5.5 I band
   HAT-9/G159 2004 Dec–2005 Jan 411        5.5 I band
   HAT-11/G159 2004 Dec–2005 Jan 331        5.5 I band
   KeplerCam 2010 Sep 16 371        1.40 Sloan i band
   FTN 2010 Oct 17 241        0.50 Sloan i band
   FTN 2011 Aug 16 285        0.85 Sloan i band
   FTN 2011 Aug 25 325        0.94 Sloan i band
   FTN 2011 Oct 4 177        0.94 Sloan i band
   BOS 2011 Oct 13 381        1.17 Sloan i band
HAT-P-41        
   HAT-6/G388 2009 May–2009 Jul 176        5.5 Sloan r band
   HAT-8/G388 2009 May–2009 Sep 3380        5.5 Sloan r band
   KeplerCam 2010 May 27 226        0.98 Sloan i band
   KeplerCam 2010 Jun 23 438        0.49 Sloan i band
   KeplerCam 2011 May 31 307        0.73 Sloan i band
   KeplerCam 2011 Jun 19 653        0.48 Sloan i band
   BOS 2011 Jul 24 238        1.17 Sloan i band

Download table as:  ASCIITypeset image

Table 2. Summary of Reconnaissance Spectroscopy Observations

Instrument HJD − 2, 400, 000 Teff⋆ log g vsin i γRVa
    (K) (cgs) (km s−1) (km s−1)
HAT-P-39          
   DS 4 obs 54807–54931 6250 4.0 16 28.54 ± 0.58 (rms)
   TRES 54934.6546 6500 4.0 16 29.25
HAT-P-40          
   TRES 55084.8821 6110 ± 80 4.21 ± 0.13 8.7 ± 0.5 −25.97
   TRES 55131.6811 5940 ± 170 4.04 ± 0.26 10.8 ± 1.5 −25.67
   TRES 55138.6609 5910 ± 170 3.81 ± 0.27 12.4 ± 1.6 −26.08
   TRES 55162.6713 6020 ± 50 3.93 ± 0.10 8.0 ± 0.5 −25.65
   TRES 55168.5775 6120 ± 80 4.15 ± 0.14 8.5 ± 0.5 −25.58
HAT-P-41          
   TRES 55319.9727 6504 ± 100 4.3 ± 0.16 23.9 ± 0.7 32.32
   TRES 55372.9209 5807 ± 223 3.94 ± 0.35 32.1 ± 2.5 29.84
   TRES 55373.9046 6430 ± 105 4.28 ± 0.17 27.5 ± 0.7 30.44

Note. aThe mean heliocentric RV of the target in the IAU system, with a systematic uncertainty of approximately 0.1 km s−1, mostly limited by how well the velocities of the standard stars have been established. We give the mean and rms RV for the four DS observations of HAT-P-39, while the velocity and classification for each TRES observation of HAT-P-39 through HAT-P-41 are listed individually.

Download table as:  ASCIITypeset image

Based on the observations summarized in Table 2, we find that all three systems have RV root mean square (rms) residuals consistent with no detectable RV variation within the precision of the measurements. All spectra were single-lined, i.e., there is no evidence that any of these targets consist of more than one star. Note that while there is a close companion to HAT-P-41 (Section 2.5), it was resolved by the TRES guider and the light from the companion did not go down the fiber. The gravities for all of the stars indicate that none of the stars are giants, though HAT-P-40 may be slightly evolved.

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. The observations were made with HIRES (Vogt et al. 1994) on the Keck-I telescope in HI and with FIES on the Nordic Optical Telescope on the island of La Palma, Spain (Djupvik & Andersen 2010). We used the high-resolution fiber (providing spectra with a resolution of R = 67, 000) for four of the FIES observations, and the medium-resolution fiber (R = 46, 000) for five of the FIES observations. The HIRES observations were reduced to RVs in the barycentric frame following the procedure described by Butler et al. (1996), while the FIES observations were reduced following Buchhave et al. (2010). The RV measurements and uncertainties are given in Tables 35 for HAT-P-39 through HAT-P-41, respectively. The period-folded data along with our best fit, described below in Section 3, are displayed in Figures 24.

Figure 2.

Figure 2. Top panel: Keck/HIRES RV measurements for HAT-P-39 shown as a function of orbital phase, along with our best-fit circular model (solid line; see Table 11) and our best-fit eccentric model (dashed line). Zero phase corresponds to the time of mid-transit. The center-of-mass velocity has been subtracted. Second panel: velocity OC residuals from the best fit. The error bars include a component from astrophysical jitter (43.0 m s−1) added in quadrature to the formal errors (see Section 3.3). Third panel: bisector spans (BSs), with the mean value subtracted. The measurement from the template spectrum is included. The BS uncertainties are internal errors determined for each spectrum from the scatter of the individual BS values measured on separate orders of the spectrum; they do not include the unknown contribution from stellar jitter. Fourth panel: FWHM of the cross-correlation functions computed from the blue regions of the Keck/HIRES spectra, with the mean value subtracted. Bottom panel: chromospheric activity index S. Note the different vertical scales of the panels. Observations shown twice are represented with open symbols.

Standard image High-resolution image
Figure 3.

Figure 3. Keck/HIRES observations of HAT-P-40. The panels are as in Figure 2. The parameters used in the best-fit model are given in Table 11.

Standard image High-resolution image
Figure 4.

Figure 4. Keck/HIRES and FIES/NOT observations of HAT-P-41. The panels are as in Figure 2. Here, we use filled circles to show Keck/HIRES observations, filled triangles to show FIES/NOT observations with the high-resolution fiber, and open triangles to show FIES/NOT observations with the low-resolution fiber. BS and S index measurements are only available for the Keck/HIRES observations. The parameters used in the best-fit model are given in Table 11.

Standard image High-resolution image

Table 3. Relative Radial Velocities, Bisector Spans, and Activity Index Measurements of HAT-P-39

