HERSCHEL OBSERVATIONS AND UPDATED SPECTRAL ENERGY DISTRIBUTIONS OF FIVE SUNLIKE WITH DEBRIS DISKS

Item Type Article

Authors Dodson-Robinson, Sarah E.; Su, Kate Y. L.; Bryden, Geoff; Harvey, Paul; Green, Joel D.

Citation HERSCHEL OBSERVATIONS AND UPDATED SPECTRAL ENERGY DISTRIBUTIONS OF FIVE SUNLIKE STARS WITH DEBRIS DISKS 2016, 833 (2):183 The Astrophysical Journal

DOI 10.3847/1538-4357/833/2/183

Publisher IOP PUBLISHING LTD

Journal The Astrophysical Journal

Rights © 2016. The American Astronomical Society. All rights reserved.

Download date 05/10/2021 14:00:39

Item License http://rightsstatements.org/vocab/InC/1.0/

Version Final published version

Link to Item http://hdl.handle.net/10150/622672 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 doi:10.3847/1538-4357/833/2/183 © 2016. The American Astronomical Society. All rights reserved.

HERSCHEL OBSERVATIONS AND UPDATED SPECTRAL ENERGY DISTRIBUTIONS OF FIVE SUNLIKE STARS WITH DEBRIS DISKS Sarah E. Dodson-Robinson1,KateY.L.Su2, Geoff Bryden3, Paul Harvey4, and Joel D. Green4,5 1 Department of Physics and Astronomy, University of Delaware, 217 Sharp Lab, Newark, DE 19716, USA; [email protected] 2 Steward Observatory, Department of Astronomy, University of Arizona, 933 North Cherry Avenue, Tucson, AZ 85721, USA 3 Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109, USA 4 Astronomy Department, University of Texas, 2515 Speedway Drive C1400, Austin, TX 78712, USA 5 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA Received 2016 July 25; revised 2016 September 29; accepted 2016 October 1; published 2016 December 16

ABSTRACT Observations from the Herschel Space Observatory have more than doubled the number of wide debris disks orbiting Sunlike stars to include over 30 systems with R>100 AU. Here, we present new Herschel PACS and reanalyzed Spitzer MIPS photometry of five Sunlike stars with wide debris disks, from Kuiper Belt size to R>150 AU. The disk surrounding HD105211 is well resolved, with an angular extent of >14″ along the major axis, and the disks of HD33636, HD50554, and HD52265 are extended beyond the PACS point-spread function size (50% of energy enclosed within radius 4 23).HD105211 also has a 24 μm infrared excess, which was previously overlooked, because of a poorly constrained photospheric model. Archival Spitzer IRS observations indicate that the disks have small grains of minimum radius amin∼3 μm, although amin is larger than the radiation- pressure blowout size in all systems. If modeled as single-temperature blackbodies, the disk temperatures would all be <60 K. Our radiative transfer models predict actual disk radii approximately twice the radius of a model blackbody disk. We find that the Herschel photometry traces dust near the source population of planetesimals. The −5 −4 disk are in the range 2×10 „L/Le„2×10 , consistent with collisions in icy planetesimal belts stirred by Pluto-size dwarf planets. Key words: circumstellar matter – infrared: planetary systems – stars: individual (eta Cru, HD 33636, HD 50554, HD 52265)

1. INTRODUCTION of five Sunlike stars with debris disks, which transition from Kuiper Belt size to R>150 AU. We use the Debris Disk Observations from the Herschel Space Observatory have Simulator (DDS)(Wolf & Hillenbrand 2005) to construct revealed a population of extrasolar debris disks with blackbody model spectral energy distributions (SEDs) and compare our temperatures <60 K and peak flux densities at wavelengths μ disk radii, grain sizes, and luminosities with other debris disks near 100 m. Photodetector Array Camera and orbiting Sunlike stars. Spectrograph (PACS; Pilbratt et al. 2010) 100 and 160 μm ( ) We begin with a discussion of target selection, observations, observations of the DUst around NEarby Stars DUNES and data reduction (Section 2). Next, we discuss the extended Sunlike targets reveal a debris disk incidence of 20%, up from structure observed in our images: our targets include one well ( 12% at the shorter Spitzer MIPS wavelengths Eiroa resolved disk and three moderately extended sources ) et al. 2013 . Far-IR observations are, therefore, opening up a (Section 3). We then present the new Herschel and Spitzer new discovery space of disks more than 100 times more photometry in Section 4. Radiative transfer models of the disk – luminous than the Edgeworth Kuiper Belt, but too cold to SEDs follow in Section 5, with conclusions, and possible future detect at 70 μm. work in Section 6. We now know of over 30 disks with radii R>100 AU (Eiroa et al. 2010; Morales et al. 2013; Duchêne et al. 2014; Pawellek et al. 2014). Since even wide disks require active 2. OBSERVATIONS AND DATA REDUCTION – μ ( collisional cascades to replenish 1 10 m grains Burns We selected five Sunlike stars with debris disks discovered ) et al. 1979 , planetesimal formation must be a robust process in Spitzer observations for Herschel followup. All targets have even at very large distances from the . Planet formation at infrared excess emission seen in both Spitzer IRS observations – 125 250 AU may even have proceeded as far as super-Earth at 32 μm (Dodson-Robinson et al. 2011; Chen et al. 2014) and (Kenyon & Bromley 2015). The wide dust disks revealed in Spitzer MIPS observations at 70 μm (Beichman et al. 2006; by Herschel and earlier (sub)millimeter/scattered light obser- Bryden et al. 2009). (HD105211 also has an MIPS 24 μm vations (e.g., Greaves et al. 1998; Ardila et al. 2004; Krist excess, which was not realized at the time of publication et al. 2005, 2010; Wyatt et al. 2005; Hines et al. 2007; Kalas (Beichman et al. 2006); see Section 5.1 for more detail.) The et al. 2007a, 2007b; Liseau et al. 2008; Golimowski et al. 2011) observing strategy was designed to ensure detection in all may be remnants of planetary systems, which orbit Sunlike Herschel PACS bandpasses. Based on the two detections at stars but are nevertheless very different than our own: different wavelengths, Dodson-Robinson et al. (2011) fit dynamical evidence indicates that the solar was likely single-temperature blackbody models to the Spitzer excess truncated between 80 and 100 AU due to photoevaporation or emission, which we extended to the Herschel PACS band- tidal interactions with nearby young stars (Kretke et al. 2012; passes. Exposure times were selected such that each target Anderson et al. 2013). Here, we present Herschel observations would be detected at the 3σ level, where σ was the quadrature