BJDa RVb σRVc BS σBS Sd Phase
(2,454,000+) (m s−1) (m s−1) (m s−1) (m s−1)    
954.81362 −51.56 7.81 ... ... 0.180 0.345
955.84224 32.48 8.82 ... ... 0.171 0.635
956.82792 61.50 9.51 ... ... 0.175 0.913
1107.13759 −90.39 8.88 −4.63 4.91 0.197 0.327
1109.07573 78.38 9.05 −13.59 7.57 0.183 0.874
1112.12944 42.24 7.88 −4.74 5.19 0.176 0.736
1135.07184 −39.87 6.92 −2.64 4.96 0.178 0.210
1172.95852 −41.57 7.30 19.77 4.58 0.176 0.900
1187.99450 −28.34 8.69 8.74 5.51 ... 0.143
1190.10225 70.06 7.12 30.09 4.51 0.181 0.738
1192.03996 −146.19 6.26 16.58 6.61 0.181 0.285
1192.05453 ... ... 16.75 3.03 0.180 0.289
1193.05376 −31.04 6.96 −1.15 4.07 0.179 0.571
1193.88084 90.74 7.48 −36.37 5.72 0.178 0.804
1197.00264 89.62 7.37 ... ... 0.180 0.685
1198.91700 4.34 10.58 ... ... 0.184 0.225
1250.89571 129.15 7.46 −8.86 4.93 0.183 0.892
1251.90337 −98.98 7.98 −6.43 4.96 0.179 0.177
1289.75983 76.11 7.50 −17.19 6.90 0.186 0.859
1312.75610 −14.73 7.40 −17.48 6.40 0.182 0.348
1313.79897 54.76 8.78 2.00 4.22 0.177 0.642
1466.10653 114.14 7.74 2.68 3.75 0.180 0.620
1468.09041 −14.54 6.61 0.53 4.41 0.183 0.180
1470.11601 −11.17 8.91 41.44 5.79 0.174 0.751
1545.08385 27.02 8.10 16.59 4.67 0.179 0.906
1545.86194 −53.41 8.31 1.59 5.47 0.178 0.125
1611.99999 24.36 16.44 −10.80 22.01 0.160 0.788
1698.75509 −64.08 9.64 −22.52 12.11 0.167 0.268
1699.76307 16.59 8.07 −10.35 7.63 0.175 0.553

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 Date calculated directly from UTC, without correction for leap seconds. 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. dChromospheric activity index.

Download table as:  ASCIITypeset image

Table 4. Relative Radial Velocities, Bisector Spans, and Activity Index Measurements of HAT-P-40

BJDa RVb σRVc BS σBS Sd Phase
(2,454,000+) (m s−1) (m s−1) (m s−1) (m s−1)    
1191.81591 18.28 3.81 −2.42 4.85 ... 0.595
1192.73706 46.75 2.96 −1.85 3.29 0.148 0.802
1192.74599 ... ... 4.16 7.06 0.146 0.804
1464.91615 49.92 3.24 −1.75 2.12 0.145 0.867
1465.96736 −41.21 3.02 2.71 4.09 0.144 0.102
1467.92695 17.59 2.93 −9.07 6.74 0.147 0.542
1468.99826 57.77 3.09 2.18 2.51 0.145 0.782
1469.73453 28.31 3.45 −11.76 3.57 0.141 0.948
1486.91164 60.46 3.33 −2.67 6.82 0.145 0.801
1521.81745 35.22 3.96 23.93 6.53 0.143 0.633
1698.11353 −49.94 3.80 4.39 3.60 0.146 0.185
1699.11678 −31.14 3.84 2.58 4.61 0.150 0.410
1701.07726 46.71 3.15 −10.44 4.24 0.145 0.850
1853.75742 −38.93 2.73 ... ... 0.148 0.105

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 Date calculated directly from UTC, without correction for leap seconds. 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. dChromospheric activity index.

Download table as:  ASCIITypeset image

Table 5. Relative Radial Velocities, Bisector Spans, and Activity Index Measurements of HAT-P-41

BJDa RVb σRVc BS σBS Sd Phase Instrumente
(2,454,000+) (m s−1) (m s−1) (m s−1) (m s−1)      
1375.97050 92.63 9.38 25.18 6.00 0.155 0.546 Keck
1375.97664 ... ... 20.47 5.29 0.155 0.549 Keck
1378.12394 −78.22 9.40 0.84 7.92 0.148 0.346 Keck
1378.58766 52.31 40.10 ... ... ... 0.518 FIESm
1379.06917 132.24 9.90 32.40 7.80 0.153 0.697 Keck
1379.54578 21.11 51.30 ... ... ... 0.874 FIESm
1380.47420 −106.99 49.00 ... ... ... 0.218 FIESm
1381.10708 −39.66 10.23 −1.39 4.57 0.150 0.453 Keck
1381.61687 112.71 49.90 ... ... ... 0.642 FIESm
1383.49357 −79.79 75.00 ... ... ... 0.339 FIESm
1400.84669 49.04 12.32 −60.12 7.44 0.139 0.780 Keck
1404.80595 −52.53 10.18 −32.58 7.48 0.148 0.250 Keck
1427.51377 112.18 40.10 ... ... ... 0.679 FIESh
1428.54090 −22.42 91.50 ... ... ... 0.060 FIESh
1429.38087 −79.42 50.00 ... ... ... 0.372 FIESh
1431.54013 −153.82 68.60 ... ... ... 0.173 FIESh
1465.87835 5.42 11.42 −9.93 6.17 0.145 0.919 Keck
1467.87264 65.94 10.34 −11.95 6.68 0.147 0.659 Keck
1469.89585 −68.65 10.18 17.06 4.94 0.143 0.410 Keck
1490.75265 −113.04 10.38 9.02 9.83 0.145 0.152 Keck
1500.73246 24.76 9.14 12.08 8.78 0.143 0.857 Keck
1704.09441 −67.20 10.27 −1.09 5.26 0.155 0.342 Keck
1814.94017 16.90 10.72 ... ... 0.148 0.487 Keck

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 Date calculated directly from UTC, without correction for leap seconds. 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 are considered in Section 3.3. dChromospheric activity index. eWe separately indicate observations obtained with FIES using the medium-resolution fiber and observations obtained with FIES using the high-resolution fiber.

Download table as:  ASCIITypeset image

In each figure, we also show the spectral-line bisector spans (BSs) computed from the Keck/HIRES spectra following Torres et al. (2007), the FWHM of the Keck/HIRES spectral lines (computed from the cross-correlation function (CCF), in a similar manner to the BSs), and the S activity index calculated following Isaacson & Fischer (2010).

2.4. Photometric Follow-up Observations

We conducted additional photometric observations of the three stars with the KeplerCam CCD camera on the FLWO 1.2 m telescope, the Spectral CCD on the 2.0 m Faulkes Telescope North (FTN) at Haleakala Observatory in HI, and the CCD imager on the Byrne Observatory at Sedgwick (BOS) 0.8 m telescope, at Sedgwick Reserve in the Santa Ynez Valley, CA. Both FTN and BOS are operated by the Las Cumbres Observatory Global Telescope (LCOGT16; T. M. Brown et al. 2012, in preparation). The observations for each target are summarized in Table 1.

The reduction of the KeplerCam images to light curves was performed as described by Bakos et al. (2010). The FTN and BOS images were reduced in a similar manner. We performed an external parameter decorrelation (EPD) and a trend filtering algorithm (TFA) to remove trends simultaneously with the light-curve modeling (for more details, see Bakos et al. 2010). The final time series, together with our best-fit transit light curvelight-curve model, are shown in the top portion of Figures 57, while the individual measurements are reported in Tables 68.

Figure 5.

Figure 5. Unbinned transit light curves for HAT-P-39, 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 6.

Figure 6. Similar to Figure 5; here we show the follow-up light curves for HAT-P-40. The facility used for each light curve is indicated next to the date of the event.

Standard image High-resolution image
Figure 7.

Figure 7. Similar to Figure 5; here we show the follow-up light curves for HAT-P-41. The facility used for each light curve is indicated next to the date of the event.

Standard image High-resolution image

Table 6. High-precision Differential Photometry of HAT-P-39

BJDa Magb σMag Mag (Orig)c Filter
(2,400,000+)        
54921.62065 0.00694 0.00127 11.10550 i
54921.62127 0.00597 0.00127 11.10530 i
54921.62187 0.00631 0.00126 11.10640 i
54921.62248 0.00358 0.00127 11.10490 i
54921.62311 0.00600 0.00127 11.10610 i
54921.62390 0.00762 0.00127 11.10910 i
54921.62452 0.00787 0.00126 11.10670 i
54921.62512 0.00663 0.00126 11.10860 i
54921.62574 0.01035 0.00126 11.10940 i
54921.62637 0.01050 0.00126 11.11030 i

Notes. aBarycentric Julian Date calculated directly from UTC, without correction for leap seconds. bThe out-of-transit level has been subtracted. These magnitudes have been subjected to the EPD and TFA procedures, which were carried out simultaneously with the transit fit. cRaw 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. Machine-readable and Virtual Observatory (VO) versions of the full table are available.

Download table as:  Machine-readable (MRT)Virtual Observatory (VOT)Typeset image

Table 7. High-precision Differential Photometry of HAT-P-40

BJDa Magb σMag Mag (Orig)c Filter
(2,400,000+)        
55456.60581 0.00871 0.00080 10.22060 i
55456.60739 0.00713 0.00067 10.21900 i
55456.60830 0.00854 0.00066 10.22070 i
55456.60930 0.00851 0.00066 10.22100 i
55456.61028 0.00935 0.00065 10.22120 i
55456.61124 0.00789 0.00065 10.21980 i
55456.61224 0.00868 0.00065 10.22070 i
55456.61319 0.00695 0.00065 10.21870 i
55456.61416 0.00727 0.00065 10.21880 i
55456.61513 0.00996 0.00064 10.22200 i

Notes. aBarycentric Julian Date calculated directly from UTC, without correction for leap seconds. bThe out-of-transit level has been subtracted. These magnitudes have been subjected to the EPD and TFA procedures, which were carried out simultaneously with the transit fit. cRaw 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. Machine-readable and Virtual Observatory (VO) versions of the full table are available.

Download table as:  Machine-readable (MRT)Virtual Observatory (VOT)Typeset image

Table 8. High-precision Differential Photometry of HAT-P-41

BJDa Magb σMag Mag (Orig)c Filter
(2,400,000+)        
55344.79616 0.01161 0.00077 10.01060 i
55344.79679 0.00980 0.00078 10.00870 i
55344.79744 0.00586 0.00078 10.00480 i
55344.79878 0.00847 0.00078 10.00830 i
55344.79946 0.01379 0.00078 10.01310 i
55344.80015 0.01168 0.00078 10.01140 i
55344.80083 0.00987 0.00077 10.00820 i
55344.80151 0.01021 0.00077 10.00830 i
55344.80219 0.01031 0.00076 10.00870 i
55344.80288 0.01025 0.00077 10.00800 i

Notes. aBarycentric Julian Date calculated directly from UTC, without correction for leap seconds. bThe out-of-transit level has been subtracted. These magnitudes have been subjected to the EPD and TFA procedures, which were carried out simultaneously with the transit fit. cRaw 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. Machine-readable and Virtual Observatory (VO) versions of the full table are available.

Download table as:  Machine-readable (MRT)Virtual Observatory (VOT)Typeset image

2.5. Adaptive Optics Imaging

We obtained high-resolution imaging of HAT-P-41 on the night of 2011 June 21 using the Clio2 near-IR imager on the MMT 6.5 m telescope in AZ. Observations were obtained with the adaptive optics (AO) system in H band and in L' band. Figure 8 shows the resulting H-band image of HAT-P-41, which easily resolves the 3farcs56 ± 0farcs02 neighbor.

Figure 8.

Figure 8. H-band AO image of HAT-P-41 obtained with MMT/Clio2 showing the ∼3farcs5 neighbor to the south.

Standard image High-resolution image

Based on these observations, we measure the H- and L'-band magnitudes of the neighbor relative to HAT-P-41 to be ΔH = 2.46 ± 0.06 mag and ΔL' = 2.6 ± 2.0 mag. Assuming that the star is a physical companion to HAT-P-41, these magnitude differences are consistent with the neighbor being a ∼0.7 M K dwarf, or roughly ∼3.5 mag fainter than HAT-P-41 in the i band.

The KeplerCam observations of HAT-P-41 described in Section 2.4 show HAT-P-41 to have an elongation in the wing of its point-spread function (PSF) due to the companion, but the seeing is not good enough to resolve the stars given the magnitude difference. We use the daophot and allstar programs (Stetson 1987, 1992) to obtain PSF-fitting photometry for the two stars and find that the neighbor is fainter in the i band by Δi ≈ 3.3 mag, with an uncertainty of at least 0.1 mag. We conclude that the broadband photometry is consistent with HAT-P-41 having a K dwarf binary companion. At the distance of HAT-P-41, the 3farcs5 angular separation corresponds to a projected physical separation of ∼1000 AU between the two stars.

3. ANALYSIS

3.1. Properties of the Parent Star

We measured the stellar atmospheric parameters for each star using the Keck/HIRES iodine-free template spectra, together with the Spectroscopy Made Easy (SME; Valenti & Piskunov 1996) package, and the Valenti & Fischer (2005) atomic line database. For each star, we obtained the following initial values and uncertainties.

  • 1.  
    HAT-P-39. Effective temperature Teff⋆ = 6325 ± 100 K, metallicity [Fe/H] = 0.14 ± 0.1 dex, stellar surface gravity log g = 4.04 ± 0.1 (cgs), and projected rotational velocity vsin i = 12.7 ± 0.5 km s−1.
  • 2.  
    HAT-P-40. Effective temperature Teff⋆ = 6140 ± 100 K, metallicity [Fe/H] = 0.25 ± 0.1 dex, stellar surface gravity log g = 4.04 ± 0.1 (cgs), and projected rotational velocity vsin i = 6.7 ± 0.5 km s−1.
  • 3.  
    HAT-P-41. Effective temperature Teff⋆ = 6007 ± 100 K, metallicity [Fe/H] = 0.06 ± 0.1 dex, stellar surface gravity log g = 3.68 ± 0.06 (cgs), and projected rotational velocity vsin i = 20.6 ± 0.5 km s−1.