1 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Table 1 Target Stars

L*/Le Age a b c c d HD Name Spectral Type V (mag) d* (pc) (This Work) L*/Le (Literature) (Gyr) Observation IDs 33636 G0VH-03 7.06 28.7 1.05 1.08 2.5 134226927[2/3] 50554 F8V 6.84 31.0 1.51 1.37 3.3 134223115[2/3], 134226899[6/7], 134226899[8/9] 52265 G0V 6.30 28.1 2.04 2.08 2.6 134223156[4/5], 134226925[8/9] 105211 ηCru F2V 4.15 19.7 7.31 7.06e 1.4f 134226237[1/2], 134226237[3/4] 202206 G6V 8.08 46.3 1.07 1.02 1.1 134221939[7/8], 134223168[6/7]

Notes. a V magnitudes were obtained from color transformations of the Tycho-2 VT photometry (Høg et al. 2000). b Distances come from the Hipparcos reduction of van Leeuwen (2007a, 2007b). c Luminosities and ages from Bonfanti et al. (2015) unless otherwise indicated. d Observation IDs recorded in Herschel Science Archive v7.1.1. Notation: 134226927[2/3] = cross-scan pair of 1342269272 and 1342269273. e from McDonald et al. (2012). f Age from Chen et al. (2014). sum of the instrument noise and the predicted confusion noise structure. Figures 1, 2, 3, and 5 contain curves of growth for reported in the Herschel Observers’ Manual.6 (In reality, some HD33636, HD50554, HD52265, and HD202206, showing of the sources have brighter 160 μm backgrounds than encircled energy fraction (EEF) as a function of circular predicted, leading to detections with low significance or aperture radius for the 100 μm (green) and, where available, nondetections.) 70 μm (blue) mosaics. For comparison, the dashed lines show Each observation consists of PACS medium-speed cross- curves of growth from the point source Vesta (Lutz 2010). For scans separated in position angle by 40°. Program OT1_sdod- HD33636, HD50554, and HD52265, the curves of growth sonr_1 includes simultaneous 100 μm/160 μm observations of are well below those of the Vesta PSF, indicating that the all targets as well as simultaneous 70 μm/160 μm observations sources have extended structure. HD50554 is elongated at   of HD 50554 and HD 105211. Observations are scan maps both 70 and 100 μm, but has a small apparent position angle ′ ″ with eight scan legs of length 3 and a cross-scan step of 4 .We shift between the two bands. Fitting an elliptical Gaussian μ / μ  also obtained 70 m 160 m observations of HD 50554, function to the source, we find a position angle (east of north) HD52265, and HD202206 from the Herschel Science ° μ ° μ ° ( ) of 54 at 70 m and 51 at 100 m. The 3 difference is too Archive program OT1_amoromar_1, PI A. Moro-Martín . small to conclusively indicate morphology changes between Data reduction, beginning with the Level 1 data cubes, was dust emitting at different wavelengths, but should be kept in performed using the Herschel Interactive Processing Environ- mind as a possible source of uncertainty in the photometry and ment version 12.1.0 (HIPE; Ott 2010). We used the HIPE SED fitting. HD202206 is less well resolved than the other implementation of the JScanam map-making algorithm sources, but is still more extended than a point source. For (Graciá-Carpio et al. 2015) to combine scans and cross-scans − some targets, the curves of growth turn downward instead of into oversampled mosaics of 1″ pixel 1 at 70 and 100 μm and ″ μ monotonically increasing with aperture size, indicating that the 2 per pixel at 160 m. The JScanam map-maker removes the fl 1/f signal-drift noise without subtracting extended emission, aperture has expanded to include pixels with negative ux densities. The photometric aperture must be kept smaller than which was an important consideration, because some of our  ( ) sources are not pointlike (see Section 3). However, JScanam the curve-of-growth turnover radius. HD 105211 Figure 4 is also leaves point sources intact, and so can be used for any type clearly resolved in all PACS bands. The extended nature of our of target. Table 1 lists the targets in our sample and their targets leads us use larger photometric apertures than the faint- ( ) Herschel Science Archive observation IDs. The table also source sizes used by the DUNES team Eiroa et al. 2013 .We includes luminosity and age estimates from the literature use 12″ as our nominal aperture size at 70 and 100 μm (see (McDonald et al. 2012; Chen et al. 2014; Bonfanti et al. 2015), Section 4), which captures the extended emission, yet avoids and luminosity estimates from this work.7 Figures 1–5 show negative-valued pixels. the final mosaics in each band for all targets. The upper-right panel of Figure 4 shows the deconvolved 70 μm image of HD105211. Examination of a point source located in the 70° cross-scan, shown in the inset plots of the 3. EXTENDED STRUCTURE original image (upper left) and the deconvolved image, reveals Determining the spatial extent of the disks is important for the extent to which deconvolution has suppressed the trefoil selecting the correct aperture sizes for photometry. Examining PSF structure. Although our deconvolution procedure does not the 70 and 100 μm mosaics in Figures 1–5 suggests that the perfectly conserve flux, one can still see that most of the disks are not point sources, but instead have some extended emission in the deconvolved image is concentrated in a narrow band, indicating a nearly edge-on disk. We performed an 6 Document HERSCHEL-HSC-DOC-0876. elliptical Gaussian source fit on the original (not deconvolved) 7 We calculate luminosities based on V-band magnitudes using 70 μm image and found s =6. 8 and σ =3 9, where the x- L x y * = 2.512()MM- * , where the variables M and L represent L direction is east–west and the y-direction is north–south. The and luminosity. We find better than 10% agreement between our luminosity ° calculations and the literature values, and base all further analysis on our own position angle is 29.7 east of north. Assuming a circular debris luminosity calculations. ring, the radius of the 70 μm emitting ring is approximately

2 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Figure 1. PACS images of HD33636. The rightmost panel shows 100 μm curves of growth for the point source Vesta (dashed line) and HD33636 (solid line). Encircled energy fraction rises more slowly with aperture radius for HD33636 than for Vesta, indicating that HD33636 is extended at 100 μm.