As described in our previous papers (e.g., Bakos et al. 2010), these initial values were used to determine the quadratic limb-darkening coefficients for each star from the Claret (2004) tables. We then used the mean stellar density, determined from the normalized semimajor axis a/R, together with the effective temperature and metallicity to determine an initial estimate of the mass and radius of each star from the Yonsei–Yale (YY) isochrones (Yi et al. 2001). This provided a refined estimate of the stellar surface gravity, which we fixed in a second iteration of SME for each star. For each system, a third iteration did not change log g appreciably, so we adopted the values from the second iteration as the final spectroscopic parameters for each star. These parameters are listed in Table 9. In this same table, we also list the available broadband photometric magnitudes from the literature and the physical parameters, such as the stellar masses and radii, which are determined from the spectroscopic parameters together with the model isochrones. As discussed in Section 3.3, we adopt the parameters assuming a circular orbit for each planet. Some of the parameters, especially the derived stellar masses and radii, depend on the eccentricity; Table 10 lists the values for these parameters when the eccentricity is allowed to vary.

Table 9. Adopted Stellar Parameters for HAT-P-39–HAT-P-41 Assuming Circular Orbits

Parameter HAT-P-39 HAT-P-40 HAT-P-41 Source
  Value Value Value  
Spectroscopic properties        
   Teff⋆ (K) 6430 ± 100 6080 ± 100 6390 ± 100 SMEa
   [Fe/H] 0.19 ± 0.10 0.22 ± 0.10 0.21 ± 0.10 SME
   vsin i (km s−1) 12.7 ± 0.5 6.9 ± 0.5 19.6 ± 0.5 SME
   vmac (km s−1) 5.04 4.50 4.97 SME
   vmic (km s−1) 0.85 0.85 0.85 SME
   γRV (km s−1) 28.42 ± 0.28 −25.0 ± 0.1 31.68 ± 0.61 DS or TRES
Photometric properties        
   V (mag) 12.422 11.699 11.087 TASS
   VIC (mag) 0.58 ± 0.15 0.77 ± 0.12 0.63 ± 0.10 TASS
   J (mag) 11.424 ± 0.020 10.367 ± 0.023 10.006 ± 0.027 2MASS
   H (mag) 11.184 ± 0.022 10.085 ± 0.018 9.777 ± 0.032 2MASS
   Ks (mag) 11.157 ± 0.020 10.009 ± 0.018 9.728 ± 0.029 2MASS
Derived properties        
   M (M) 1.404 ± 0.051 1.512+0.045− 0.109 1.418 ± 0.047 YY+a/R+SMEb
   R (R) 1.625+0.081− 0.062 2.206 ± 0.061 1.683+0.058− 0.036 YY+a/R+SME
   log g (cgs) 4.16 ± 0.03 3.93 ± 0.02 4.14 ± 0.02 YY+a/R+SME
   L (L) 4.05 ± 0.48 6.00 ± 0.61 4.25 ± 0.41 YY+a/R+SME
   MV (mag) 3.21 ± 0.14 2.83 ± 0.13 3.16 ± 0.12 YY+a/R+SME
   MK (mag,ESO) 2.12 ± 0.10 1.49 ± 0.07 2.04 ± 0.06 YY+a/R+SME
   Age (Gyr) 2.0 ± 0.4 2.7+0.9− 0.3 2.2 ± 0.4 YY+a/R+SME
   AV (mag)c 0.171 ± 0.135 0.353 ± 0.127 0.248 ± 0.134 YY+a/R+SME
   Distance (pc) 642 ± 29 501 ± 16 344+12− 8 YY+a/R+SME
   log R'HKd −4.85 ± 0.07 −5.12 ± 0.16 −5.04 ± 0.04 Keck/HIRES

Notes. aSME = "Spectroscopy Made Easy" package for the analysis of high-resolution spectra (Valenti & Piskunov 1996). These parameters rely primarily on SME, but also have a small dependence 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. cV-band extinction determined by comparing the measured 2MASS and TASS photometry for the star to the expected magnitudes from the YY+a/R+SME model for the star. We use the Cardelli et al. (1989) extinction law. dChromospheric activity index defined in Noyes et al. (1984) determined from the Keck/HIRES spectra following Isaacson & Fischer (2010). In each case, we give the average value and the standard deviation from the individual spectra.

Download table as:  ASCIITypeset image

Table 10. Derived Stellar Parameters for HAT-P-39–HAT-P-41 Allowing Eccentric Orbitsa

Parameter HAT-P-39 HAT-P-40 HAT-P-41 Source
  Value Value Value  
M (M) 1.400+0.102− 0.069 1.504 ± 0.103 1.405 ± 0.066 YY+a/R+SME
R (R) 1.622+0.294− 0.167 2.422 ± 0.154 1.525+0.177− 0.133 YY+a/R+SME
log g (cgs) 4.16 ± 0.10 3.85 ± 0.04 4.22 ± 0.07 YY+a/R+SME
L (L) 4.01+1.78− 0.81 6.98+1.27− 0.97 3.68+1.00− 0.68 YY+a/R+SME
MV (mag) 3.22 ± 0.31 2.67 ± 0.18 3.31 ± 0.25 YY+a/R+SME
MK (mag, ESO) 2.12 ± 0.29 1.31 ± 0.14 2.25 ± 0.22 YY+a/R+SME
Age (Gyr) 2.0+0.5− 0.6 2.9+0.8− 0.6 1.5 ± 0.6 YY+a/R+SME
AV (mag) 0.156+0.152− 0.111 0.317 ± 0.126 0.332 ± 0.134 YY+a/R+SME
Distance (pc) 641+115− 66 548 ± 36 311+36− 27 YY+a/R+SME

Note. aQuantities and abbreviations are as in Table 9, which gives our adopted values, determined assuming circular orbits. We do not list parameters that are independent of the eccentricity.

Download table as:  ASCIITypeset image

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 9. In each case, the stellar properties and their 1σ and 2σ confidence ellipsoids 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). Note the significant change in the estimated temperature, particularly for HAT-P-41, between the first and second SME iterations. As shown recently by Torres et al. (2012), when SME is applied in the iterative fashion described in this paper to TEP host stars, the temperature after the second iteration tends to be systematically higher than after the first iteration for stars with Teff ≳ 6200 K. This is presumably due to hot stars having fewer and broader lines than cool stars, making it difficult to simultaneously infer the surface gravity, temperature, and metallicity from their spectra. When a star hosts a transiting planet, the strong constraint on the stellar density, which comes from the transit, makes it possible to reduce the number of free parameters in modeling the spectrum, yielding a more accurate determination of those parameters.

Figure 9.