22 ssxy+=7. 9, or 154 AU, in rough agreement with the correction A is fi ( ) SED tting results see Section 5 . The inner edge of the ring is 1 not resolved, possibly due to the high inclination of the disk. A = .1() EEF()ab´ EEF ()

For a=18″ and b=12″,wefind aperture corrections of 4. PHOTOMETRY 1.201 at 70 μm, 1.241 at 100 μm, and 1.384 at 160 μm. For all targets except for HD105211, our aperture Uncertainties have several components: Poisson noise from photometry is based on the PACS ipipe script the source, sky background, PACS absolute flux calibration, L3_pointSourceAperturePhotometry.py8, which and residual 1/f noise, which may elude the map-making uses the HIPE tasks annularSkyAperturePhotometry algorithm. Since the pixels in each final mosaic are not native, and photApertureCorrectionPointSource. For but instead are reconstructed from multiple scans of the source, HD33636, HD50554, and HD202206, the aperture sizes they are not independent and the method of estimating random ( σ are 12″ at 70 and 100 μm, with background subtraction based noise as sskyN pix where is the standard deviation of the sky fl ) on sky annuli of 25″–35″. For HD52265, we use an 8″ ux and Npix is the number of pixels in the aperture does not aperture and an 18″–22″ sky annulus at both 70 and 100 μmto work. Instead, the usual procedure for calculating the random exclude a bright source west of the target. At 160 μm, where component of the PACS-image error budget is to place apertures of the same size and shape used for the science confusion noise is most severe, we use a 14″ aperture and target on the sky background and calculate the standard decrease the sky annulus to 24″–28″ for HD33636 and deviation of the aperture-corrected flux from each. Following HD202206. For HD50554 at 160 μm we use an aperture of ( ) ″ ″– ″ Balog et al. 2014 , we use six apertures placed evenly around 10 and a sky annulus of 20 24 in order to avoid a nearby the outer edge of the background sky annulus for all targets  μ source. HD 52265 is undetected at 160 m due to confusion. except HD105211. The HD105211 sky background is  For the extended source HD 105211, we use the Aperture structured, with an extended bright source directly to the west Photometry Tool (Laher et al. 2012). We choose an elliptical of the disk (Figure 4; see especially the 160 μm image at the aperture of semimajor axis 18″ and semiminor axis 12″ at an lower right). To estimate photometric error for HD105211, we angle of 30° east of north. The sky annulus has major axis 30″– use 12 elliptical apertures placed randomly on the image, 45″ and minor axis 20″–30″ at 70 and 100 μm, and major axis excluding the bright source. Finally, we include conservative 30″–44″ and minor axis 20″–29 3 at 160 μm. The elliptical estimates of the systematic errors—both from absolute flux aperture corrections are based on Table2 of Balog et al. calibration and repeatability—of 7% of the source flux (Balog (2014), which reports EEF as a function of circular aperture et al. 2014). Quoted uncertainties are the quadrature sum of the radius for each PACS filter. For an aperture with semimajor random and systematic errors. Table 2 lists the measured PACS axis a and semiminor axis b, the multiplicative aperture flux densities. As a consistency check, we compare our PACS 70 μm flux 8 μ Author: Markus Nielbock; available in HIPE through the “Scripts  PACS densities with Spitzer MIPS 70 m observations. Published Useful Scripts” menu. values come from the catalog of Chen et al. (2014), which

3 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Figure 2. PACS images of HD50554. The upper-right plot shows curves of growth for the point source Vesta (dashed lines) and HD50554 (solid lines), with 70 μm in blue and 100 μm in green. The slower rise of HD50554ʼs curves of growth compared with Vestaʼs indicates extended structure, which is clearly visible in the 70 and 100 μm images. includes photometry from a variety of literature sources. We 160 μm; observing day 614) and 1342203030/1342203031 also present new photometry of both the 24 and 70 μm images (70/160 μm; observing day 457). The scan/cross-scan pairs based on the MIPS team in-house pipeline (Gordon et al. 2005; were processed with the same JScanam map-making script as Su et al. 2010).9 MIPS did not observe HD202206 at 24 μm. the science observations. We then performed photometry with We use PSF extraction to find 24 μm flux densities for the the nominal aperture and sky annuli used for HD 33636 and remaining targets. All targets have MIPS 70 μm data. We use HD 202206 (our compact, uncrowded targets). Our results PSF extraction to measure 70 μm flux densities for all targets match those of the instrument team (Table12, Balog except HD105211, for which we use an aperture with radius et al. 2014) to within 1% at all wavelengths, demonstrating 35″ and a sky annulus of 39″–65″. the high accuracy of our data pipeline for pointlike sources with ( Table 2 lists the MIPS flux densities for each target. The two clean backgrounds see Section 5 for caveats about the  ) MIPS 24 μm analysis pipelines yield flux densities that are HD 105211 observations . consistent at the 1–2σ level, although there is a clear In the next section, we combine Spitzer and Herschel discrepancy in how the error bars are computed for observations to extend SED coverage to the peak wavelengths HD105211. The MIPS 70 μm flux densities from the two of dust emission, using a radiative transfer model to derive dust pipelines are fully consistent for each source except properties. HD105211. The low HD105211 70 μm flux density of 474 mJy reported by Beichman et al. (2006) is probably due to 5. SPECTRAL ENERGY DISTRIBUTIONS their choice to use a uniform aperture radius of 1.5 camera The disks in this study are cold enough so that observations pixels (7 8) for the entire survey. Since the spatial extent of the at 70 μm and shorter wavelengths probe only the increasing HD105211 disk in our PACS 70 μm images is >14 on the side of the Fν(λ) curve. The Herschel photometry extends major axis, Beichman et al. (2006) may have underestimated coverage to the peak and the decreasing side of Fν(λ). We use the flux density. The remaining PACS and MIPS 70 μm fluxes the DDS10 (Wolf 2003; Wolf & Hillenbrand 2005) to fit SED are consistent at the 1σ level, indicating good agreement models to the Spitzer IRS/MIPS and Herschel PACS between the Spitzer and Herschel fluxes. photometry. The DDS takes as input a stellar spectrum, an As a final test of our photometry pipeline, we measure fluxes inner and outer radius for the dust ring, a dust mass, an from the PACS calibrator star αCeti (HD 18884). We use analytical or numerical description of the dust number density n scan/cross-scan AOR pairs 1342212853/1342212854 (100/ (r)(where r is the distance from the star), and a grain size -q distribution of the form n ()aaaµ ; min aa max (a is the 9 New photometry presented in Table 2 is labeled “Su et al. (2010)” under Analysis Method. 10 www1.astrophysik.uni-kiel.de/dds/