Figure 9. Model isochrones from Yi et al. (2001) for the metallicities of HAT-P-39 (top), HAT-P-40 (center), and HAT-P-41 (bottom). For HAT-P-39 and HAT-P-41, the isochrones are shown for ages of 0.2 Gyr, 0.6 Gyr, and 1.0 Gyr–3.0 Gyr in steps of 0.5 Gyr (left to right), while for HAT-P-40 the isochrones are shown for ages of 0.2 Gyr, 0.6 Gyr, and 1.0 Gyr–4.5 Gyr in steps of 0.5 Gyr (left to right). The adopted values of Teff⋆ and a/R are shown together with their 1σ and 2σ confidence ellipsoids. In each plot, the initial values of Teff⋆ and a/R from the first SME and light-curve analyses are represented with a triangle.

Standard image High-resolution image

We determine the distance to and extinction of each star by comparing the J, H, and KS magnitudes from the Two Micron All Sky Survey (2MASS) catalog (Skrutskie et al. 2006), and the V and IC magnitudes from the TASS Mark IV catalog (Droege et al. 2006), to the expected magnitudes from the stellar models. We use the transformations by Carpenter (2001) to convert the 2MASS magnitudes to the photometric system of the models (ESO), and use the Cardelli et al. (1989) extinction law, assuming a total-to-selective extinction ratio of RV = 3.1, to relate the extinction in each bandpass to the V-band extinction AV. The resulting AV and distance measurements are given in Table 9. We find total V-band extinctions of AV = 0.171 ± 0.135, 0.353 ± 0.127, and 0.248 ± 0.134 mag for HAT-P-39 through HAT-P-41, respectively. For comparison, the total line-of-sight extinctions in each direction, estimated from the Schlegel et al. (1998) dust maps, are 0.146 mag, 0.724 mag, and 0.513 mag. Following Bonifacio et al. (2000), we estimate the expected distance-corrected extinction to each source to be 0.113 mag, 0.306 mag, and 0.157 mag, respectively. For HAT-P-39 and HAT-P-40, the measured and expected values are consistent. For HAT-P-41, the broadband photometry appears to point to a slightly redder star than expected based on the spectroscopic temperature and expected extinction. As noted in Section 2.5, HAT-P-41 has a close companion that is unresolved in the 2MASS or TASS catalogs. This companion is the probable cause of the discrepancy between the expected and observed magnitudes of HAT-P-41.

As discussed in Section 3.3, both HAT-P-39 and HAT-P-41 show high RV jitter, with values of 43.0 m s−1 and 33.4 m s−1, respectively. While these values are higher than for most exoplanet host stars, they are typical of F dwarfs with projected rotation velocities greater than 10 km s−1. Using the empirical relation between vsin i and jitter found by Saar et al. (2003), the expected jitter for an F dwarf with vsin i = 10 km s−1 is ∼30 m s−1. As discussed in Hartman et al. (2011) in the context of HAT-P-32 and HAT-P-33, two rapidly rotating F dwarfs with jitter values in excess of those seen for HAT-P-39 and HAT-P-41, the physical origin of this jitter is not clear, though Saar et al. (1998) argue that for F dwarfs it is probably due to convective inhomogeneities on the stellar surfaces that vary in time.

3.2. Excluding Blend Scenarios

The analyses of our reconnaissance spectroscopic observations, discussed in Section 2.2, rule out the most obvious astrophysical false positive scenarios for HAT-P-39 through HAT-P-41. Additionally, the spectral-line BS analyses that we conducted (Figures 24) provide constraints on more subtle blend scenarios similar to that presented in Torres et al. (2004). However, because HAT-P-39 and HAT-P-41 have high RV jitter, and consequently, high BS scatter (∼80 m s−1 and ∼30 m s−1, respectively), relative to their RV semiamplitudes (63.6 ± 10.4 m s−1 and 92.5 ± 11.6 m s−1, respectively), the BS test provides a less stringent constraint on possible blend scenarios than it does in a case such as HAT-P-40 (BS scatter of ∼10 m s−1 and RV semiamplitude of 58.1 ± 2.9 m s−1). To provide additional support for the planetary interpretation of the observations of each system, we conduct detailed blend analyses of the light curves (including both the HATNet discovery light curves and all available follow-up light curves) and absolute photometry following Hartman et al. (2011).

In Figure 10 we show, for each system, the histogram of Δχ2 values between the best-fit transiting planet model and the best-fit blend model for simulated data sets having the same noise properties as the observed residuals from the best-fit blend model (see Hartman et al. 2011 for a more detailed discussion). In this same figure, we also show the Δχ2 difference between the best-fit models applied to the observations. We find that for HAT-P-39 and HAT-P-41, we can reject blend scenarios involving combinations of three stars with greater than 3σ and 5σ confidence, respectively, based solely on the photometry. For HAT-P-39, the detailed shape of the transit, as determined from the follow-up light curves, contributes most of the χ2 difference between the models, while for HAT-P-41 it is the lack of out-of-transit variations, as determined from the HATNet light curve, that contributes most of the χ2 difference. For HAT-P-40, we are unable to rule out blend scenarios based solely on the photometry however; in this case, the lack of BS or FWHM variations rules out such blends. To quantify this, we simulate the CCF of blended systems, which could plausibly fit the photometric data (configurations which cannot be rejected with greater than 5σ confidence) and find that in all cases either the RV or FWHM of the blended configuration varies by several km s−1, or the BS varies by greater than ∼100 m s−1, greatly exceeding the observational limits on any such variations. Similarly for HAT-P-39, we find that stellar blend configurations which cannot be rejected with greater than 5σ confidence predict greater than 500 m s−1 variations in the RV or BS of the Keck spectra, which are well above the observational constraints.

Figure 10.

Figure 10. Histogram of Δχ2 values between the best-fit transiting planet model and the best-fit blend model for simulated data sets having the same noise properties as the observed residuals from the best-fit blend model (see Hartman et al. 2011 for a more detailed discussion). The histogram is shown separately for each system. In each panel, the arrow shows the χ2 difference between the models fitted to the actual observations. For HAT-P-41, the best-fit blend scenario is rejected based on the photometry with greater than 5σ confidence. For HAT-P-39 and HAT-P-40, blend scenarios which cannot be rejected based solely on the photometry are rejected based on spectroscopic information (limits on variations in RV, and in the BS and FWHM of the spectral-line profiles).

Standard image High-resolution image

As discussed in Section 2.5, HAT-P-41 has a close neighbor that we estimate to be ∼3.5 mag fainter in the i band. While such a neighbor could, in principle, be eclipsed by an object that would produce a ∼1% dip in the blended light curve, as we have shown here, the detailed shape of the light curve cannot be produced using physically possible combinations of stars (i.e., stars with parameters determined from stellar evolution models). We note that the KeplerCam observations show no evidence for variations in the flux centroid that correlate with the photometry, providing further evidence that the observed variation is not due to a deep eclipse in the poorly resolved neighbor.