4 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Figure 3. PACS images of HD52265. The upper-right plot shows curves of growth for the point source Vesta (dashed lines) and HD52265 (solid lines), with 70 μm in blue and 100 μm in green. Curves of growth suggest extended structure in the 70 and 100 μm images, although the background confusion may mimic extended structure. Confusion noise prevents detection of HD52265 at 160 μm. grain radius and we set q=3.5). The assumption of a power- the Dodson-Robinson et al. (2011) model photospheres and the law grain size distribution is reasonable for the larger grains, allWISE photometry (Wright et al. 2010; Cutri et al. 2013) and but dynamical models that include loss processes such as found good agreement. The allWISE photometry was particu- Poynting–Robertson drag predict that for grains with larly important for HD105211, which has conflicting photo- a10 μm, the size distribution turns over so that the grain sphere models in the literature due to poor-quality 2MASS abundance increases with radius (Wyatt et al. 2011; Kenyon & photometry; and for HD52265, which has a poor-quality Bromley 2016). Since the far-infrared-emitting dust grains are 2MASS J magnitude. Table 3 lists the fitted parameters for most likely part of a collisional cascade that includes large each SED, and Figures 6–10 show the model SEDs. We begin grains, we set amax=1 mm for all models. Larger particles our discussion of SEDs with a detailed look at the photosphere would be present, but the grain size distribution is so steep that model and mid-IR photometry of HD105211, which requires -12 the larger grains would contribute only (amax 1mm ) = extra care because of the imprecise 2MASS data, and then 3% of the total thermal emission. The DDS also allows the user move on to the ensemble properties of our disks. to set the chemical composition of the grains; here we assume 100% astronomical silicate (Laor & Draine 1993). Since we do not explore the composition options, our model SEDs are 5.1. HD 105211 Mid-IR Photometry and Photosphere plausible but not unique. In particular, the grains could have HD105211 is bright enough to have saturated the 2MASS organic components or ice mantles, although ice would detectors, so the JHK flux densities have >20% errors. The eventually succumb to photodesorption. However, the small near-IR data is important for constraining the Rayleigh–Jeans ’ -2 number of spectral data points that cover the bulk of the disks side of the spectrum, as the turnover to an Fn µ l power law emission do not warrant extremely flexible fits. happens in the H band for HD105211. A further issue is that The parameters we explore are minimum grain size amin, the the SIMBAD database identifies HD105211 as a spectroscopic debris ringʼs distance from the star R, the width of the debris binary, although Eggleton & Tokovinin (2008) find that a ring σ (assuming a Gaussian number-density profile centered at single star is the most probable configuration. If HD105211 is R), and the mass of grains smaller than amax=1 mm. Input star a binary, a single photosphere model might not accurately spectra are Atlas9 model photospheres (Kurucz 1992) origin- reproduce the emission from both components. Beichman et al. ally fitted by Dodson-Robinson et al. (2011), based on (2006) presented the first Spitzer observations of HD105211, Hipparcos/Tycho BT and VT magnitudes (Perryman reporting an infrared excess at 70 μm but not at 24 μm. Their et al. 1997; Høg et al. 2000) and 2MASS JHK magnitudes measured excess was based on a model photosphere with (Cutri et al. 2003). Where available, we added RI photometry Teff=6600 K, which was fitted to BVJHK photometry and from Bessell (1990). We also checked for consistency between predicted a photospheric flux density of 363.4 mJy at 24 μm.

5 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Figure 4. PACS images of HD105211. The disk is well resolved at all PACS wavelengths. The deconvolved 70 μm image in the upper-right panel was created using the point-spread function (PSF) measured from a point source in the 70° cross-scan. The insets in the 70 μm plot panels show the point source before deconvolution (left), with the tri-lobe structure visible, and after deconvolution (right).

Saffe et al. (2008) then measured Teff=6900 K from There is still concern about the IRS data, however. Since slit spectroscopy, which agrees with the optical spectral type loss led to 5%–10% flux density underestimates in IRS, F2V of Gray et al. (2006). Dodson-Robinson et al. (2011) Dodson-Robinson et al. (2011) calculated absolute flux added RI photometry from Bessell (1990) to the photospheric densities by assuming that the shortest-wavelength data points fit constraints, which yielded a best-fit photosphere model with of the SL1 module (beginning at 7.58 mm) traced the Teff=7250 K. Moving to the new, hotter photosphere model photosphere. The LL2 and LL1 data were then spliced to the yields a higher near-IR/mid-IR flux ratio and a lower predicted SL1 data so as to make a continuous function. However, the ( – μ ) photospheric 24 μm flux density of 321.9 mJy. Based on the data at the long-wavelength end of LL2 19 20 m have large Dodson-Robinson et al. (2011) model photosphere and our re- error bars, and there appears to be a drop in fractional excess fl (> μ  reduction of the MIPS data, HD105211 has a 13σ excess at ux at the short-wavelength end of LL1 20 m; see Figure 2 )  24 μm (see Table 2). Chen et al. (2014) computed a third model of Dodson-Robinson et al. 2011 . It is possible that HD 105211 μ fl has multiple debris belts, which would account for the rise and photosphere with a predicted 24 m ux of 347.1 mJy, which – μ combined with the new MIPS re-reduction yields a 7σ excess at fall of fractional excess at 19 20 m, but the IRS data are not μ conclusive. What is concerning is that the IRS data differ from 24 m. The allWISE W1 photometry best corresponds to the ( fl ( ) the allWISE W4 data after conversion to monochromatic ux Dodson-Robinson et al. 2011 photosphere model, with a μ ) ∼ σ ( μ fl σ at 22.1 m by 5 . Beichman et al. 2006 and our MIPS re- monochromatic 3.35 m ux density only 0.11 higher than reduction disagree on the errors in the MIPS 24 μm flux, but the model. As the star is saturated in W2 (see Figure 9),wedo σ= fi adopting 22.5 mJy from Beichman et al. 2006, our MIPS not re- t the model photosphere with the allWISE data added. 24 μm flux density is consistent with the IRS data.) The ( ) We adopt the Dodson-Robinson et al. 2011 photosphere disagreement between W4 and IRS may be due to the W4 color model for this work, but note that the accuracy of the spectral correction; the bandpass is broad and the SED is no longer fi type and would bene t from additional photospheric. The IRS, allWISE, and MIPS data all indicate an data. James (2013) finds an intrinsic scatter of about two infrared excess at 22–24 μm, but the level of excess is still open spectral subtypes in a cross-correlation analysis of their spectral to debate. library—for example, the spectrum of a star with type F8V We also have some concerns about the photometry at might best match the F6V template spectrum—a level of 160 μm due to the bright source east of the target, which may uncertainty consistent with the differences between the have led to over-subtraction of the sky background. We find it analyses of Dodson-Robinson et al. (2011), Chen et al. difficult to reproduce the “peakiness” of the SED at 100 μm; (2014), Saffe et al. (2008), and Gray et al. (2006). dust models that fit the 32/70 μm flux ratio tend to over-predict

6 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Figure 5. PACS images of HD202206. 70 and 100 μm curves of growth do not conclusively show extended structure. the 160 μm flux density, while models that reproduce the 70/ the presence of superheated small grains by computing the 100 μm flux ratio tend to under-predict the 32 μm flux density. parameter G = RRBB, the ratio of the disk radius from our A low abundance of grains near the minimum size, as SED models (in which dust temperature depends on grain size) suggested by collisional cascade models which predict non- to the disk radius if only blackbody grains were present: ( power-law size distributions Wyatt et al. 2011; Kenyon & T = 5100 K() 1ml m () 2 Bromley 2016), might improve the SED fit by decreasing the BB max 12 32/70 μm flux ratio. The SED model shown in Figure 9 is a ⎛ 278 K ⎞2 ⎛ L ⎞ μ RBB = ⎜ ⎟ ⎜ * ⎟ au.() 3 compromise between the opposing constraints at 32 and 70 m. ⎝ T ⎠ ⎝ L ⎠ Reducing the maximum grain size contributing to the emission BB  from the value of 1 mm assumed for all disks could reduce the In Equation (3), TBB is the temperature of blackbody grains, μ fl predicted 160 m ux density, but we doubt the physical λmax is the wavelength of maximum emission, and RBB is the realism of any model which includes small grains without a radius of the ring of blackbody dust. The disks in our sample continuous size distribution extending to macroscopic grains. have 1.7<Γ<2.7 (Table 3). Although four out of five of our The uncertainties in the PACS and MIPS photometry disks are not well enough resolved to measure disk sizes (Section 4) and the underlying photosphere model must be directly from the images, our SED-based values of Γ agree well kept in mind when discussing the dust properties. Based on the ( infrared excess at both 24 μm and 70–160 μm, we strongly with studies of resolved debris disks e.g., Rodriguez & ) suggest followup observations of HD105211 with both JWST Zuckerman 2012; Booth et al. 2013 . Each disk must contain a and ALMA. population of small grains that do not emit efficiently at ~100m m wavelengths, and so heat to above the local blackbody temperature. Given the presence of small grains, 5.2. Dust Properties our sources are most likely self-stirred or planet-stirred We now turn our attention to the properties of the dust in our planetesimal belts with grain-producing collisional cascades, systems, including the disk radius, minimum grain size, which is consistent with our assumption of q = 3.5 (Pan & location of dust relative to the parent planetesimal population, Schlichting 2012; Matthews et al. 2014, p. 521). and possibility of planet sculpting. First, we find that all of the Dodson-Robinson et al. (2011) estimated blackbody tem- dust SEDs peak between 80 and 120 μm, which is typical of peratures TBB for our target disks based on Spitzer IRS and debris disks from the DUNES program (Krivov et al. 2013). MIPS (spectro)photometry at 32 and 70 μm. For the warmer We do not find any sources which are brighter at 160 μm than debris systems HD50554, HD52265, and 105211, their at 100 μm, a feature which could indicate either a background estimates agree well with ours. For HD33636 and in the beam (Gáspár & Rieke 2014) or a disk of HD202206, the peak flux density is much redder than 70 μm unstirred, primordial grains (Krivov et al. 2013). We diagnose and the Herschel observations indicate colder blackbody

7 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Table 2 Photometry

Instrument Wavelength Analysis Method HD 33636 HD 50554 HD 52265 HD 105211 HD 202206 MIPS 24 μm Chen et al. (2014) 43.9±0.9a,b 51.0±1.0b 76.4±1.5b 367.9±22.5c K MIPS 24 μm Su et al. (2010) 42.5±0.5 48.6±0.5 76.2±0.8 374.6±3.8 K MIPS 70 μm Chen et al. (2014) 35.0±2.7d 42.0±4.5d 38.0±5.4d 473.7±19.8c 28.9±3.4e MIPS 70 μm Su et al. (2010) 38.1±2.4 43.5±3.0 40.5±6.0 702±38 27.9±3.4

PACS 70 μm this work K 48.1±4.3 40.3±5.5 692.0±52.6 33.1±3.7 PACS 100 μm this work 42.5±6.1 39.5±10.0 37.3±11.1 704.0±95.1 43.4±4.9 PACS 160 μm this work 30.4±10.2 23.7±8.8f undetected 440.3±146.0 36.1±7.5

Notes. a Flux density units are mJy. Photometry has not been color-corrected. b Original source: IPAC IRSA Spitzer Enhanced Imaging Products Catalog, irsa.ipac.caltech.edu/data/SPITZER/Enhanced/Imaging. c Original source: Beichman et al. (2006). d Original source: Trilling et al. (2008). e Original source: Bryden et al. (2009). f Detection significance is between 2σ and 3σ. temperatures than the Spitzer-based photometry. We emphasize spectral types in our sample. Intriguingly, Pawellek et al. that the SEDs modeled here are not based on blackbody grains (2014) and Pawellek & Krivov (2015) show that amina rp, the —our comparison with single-temperature blackbody models ratio of the true minimum grain size to the minimum predicted simply highlights the importance of far-IR photometry in value from Equation (4)(now with áQapr ()ñ calculated self- characterizing the debris rings. consistently from grain optical properties), is a decreasing The Spitzer IRS spectra provide the best constraint on function of stellar luminosity, approaching unity for A-type minimum grain size. In the Rayleigh regime, where 2πa =λ, stars but with larger values for solar-type stars. Our results the grain opacity κ declines with wavelength as klµ -b, where β=2 for simple conductors and insulators agree well with theirs: the most luminous star in our sample,  / > (Draine 2006).At32μm, the center of the spectral window HD 105211 with L* Le 7, has amina rp ~ 2, while our least used by Dodson-Robinson et al. (2011) to detect infrared luminous star, HD33636 with L*/Le=1.05, has excesses, 2πa = λ for λ=5.1 μm. We find that the lack of amin/arp∼6. Spitzer IRS fractional excesses at 20 μm (with the exception of We suggest that our Herschel observations most likely trace HD105211) and the small, though detectable, excesses at dust near the “birth ring” (Strubbe & Chiang 2006), where the 32 μm exclude grains smaller than ∼3–4 μm, as such grains source planetesimals reside. The largest grains that can emit would emit efficiently from 20–30 μm. The minimum grain like blackbodies at 160 μm have radii set by 2πa≈λ, such that sizes predicted by our best-fit model SEDs are in good a≈25 μm. The lifetime τ of a grain against Poynting– agreement with the Pawellek & Krivov (2015) SED-based Robertson drag is measurements from 32 objects (see their Figure 4): all of their ∼ μ < 16prcRa22 target stars with L< LL*  10 have 2 m t = ,5() a <∼10 μm. min 3L* There is, of course a degeneracy between R and amin.Asan example, one model of the HD50554 debris disk has R=40 where c is the speed of light. For 25 μm grains, we find AU and amin = 5.5m m, while another has R=45 AU and 0.2 Gyrt 2 Gyr for all of the disks in our sample. The χ2 amin = 4.5m m; the difference between their reduced quantity star age/τ for 25 μm grains is between 0.7 and 6.5 for statistics is only 0.2. Yet the value of R from the SED of the all systems except one; HD50554 has age estimates which  resolved source HD 105211 is in rough agreement with the differ by a factor of 10 (3.3 Gyr from Bonfanti et al. (2015) measured size from the image (175 AU versus 154 AU), versus 0.33 Gyr from Chen et al. 2014), and so star age/τ is suggesting that we have broken the degeneracy for at least one fi fi either 1.6 or 16. Since the grains that can contribute source. Indeed, we nd it dif cult to reproduce the measured fi μ 32/70 μm flux ratios without setting a minimum grain size signi cantly to the 160 m emission have lifetimes against / larger than the radiation-pressure blowout size of (Burns radiative drag of at least 1 4 the star age, they are likely to be et al. 1979) destroyed by collisions before migrating significantly. Simi- larly, all systems have β<0.05 (where b = FFpr g; Fpr is the 6LQáñpr () a radiation-pressure force given áQa()ñ=1 and F is the a  * ,4() pr g rp gravitational force) for 25 μm grains, so their orbits are not 16prGM* c substantially modified by radiation pressure. The dust ring where L* and M* are the star luminosity and mass, áQapr ()ñ is widths from our SED fits are likely close to the true widths of the radiation-pressure coupling coefficient averaged over all the source planetesimal belts, with two caveats: (a) mutual frequencies in the stellar spectrum, and ρ is the particle density grain collisions will increase the eccentricity dispersion and, (2.2 g cm−3 for astronomical silicate). In the geometric optics therefore, the ring width (Thébault 2009), and (b) our models limit where áQa()ñ=1 for all values of a,wefind blowout- do not account for the possibility that the small grains have a size limits of 0.5–2.2 μm for the range of star luminosities and wider spatial distribution than the large grains—we assume all