While we can rule out, for each of the systems, blend scenarios involving only stellar-mass components, we cannot rule out scenarios involving binary star systems with one component hosting a transiting planet. Indeed, for HAT-P-41 we find that including a ∼0.7 M star in the system provides a slightly better fit to the photometric observations. In this case, we actually know that there is a faint companion (though it is either unresolved in our light curves or in the available absolute broadband photometry measurements), so it is reassuring to find that the blend analysis of the photometric data also points to the existence of this companion. For HAT-P-39, we can rule out binary companion stars with M < 1.24 M, while for HAT-P-40 and HAT-P-41, we cannot rule out binary companions of any mass up to that of the mass inferred for the brighter star in the system. While massive binary companions in general should be easier to detect in the spectrum, if the two stars have very similar average velocities, the resulting variations in BS, FWHM, and RV measurements can be less than the constraints set by the observations.

We conclude that each system presented here contains a transiting planet; however, we cannot definitively claim that these are all single stars. High-resolution imaging, or further high-precision RVs would be needed to rule out, or discover, binary star companions. There is no evidence that either HAT-P-39 or HAT-P-40 is a binary system, so we treat each of these as single stars in the analysis that follows. For HAT-P-41, there is a resolved neighbor that we account for in our analysis of the system.

3.3. Global Modeling of the Data

We modeled the HATNet photometry, the follow-up photometry, and the high-precision RV measurements using the procedure described by Bakos et al. (2010). Following the discussion by Eastman et al. (2012), we made two important changes to our analysis procedure compared to what was done in Bakos et al. (2010). As noted in Section 3.5.3 of Eastman et al. (2012), there is a common mistake in the implementation of the Metropolis–Hastings (M-H) algorithm for conducting a Markov Chain Monte Carlo (MCMC) analysis whereby the Markov Chain is not increased when a proposed transition is rejected. We discovered that the implementation we have been using for the analysis of HATNet planets has made this mistake and we have corrected it for the analysis of the planets presented in this paper. We found that this bug tends to inflate the error bars on the determined parameters by a factor of a few parts in a hundred. Errors given in previous discovery papers may thus be slightly overestimated. The second significant change that we have made is to use $\sqrt{e} \cos \omega$ and $\sqrt{e} \sin \omega$ as jump parameters, rather than ecos ω and esin ω. Previously, we had been assuming uniform priors on the latter jump parameters, which amounts to assuming a linear prior on the eccentricity e, creating a bias toward measuring nonzero eccentricities. As other authors have noted, using $\sqrt{e} \cos \omega$ and $\sqrt{e} \sin \omega$ leads to a uniform prior on e. We found that this change had a much more significant impact on our determined parameters than did correcting the bug in our implementation of M-H. For example, for HAT-P-39b, the eccentricity that we find is 0.094 ± 0.086 compared with 0.161 ± 0.094 when using our old jump parameters.

We also made a few minor changes to the analysis for the particular planets presented in this paper. For the analysis of HAT-P-41, we allowed independent RV zero points for the Keck/HIRES RVs, the high-resolution NOT/FIES RVs, and the medium-resolution NOT/FIES RVs. To account for the contribution from the neighbor to the photometry, we also fixed the third light in the i band to 4%, based on our PSF-fitting analysis of the KeplerCam observations of this system.

For each of the planets, we added in quadrature a component of stellar jitter to the formal Keck/HIRES RV errors such that χ2 per degree of freedom is unity. For HAT-P-41, we did not add a jitter term to the FIES/NOT RV errors because the formal errors for these observations exceeded the scatter in the RV residuals. We find that both HAT-P-39 and HAT-P-41 require relatively high jitter values (43.0 m s−1 and 33.4 m s−1, respectively), while HAT-P-40 requires significantly less jitter (6.2 m s−1). To see whether the excess RV scatter for HAT-P-39 or HAT-P-41 could be due to additional planets in these systems, we examined the Lomb–Scargle frequency spectra (Lomb 1976; Scargle 1982; Press & Rybicki 1989) of the RV residuals of both stars and found no significant peaks. As discussed in Section 3.1, these values are consistent with the jitter seen in other stars with comparable spectral types and projected rotation velocities. Note that despite the significant RV scatter, the orbital variation is detected at high confidence for all three planets. Quantitatively, for HAT-P-39 the orbital variation is detected with ∼6σ confidence, for HAT-P-40 it is detected with ∼20σ confidence, and for HAT-P-41 it is detected with ∼8σ confidence.

For each planet, we performed the fit both allowing eccentricity to vary and fixing it to zero. The resulting parameters for each system are listed in Table 11, assuming circular orbits, and in Table 12 allowing eccentric orbits. We use a Lucy & Sweeney (1971) test to determine the significance of the measured eccentricities of each system. We find that the observations of all three systems are consistent with the planets being on circular orbits (the circular orbit hypothesis is rejected with 45% confidence, 86% confidence, and 91% confidence, for HAT-P-39b, HAT-P-40b, and HAT-P-41b, respectively; for reference the confidence level would need to be greater than 99.7% for the detection of a nonzero eccentricity to be significant at the 3σ level, which is generally taken as a minimum level of significance). In the past, we have generally presented parameters for systems allowing the eccentricity to vary, even in cases where the observations are consistent with a circular orbit, on the grounds that this provides a more conservative estimate of the errors. However, as discussed recently by Anderson et al. (2012), the best-fit parameters that result from allowing the eccentricity to vary are often biased relative to the circular orbit values, and in most cases further follow-up observations, such as occultation observations, reveal the planets to be on circular orbits after all. We therefore suggest adopting the circular orbit parameters as the most probable values for each planet. These are given in Table 11. For reference, Table 12 lists the resulting parameters when the eccentricities are varied.

Table 11. Adopted Orbital and Planetary Parameters for HAT-P-39b–HAT-P-41b Assuming Circular Orbits