8 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Table 3 Dust Ring Parameters from SED Fits and Images ( )a σ ( )b (μ )c d ( )e ( )f Γg h Star RSED AU AU amin m Mdust (M) TBB K RBB AU Ldust L* HD 33636 85 12 3.0 10−8 46 37 2.3 5.4×10−5 HD 50554 45 4 4.5 3×10−9 59 27 1.7 4.1×10−5 HD 52265 70 10 3.0 3×10−9 58 32 2.2 2.0×10−5 HD 105211 175 20 4.1 6.6×10−8 49 86 2.0 5.4×10−5 HD 202206 105 20 3.0 4×10−8 45 40 2.7 1.3×10−4

Notes. a Distance of the center of the dust ring from the star, computed from SED. b Half-width of debris ring, assuming Gaussian number-density profile for dust. c Minimum grain size. d Dust mass. e Temperature of dust grains, assuming blackbody emission. f Radius of disk of blackbody grains. g R/RBB. h Ratio of dust luminosity to star luminosity.

Figure 6. Spectral energy distribution of the HD33636 star-disk system. The Figure 7. Spectral energy distribution of the HD50554 star-disk system. The inset plot at the lower left shows the best-fit dust ring model, which is centered lower-left inset shows the best-fit Gaussian ring model with center 45 AU and at 85 AU and has width 12 AU. Observational data include Hipparcos BV, RI width 4 AU. Observational data include Hipparcos BV, 2MASS JHK, allWISE, from Bessell (1990), 2MASS JHK, allWISE, Spitzer IRS and MIPS, and PACS Spitzer IRS and MIPS, and Herschel PACS. 100/160 μm. Finally, many lines of evidence suggest that at least some grain sizes are well mixed throughout the dust annulus. In ( reality, radiation pressure might push the smallest grains to planetesimal belts are sculpted by planets e.g., Quillen & ( ) Thorndike 2002; Chiang et al. 2009; Boley et al. 2012;Su& wide or eccentric orbits Burns et al. 1979; Gáspár et al. 2012 ) β  Rieke 2014, p. 318; Sai et al. 2015 . Interestingly, the disk and create structures similar to the Pictoris, HR 8799, and surrounding HD50554 is a true Kuiper Belt analog in its  “ ” ( HD 95086 halos seen in scattered light Augereau spatial distribution, with a best-fit radius of 45 AU and ring et al. 2001; Su et al. 2009, 2015; Ballering et al. 2016). half-width of 4 AU. Since HD50554 is slightly younger than Pushing the small grains in the HD105211 system to wider the Sun (3.3 Gyr; Bonfanti et al. 2015), its disk may be similar orbits might help resolve the conflict between the 32/70 μm to the Kuiper Belt at an earlier stage of evolution and may have flux ratio and the 160 μm flux density discussed in Section 5.1. a Neptune-like planet at its inner edge. In N-body simulations The HD105211 disk also has the smallest amin/arp size ratio of of planetesimal belts with giant planets orbiting just inside their any disk in the sample, and so is most likely to be affected by inner edges, Rodigas et al. (2014) find that the width of the / radiation pressure. Followup with ground-based adaptive optics planetesimal belt increases with both planet mass and planet fi fi or the Hubble Space Telescope could detect scattered light from disk semimajor axes. Here, we nd that the width of the best- t dust ring roughly increases with distance from the star, from blowout-size grains. If our Herschel photometry traces the birth σ= =  σ= > 4AUatR 45 AU for HD 50554 to 20 AU for both rings, our systems can be added to the list of 30 debris disks HD202206 (R=105 AU) and HD105211 (R=175 AU), with radii larger than the probable solar nebula radius of 80 AU although the width of the ring becomes less observationally ( ) Kretke et al. 2012 . Planetesimal formation seems to be a constrained as R increases. However, we have no direct robust process even out to extremely large distances (Kenyon constraints on the presence or absence of planets on wide & Bromley 2013). orbits, and we measure fractional dust luminosities of

9 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

Figure 8. Spectral energy distribution of the HD52265 star-disk system. The Figure 10. Spectral energy distribution of the HD202205 star-disk system. lower-left inset shows the best-fit Gaussian ring model with center 70 AU and The lower-left inset shows the best-fit Gaussian ring model with center 105 AU width 10 AU. Observational data include Hipparcos BV, 2MASS HK, and width 20 AU. Observational data include Hipparcos BV, 2MASS JHK, allWISE, Spitzer IRS and MIPS, and Herschel PACS. HD52265 was not allWISE, Spitzer IRS and MIPS, and Herschel PACS. detected at 160 μm. 2MASS J photometry was flagged as low quality and was not included in the analysis. range, which are larger than the radiation-pressure blowout size. All targets are younger than the Sun (although still on the main sequence), and so we may be seeing Kuiper Belt analogs at earlier phases of evolution. The disks have luminosities consistent with dust production from icy planetesimals stirred by Pluto-size bodies (Kenyon & Bromley 2010). We also find that the Herschel 160 μm emission likely emanates from the birth ring of planetesimals, and so we add our targets to the growing list of Sunlike stars with planetesimals beyond the radius of the solar nebula. One caveat is that our SED models may not be unique: we assumed 100% astronomical silicate grains and did not consider mixtures of silicate, ice, and/or organic grains (although icy grain mantles will likely photoevaporate). Another unexplored possibility is that differ- ent grain sizes have different radial distributions. Although three of the stars host planets discovered by Doppler searches (HD50554, HD52265, and HD202206), the planets are on short-period orbits, which should be dynamically decoupled from the debris disks. Any planets that Figure 9. Spectral energy distribution of the HD105211 star-disk system. The sculpt the debris disks in this sample have not yet been lower-left inset shows the best-fit Gaussian ring model with center 175 AU and detected. HD33636 has an M6 companion in a ∼3 AU orbit width 20 AU. Observational data include Hipparcos BV, Bessell (1990) RI, (Bean et al. 2007), indicating that the debris disk at 85 AU is 2MASS JHK, allWISE, Spitzer IRS and MIPS, and Herschel PACS. circumbinary. HD105211 may also be a spectroscopic binary, HD105211 is saturated in the allWISE W2 filter. although Eggleton & Tokovinin (2008) find a single star more probable. HD33636 and (possibly) HD105211 complement L L =-10--54 10 (Table 3), which are consistent with dust * ∼30 previous detections of circumbinary debris disks, although dust production from icy planetesimals stirred by Pluto-size there is some indication that the frequency of cold debris disks dwarf planets (Kenyon & Bromley 2010). may be higher among single stars (Rodriguez et al. 2015). There is no conclusive evidence that any of our targets host “holey” debris disks, which have planets in the gaps (Kennedy 6. CONCLUSIONS AND FUTURE WORK & Wyatt 2014; Meshkat et al. 2015), although the IRS Here, we have presented new Herschel and updated Spitzer spectrum of HD105211 may have a peak at 19 μm (see photometry of five Sunlike stars with debris disks of Kuiper Section 5.1). Belt size and larger. Both extended structure in the images and Based on target , ALMA followup would be SED fits from the DDS (Wolf & Hillenbrand 2005) indicate possible for HD33636, HD52265, HD105211, and that the disks are not composed of blackbody grains; instead, HD202206. The most compelling target is HD105211 due small grains are present and disks are wider than their to both its brightness—0.7 Jy at 100 μm—and its spatial extent blackbody radii. As seen by Booth et al. (2013), the minimum of >14″, which is wider than the disk surrounding the T Tauri grain sizes in the best-fit SEDs are in the 3.5–4.5 μm size star TWHydrae. HD33636 and HD52265 are also good