Parameter HAT-P-39b HAT-P-40b HAT-P-41b
  Value Value Value
Light-curve parameters      
   P (days) 3.543870 ± 0.000005 4.457243 ± 0.000010 2.694047 ± 0.000004
   Tc (BJD)a 2455208.75049 ± 0.00041 2455813.17584 ± 0.00054 2454983.86167 ± 0.00107
   T14 (days)a 0.1745 ± 0.0017 0.2557 ± 0.0014 0.1704 ± 0.0012
   T12 = T34 (days)a 0.0178 ± 0.0017 0.0196 ± 0.0009 0.0166 ± 0.0009
   a/R 6.74 ± 0.25 5.92+0.06− 0.14 5.44+0.09− 0.15
   ζ/Rb 12.75 ± 0.07 8.48 ± 0.04 13.01 ± 0.06
   Rp/R 0.0993 ± 0.0025 0.0807 ± 0.0014 0.1028 ± 0.0016
   b2 0.122+0.069− 0.061 0.030+0.045− 0.019 0.049+0.051− 0.029
   bacos i/R 0.349+0.085− 0.120 0.174+0.091− 0.077 0.222+0.088− 0.093
   i (deg) 87.0 ± 1.0 88.3 ± 0.9 87.7 ± 1.0
Limb-darkening coefficientsc      
   c1, i (linear term) 0.1787 0.2190 0.1908
   c2, i (quadratic term) 0.3812 0.3650 0.3746
   c1, r ⋅⋅⋅ ⋅⋅⋅ 0.2658
   c2, r ⋅⋅⋅ ⋅⋅⋅ 0.3814
   c1, I ⋅⋅⋅ 0.1987 ⋅⋅⋅
   c2, I ⋅⋅⋅ 0.3646 ⋅⋅⋅
   c1, R 0.2335 ⋅⋅⋅ ⋅⋅⋅
   c2, R 0.3871 ⋅⋅⋅ ⋅⋅⋅
RV parameters      
   K (m s−1) 63.6 ± 10.4 58.1 ± 2.9 92.5 ± 11.6
   e 0 (fixed) 0 (fixed) 0 (fixed)
   RV jitter (m s−1)d 43.0 6.2 33.4
Planetary parameters      
   Mp (MJ) 0.599 ± 0.099 0.615 ± 0.038 0.800 ± 0.102
   Rp (RJ) 1.571+0.108− 0.081 1.730 ± 0.062 1.685+0.076− 0.051
   C(Mp, Rp)e 0.09 0.36 0.10
   ρp (g cm−3) 0.19 ± 0.04 0.15 ± 0.01 0.20 ± 0.03
   log gp (cgs) 2.77 ± 0.09 2.71 ± 0.03 2.84 ± 0.06
   a (AU) 0.0509 ± 0.0006 0.0608+0.0006− 0.0015 0.0426 ± 0.0005
   Teq (K)f 1752 ± 43 1770 ± 33 1941 ± 38
   Θg 0.027 ± 0.005 0.029 ± 0.002 0.028 ± 0.004
   〈F〉 (109 erg s−1 cm−2)h 2.13 ± 0.21 2.22 ± 0.17 3.20 ± 0.25

Notes. aReported times are in Barycentric Julian Date calculated directly from UTC, without correction for leap seconds. Tc: reference epoch of mid-transit that minimizes the correlation with the orbital period. 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. bReciprocal of the half duration of the transit used as a jump parameter in our MCMC analysis in place of a/R. It is related to a/R by the expression $\zeta /R_\star = a/R_\star (2\pi (1+e\sin \omega))/(P \sqrt{1 - b^{2}}\sqrt{1-e^{2}})$ (Bakos et al. 2010). cValues for a quadratic law, adopted from the tabulations by Claret (2004) according to the spectroscopic (SME) parameters listed in Table 9. dError term, either astrophysical or instrumental in origin, added in quadrature to the formal Keck/HIRES RV errors such that χ2 per degree of freedom is unity. For HAT-P-41, we did not add a jitter term to the FIES/NOT RV errors because the formal errors for these observations exceeded the scatter in the RV residuals. eCorrelation coefficient between the planetary mass Mp and radius Rp. fPlanet equilibrium temperature averaged over the orbit, calculated assuming a Bond albedo of zero, and that flux is reradiated from the full planet surface. gThe Safronov number is given by Θ = (1/2)(Vesc/Vorb)2 = (a/Rp)(Mp/M) (see Hansen & Barman 2007). hIncoming flux per unit surface area, averaged over the orbit.

Download table as:  ASCIITypeset image

Table 12. Orbital and Planetary Parameters for HAT-P-39b–HAT-P-41b Allowing Eccentric Orbitsa

Parameter HAT-P-39b HAT-P-40b HAT-P-41b
  Value Value Value
Light-curve parameters      
   a/R 6.74 ± 0.77 5.39 ± 0.28 5.97 ± 0.52
   ζ/R 12.76 ± 0.08 8.49 ± 0.04 12.99 ± 0.06
   i (deg) 87.1+1.1− 1.5 88.0+0.9− 1.2 87.9 ± 0.9
RV parameters      
   K (m s−1) 63.8 ± 10.6 55.4 ± 2.5 95.6 ± 10.4
   $\sqrt{e} \cos \omega$ 0.094 ± 0.189 0.041 ± 0.064 −0.189 ± 0.100
   $\sqrt{e} \sin \omega$ −0.008 ± 0.263 0.300+0.074− 0.137 −0.293+0.244− 0.120
   ecos ω 0.022 ± 0.074 0.012 ± 0.018 −0.068 ± 0.039
   esin ω −0.001 ± 0.114 0.092 ± 0.051 −0.105 ± 0.086
   e 0.094 ± 0.086 0.095 ± 0.048 0.139 ± 0.063
   ω (deg) 188 ± 115 83 ± 42 237 ± 38
   RV jitter (m s−1) 42.7 4.1 27.5
Secondary eclipse parameters      
   Ts (BJD) 2455189.310 ± 0.168 2455819.895 ± 0.052 2454985.092 ± 0.068
   Ts, 14 0.1741 ± 0.0378 0.3048 ± 0.0300 0.1397 ± 0.0241
   Ts, 12 0.0176 ± 0.0084 0.0239 ± 0.0029 0.0135 ± 0.0027
Planetary parameters      
   Mp (MJ) 0.596 ± 0.099 0.584 ± 0.035 0.812 ± 0.094
   Rp (RJ) 1.565+0.292− 0.169 1.900 ± 0.127 1.529+0.185− 0.136
   C(Mp, Rp) 0.08 0.03 0.32
   ρp (g cm−3) 0.19+0.10− 0.06 0.11 ± 0.02 0.28+0.09− 0.07
   log gp (cgs) 2.77 ± 0.14 2.60 ± 0.06 2.93 ± 0.09
   a (AU) 0.0509+0.0012− 0.0009 0.0607 ± 0.0014 0.0424 ± 0.0007
   Teq (K) 1751+137− 87 1843 ± 60 1879 ± 88
   Θ 0.027 ± 0.006 0.025 ± 0.002 0.032 ± 0.004
   〈F〉 (109 erg s−1 cm−2) 2.12+0.85− 0.38 2.60 ± 0.34 2.82+0.65− 0.44

Note. aQuantities and definitions are as in Table 11, which gives our adopted values, determined assuming circular orbits. Here, we do not list parameters that are effectively independent of the eccentricity.

Download table as:  ASCIITypeset image

4. DISCUSSION

We have presented the discovery of three new transiting planets which we show on mass–radius and equilibrium temperature–radius diagrams in Figure 11. As seen in the mass–radius diagram, planets generally have radii with 0.6 RJ < R < 1.5 RJ over a broad mass range spanning over two orders of magnitude, except for in the range 0.4 MJ < M < 1.5 MJ, where planets are found with radii as large as ∼2.0 RJ (if WASP-12b is excluded, then the mass range is 0.4 MJ < M < 1.0 MJ). Applying the Kolmogorov–Smirnov test (e.g., Press et al. 1992), we find that there is only a 0.7% chance that the masses of planets with M > 0.4 MJ and R > 1.5 RJ are drawn from the same distribution as the masses of planets with M > 0.4 MJ and R < 1.5 RJ. The three planets presented here fall in the population of large radius (R > 1.5 RJ), sub-Jupiter-mass planets, which we refer to as highly inflated planets.