10 The Astrophysical Journal, 833:183 (11pp), 2016 December 20 Dodson-Robinson et al.

ALMA targets based on their extended structure in the Gáspár, A., Psaltis, D., Özel, F., Rieke, G. H., & Cooney, A. 2012, ApJ, Herschel PACS images. We also suggest JWST followup of 749, 14 HD105211 using the (a) the MIRI coronagraph with the Gáspár, A., & Rieke, G. H. 2014, ApJ, 784, 33 μ fi ( ) Golimowski, D. A., Krist, J. E., Stapelfeldt, K. R., et al. 2011, AJ, 142, 30 23 m lter and b the MIRI low-resolution spectrometer. Gordon, K. D., Rieke, G. H., & Engelbracht, C. W. 2005, PASP, 117, 503 Based on the allWISE W1 photometry, we find that the the Graciá-Carpio, J., Wetzstein, M., Roussel, H. & the PACS Instrument Control Dodson-Robinson et al. (2011) photosphere model is likely Centre Team 2015, arXiv:1512.03252 accurate, which indicates that HD105211 has a strong infrared Gray, R. O., Corbally, C. J., Garrison, R. F., et al. 2006, AJ, 132, 161 μ Greaves, J. S., Holland, W. S., Moriarty-Schieven, G., et al. 1998, ApJL, excess at 23 m. Observations with JWST would allow us to 506, L133 precisely locate the inner edge of the cold debris ring and Hines, D. C., Schneider, G., Hollenbach, D., et al. 2007, ApJL, 671, L165 possibly detect a hot inner debris belt, if it exists. Høg, E., Fabricius, C., Makarov, V. V., et al. 2000, A&A, 355, L27 James, D. J. 2013, PASP, 125, 1087 Funding for this work was provided by NASA research Kalas, P., Duchêne, G., Fitzgerald, M. P., & Graham, J. R. 2007a, ApJL, 671, L161 support agreement 1524391. We thank Roberta Paladini of the Kalas, P., Fitzgerald, M. P., & Graham, J. R. 2007b, ApJL, 661, L85 HASA Herschel Science Center for guidance on data reduction Kennedy, G. M., & Wyatt, M. C. 2014, MNRAS, 444, 3164 and Sebastian Wolf for developing the publicly available Kenyon, S. J., & Bromley, B. C. 2010, ApJS, 188, 242 Debris Disk Simulator tool. John Gizis and Neal Evans Kenyon, S. J., & Bromley, B. C. 2013, AJ, 143, 63 Kenyon, S. J., & Bromley, B. C. 2015, ApJ, 806, 42 provided helpful input on data analysis methods. This research Kenyon, S. J., & Bromley, B. C. 2016, ApJ, 817, 51 has made use of the following resources: (1) Herschel Kretke, K. A., Levison, H. F., Buie, M. W., & Morbidelli, A. 2012, AJ, 143, 91 Interactive Processing Environment (HIPE) and the ipipe Krist, J. E., Ardila, D. R., Golimowski, D. A., et al. 2005, AJ, 129, 1008 scripts for data reduction. (2) VizieR catalogue access tool, Krist, J. E., Stapelfeldt, K. R., Bryden, G., et al. 2010, AJ, 140, 1051 CDS, Strasbourg, France. The original description of the Krivov, A. V., Eiroa, C., Löhne, T., et al. 2013, ApJ, 772, 32 ( ) / Kurucz, R. L. 1992, in IAU Symp. 149, The Stellar Populations of , VizieR service was published in A&AS 143, 23. 3 NASA ed. B. Barbuy & A. Renzini (Dordrecht: Kluwer), 225 IPAC Infrared Science Archive, which is operated by the Jet Laher, R. R., Gorjian, V., Rebull, L. M., et al. 2012, PASP, 124, 737 Propulsion Laboratory, California Institute of Technology, Laor, A., & Draine, B. T. 1993, ApJ, 402, 441 under contract with the NASA. (4) SIMBAD database, Liseau, R., Risacher, C., Brandeker, A., et al. 2008, A&A, 480, L47 ( ) Lutz, D. 2010, PACS Photometer Point-spread Function. Herschel Science operated at CDS, Strasbourg, France; Wenger et al. 2000 . Center, PICC-ME-TN-033, version 1.01 (5) Herschel Science Archive, www.cosmos.esa.int/web/ Matthews, B. C., Krivov, A. V., Wyatt, M. C., Bryden, G., & Eiroa, C. 2014, in herschel/science-archive. (6) Astropy python library for Protostars and Planets VI, ed. H. Beuther et al. (Tuscon, AZ: Univ. Arizona astronomy, www.astropy.org. (7) APLpy (Astronomical Plot- Press) ting Library in Python), aplpy.github.io. (8) Data products from McDonald, I., Zijlstra, A. A., & Boyer, M. L. 2012, MNRAS, 427, 343 fi Meshkat, T., Bailey, V. P., Su, K. Y. L., et al. 2015, ApJ, 800, 5 the Wide- eld Infrared Survey Explorer, which is a joint Morales, F. Y., Bryden, G., Werner, M. W., & Stapelfeldt, K. R. 2013, ApJ, project of the University of California, Los Angeles, and the Jet 776, 111 Propulsion Laboratory/California Institute of Technology, Ott, S. 2010, in ASP Conf. Ser. 434, Astronomical Data Analysis Software and funded by the National Aeronautics and Space Administration. Systems XIX, ed. Y. Mizumoto, K.-I. Morita, & M. Ohishi (San Francisco, CA: ASP), 139 Pan, M., & Schlichting, H. E. 2012, ApJ, 747, 113 REFERENCES Pawellek, N., & Krivov, A. V. 2015, MNRAS, 454, 3207 Pawellek, N., Krivov, A. V., Marshall, J. P., et al. 2014, ApJ, 792, 65 Anderson, K. R., Adams, F. C., & Calvet, N. 2013, ApJ, 774, 9 Perryman, M. A. C., Lindegren, L., Kovalevsky, J., et al. 1997, A&A, 323, L49 Ardila, D. R., Golimowski, D. A., Krist, J. E., et al. 2004, ApJ, 617, 147 Pilbratt, N. M., Riedinger, J. R., Passvogel, T., et al. 2010, A&A, 518, L1 Augereau, J. C., Nelson, R. P., Lagrange, A. M., Papaloizou, J. C. B., & Popesso, P., Magnelli, B., Buttiglione, S., et al. 2012, arXiv:1211.4257 Mouillet, D. 2001, A&A, 370, 447 Quillen, A. C., & Thorndike, S. 2002, ApJL, 578, L149 Ballering, N. P., Su, K. Y. L., Rieke, G. H., & Gáspár, A. 2016, ApJ, 823, 108 Rodigas, T. J., Malhotra, R., & Hinz, P. M. 2014, ApJ, 780, 65 Balog, Z., Müller, T., Nielbock, M., et al. 2014, ExA, 37, 129 Rodriguez, D. R., Duchêne, G., Henry, T., et al. 2015, MNRAS, 449, 3160 Bean, J. L., McArthur, B. E., Benedict, G. F., et al. 2007, AJ, 134, 749 Rodriguez, D. R., & Zuckerman, B. 2012, ApJ, 745, 147 Beichman, C. A., Bryden, G., Stapelfeldt, K. R., et al. 2006, ApJ, 652, 1674 Saffe, C., Gómez, M., Pintado, O., & González, E. 2008, A&A, 490, 297 Bessell, M. S. 1990, A&AS, 83, 357 Sai, S., Itoh, Y., Fukugawa, M., Shibai, H., & Sumi, T. 2015, PASJ, 67, 20 Boley, A. C., Payne, M. J., Corder, S., et al. 2012, ApJ, 750, 21 Strubbe, L. E., & Chiang, E. I. 2006, ApJ, 648, 652 Bonfanti, A., Ortolani, S., Piotto, G., & Nascimbeni, V. 2015, A&A, 575, 18 Su, K. Y. L., Morrison, S., Malhotra, R., et al. 2015, ApJ, 799, 146 Booth, M., Kennedy, G., Sibthorpe, B., et al. 2013, MNRAS, 428, 1263 Su, K. Y. L., & Rieke, G. 2014, in Proc. IAU Symp. 299, Exploring the Bryden, G., Beichman, C. A., Carpenter, J. M., et al. 2009, ApJ, 705, 1226 Formation and Evolution of Planetary Systems, 318 Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icar, 40, 1 Su, K. Y. L., Rieke, G., Stapelfeldt, K., et al. 2010, BAAS, 42, 349 Chen, C. H., Mittal, T., Kuchner, M., et al. 2014, ApJS, 211, 25 Su, K. Y. L., Rieke, G. H., Stapelfeldt, K. R., et al. 2009, ApJ, 705, 314 Chiang, E., Kite, E., Kalas, P., Graham, J. R., & Clampin, M. 2009, ApJ, Thébault, P. 2009, A&A, 505, 1269 693, 734 Trilling, D. E., Bryden, G., Beichman, C. A., et al. 2008, ApJ, 674, 1086 Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, VizieR Online Data van Leeuwen, F. 2007a, Astrophysics and Space Science Library, Vol. 350 Catalog, 2246 (Berin: Springer) Cutri, R. M., Wright, E. L., Conrow, T., et al. 2013, VizieR Data Catalog, 2328 van Leeuwen, F. 2007b, A&A, 474, 653 Dodson-Robinson, S. E., Beichman, C. A., Carpenter, J. M., & Bryden, G. Wenger, M., Ochsenbein, F., Egret, D., et al. 2000, A&AS, 143, 9 2011, AJ, 141, 11 Wolf, S. 2003, CoPhC, 150, 99 Draine, B. T. 2006, ApJ, 636, 1114 Wolf, S., & Hillenbrand, L. A. 2005, CoPhC, 171, 208 Duchêne, G., Arriaga, P., Wyatt, M., et al. 2014, ApJ, 784, 148 Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, AJ, 140, 1868 Eggleton, P. P., & Tokovinin, A. A. 2008, MNRAS, 389, 869 Wyatt, M. C., Clarke, C. J., & Booth, M. 2011, CeMDA, 111, 1 Eiroa, C., Fedele, D., Maldonado, J., et al. 2010, A&A, 518, L131 Wyatt, M. C., Greaves, J. S., Dent, W. R. F., & Coulson, I. M. 2005, ApJ, Eiroa, C., Marshall, J. P., Mora, A., et al. 2013, A&A, 555, A11 620, 492

11