Figure 11.

Figure 11. Left: mass–radius diagram of TEPs. HAT-P-39b through HAT-P-41b are indicated with triangles. Jupiter and Saturn are marked with filled squares. Right: equilibrium temperature vs. radius; the mass of each planet is indicated by the color of its symbol.

Standard image High-resolution image

As has been repeatedly noted (the earliest reference being Guillot 2005), the radii of close-in gas-giant planets are strongly correlated with the degree of irradiation (variously traced by the planet equilibrium temperature estimated by adopting a constant albedo, typically zero, for all planets, and making an assumption about the heat redistribution, or traced by the bolometric surface flux). As is evident in Figure 11, the degree to which planets are inflated depends on their masses, with lower mass planets showing a stronger correlation between temperature and radius. This has also been previously noted (e.g., Enoch et al. 2012) and has been taken as evidence for some theoretical models of the inflation process (e.g., Batygin et al. 2011; Laughlin et al. 2011). The planets presented here generally follow the established empirical trends, though they are somewhat more inflated than other planets with comparable equilibrium temperatures and semimajor axes. The empirical relation given by Enoch et al. (2012), which gives a prediction for the radius as a function of Teq and a for planets with 0.5 MJ < Mp < 2.0 MJ, yields radii of 1.52 RJ, 1.63 RJ, and 1.58 RJ for HAT-P-39b through HAT-P-41b, respectively. The formula given by Béky et al. (2011), which uses Teq and [Fe/H] as independent variables and was derived for planets with 0.3 MJ < Mp < 0.8 MJ, predicts radii of 1.31 RJ, 1.30 RJ, and 1.37 RJ. In all cases, the predicted radii are smaller than the measured values (1.57 RJ, 1.73 RJ, and 1.68 RJ), though the Enoch et al. (2012) relation, which was determined including more recent discoveries, gives values that are closer to the observations.

All three of the planets presented in this paper orbit relatively massive stars, with 1.4 M < M < 1.52 M. The existence of hot Jupiters around such stars, as revealed by transit surveys, can be contrasted with the results from Doppler surveys targeting evolved stars with m⋆ > 1.5 M. With two exceptions (Johnson et al. 2010; Sato et al. 2012), such stars show a strong deficit of planets on orbits smaller than 1 AU (Johnson et al. 2007). It is unclear whether this deficit is due to the destruction of close-in planets as their host stars evolve to the sub-giant branch, or if it is a relic of the planet formation process (see Bowler et al. 2010 and references therein). While HAT-P-39–HAT-P-41 have masses near or below the 1.5 M threshold adopted for the Doppler surveys, the discovery of these close-in planets, and others around stars with M > 1.4 M (e.g., HAT-P-7, HAT-P-33, Kepler-14, KOI-428, OGLE2-TR-L9, TrES-4, WASP-33, and WASP-79) suggests that the absence of close-in planets around evolved massive stars is most likely due to planet destruction, rather than due to a relic of the planet formation process.

We find that HAT-P-41 has a neighbor which has near-IR photometry consistent with it being a 0.7 M star at the same distance as HAT-P-41. HAT-P-41 is thus one of number of hot Jupiter host stars with close visually resolved neighbors (e.g., HD 189733, Bakos et al. 2006; HAT-P-1, Bakos et al. 2007, and many others), though it is not known what fraction of these are physical companions. Such companions may be responsible for driving planets into close-in orbits via the Kozai Mechanism (Fabrycky & Tremaine 2007), a hypothesis that has gained traction recently with the discovery that many close-in planets are on high obliquity orbits (e.g., Triaud et al. 2010; Albrecht et al. 2012), but a complete survey to determine the frequency of companions to hot Jupiter host stars in a statistically robust way has not yet been published.

Finally, we note that all three of these planets are good targets for measuring the Rossiter–McLaughlin effect (R-M; Rossiter 1924; McLaughlin 1924) as they orbit bright stars with relatively high-projected rotation velocities. For HAT-P-39b and HAT-P-41b, the expected R-M semiamplitude is over 100 m s−1. For HAT-P-40b, the signal amplitude is lower, but this is compensated by the long duration of the transit. The three stars are also positioned closely below (HAT-P-40) and above (HAT-P-39 and HAT-P-41) the 6250 K effective temperature threshold found by Winn et al. (2010) to separate planets on orbits that are well aligned with the spin axes of their host stars from planets that are on high obliquity orbits (see also Albrecht et al. 2012).

HATNet operations have been funded by NASA grants NNG04GN74G, NNX08AF23G, and SAO IR&D grants. J.H. acknowledges partial support from NSF grant AST-1108686. G.Á.B., Z.C., and K.P. acknowledge partial support from NASA grant NNX09AB29G. G.T. acknowledges partial support from NASA grant NNX09AF59G. We acknowledge partial support also from the Kepler mission under NASA Cooperative Agreement NCC2-1390 (PI: D.W.L.). 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 (A201Hr, A289Hr, and A284Hr), NASA (N049Hr, N018Hr, N167Hr, N029Hr, N108Hr, and N154Hr), and the NOAO Gemini/Keck time-exchange program (G329Hr). This paper presents observations made with the Nordic Optical Telescope, operated on the island of La Palma jointly by Denmark, Finland, Iceland, Norway, and Sweden, in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofisica de Canarias. This paper uses observations obtained with the facilities of the Las Cumbres Observatory Global Telescope. The Byrne Observatory at Sedgwick (BOS) is operated by the Las Cumbres Observatory Global Telescope Network and is located at the Sedgwick Reserve, a part of the University of California Natural Reserve System. Data presented in this paper are based on observations obtained at the HAT station at the Submillimeter Array of SAO and the HAT station at the Fred Lawrence Whipple Observatory of SAO. The authors wish to 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. Finally, we thank the referee for a rapid and helpful report which improved the quality of this paper.

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 (A201Hr, A289Hr, and A284Hr), NASA (N049Hr, N018Hr, N167Hr, N029Hr, N108Hr, and N154Hr), and the NOAO Gemini/Keck time-exchange program (G329Hr). Based in part on observations made with the Nordic Optical Telescope, operated on the island of La Palma jointly by Denmark, Finland, Iceland, Norway, and Sweden, in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrofisica de Canarias. Based in part on observations obtained with facilities of the Las Cumbres Observatory Global Telescope. Observations reported here were obtained at the MMT Observatory, a joint facility of the Smithsonian Institution and the University of Arizona.

  • 16 
Please wait… references are loading.
10.1088/0004-6256/144/5/139