A High Contrast Survey for Extrasolar Giant with the Simultaneous Differential Imager (SDI)

Item Type text; Electronic Dissertation

Authors Biller, Beth Alison

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author.

Download date 03/10/2021 23:58:44

Link to Item http://hdl.handle.net/10150/194542 A HIGH CONTRAST SURVEY FOR EXTRASOLAR GIANT PLANETS WITH THE SIMULTANEOUS DIFFERENTIAL IMAGER (SDI)

by Beth Alison Biller

A Dissertation Submitted to the Faculty of the DEPARTMENT OF In Partial Fulfillment of the Requirements For the Degree of DOCTOR OF PHILOSOPHY In the Graduate College THE UNIVERSITY OF ARIZONA

2 0 0 7 2

THE UNIVERSITY OF ARIZONA GRADUATE COLLEGE

As members of the Dissertation Committee, we certify that we have read the dis- sertation prepared by Beth Alison Biller entitled “A High Contrast Survey for Extrasolar Giant Planets with the Simultaneous Differential Imager (SDI)” and recommend that it be accepted as fulfilling the dissertation requirement for the Degree of Doctor of Philosophy.

Date: June 29, 2007 Laird Close

Date: June 29, 2007 Don McCarthy

Date: June 29, 2007 John Bieging

Date: June 29, 2007 Glenn Schneider

Final approval and acceptance of this dissertation is contingent upon the candi- date’s submission of the final copies of the dissertation to the Graduate College.

I hereby certify that I have read this dissertation prepared under my direction and recommend that it be accepted as fulfilling the dissertation requirement.

Date: June 29, 2007 Dissertation Director: Laird Close 3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of requirements for an advanced degree at The University of Arizona and is deposited in the Univer- sity Library to be made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permis- sion, provided that accurate acknowledgment of source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the head of the major department or the Dean of the Graduate College when in his or her judgment the proposed use of the material is in the interests of scholarship. In all other instances, however, permission must be obtained from the author.

SIGNED: Beth Alison Biller 4

ACKNOWLEDGMENTS

Science is above all else a collaborative enterprise. First of all, huge thanks to my advisor, Laird Close, for his patience and unwavering support. I would also like to thank the many other scientific collaborators (including my thesis committee) I’ve worked with over the course of this thesis: Eric Nielsen, Don Mc- Carthy, Karl Stapelfeldt, Michael Liu, Markus Kasper, Wolfgang Brandner, Rainer Lenzen, Elena Masciadri, Thomas Henning, Markus Hartung, John Trauger, Eric Mamajek, Aigen Li, Massimo Marengo, John Bieging, Glenn Schneider, Phil Hinz, William Hoffman, Guido Brusa, Douglas Miller, Stephan Kellner, Craig Kulesa, Matthew Kenworthy, Michael Lloyd-Hart, Francois Wildi, Dan Potter, and Ben Oppenheimer. Thanks to the supportive Steward community past and present, including but hardly limited to Iva Momcheva, Jane Rigby, Kim Chapman, Doris Tucker, Michelle Cournoyer, Erin Carlson, Catalina Diaz-Silva, Karen Knierman, Abby Hedden, Wayne Schlingman, Jeff Fookson, Neal Lauver, Janice Lee, Lei Bai, Patrick Young, Jackie Monkiewicz, Eric Nielsen, Kristian Finlator, Moire Prescott, Chien Peng, Matt Kenworthy, Vidya Vaitheeswaran, John Codona, Dave Sudarsky, and Tami Rogers. Thanks to Ry for reading more drafts of this thesis than anyone else – some people have a personal trainer, I’m lucky to have a personal editor. I gratefully acknowledge financial support from NASA through the Graduate Student Researchers Program and future support through the Hubble Fellowship Program. Thanks to my family, as always... after only 29 , I’m finally getting a job. And thanks to my Tucson “family” as well – Hayley, Jillian, Lori, Jeremy, Teresa, Charleen, Doug, Ching, Brandye, Emma, Fonda, Lori, Raven, Monica, Amy, Georgia, Buzz, Lana, Ziva, Susan, Erika, Taylor, Trinity, Madeline, Sarah, and many, many others. You know who you are. 5

DEDICATION

To my grandmother, Rosagene Baron.

To my parents and brother, Larry, Sari, and Alan Biller.

To the amazing women of Midriff Crisis – Brandye, Emma, Fonda, Lori, Raven, Monica, and Amy.

And, last but certainly not least, to Ry. He’s already dedicated a play to me. It’s time I caught up. 6

TABLE OF CONTENTS

LIST OF FIGURES ...... 8 LIST OF TABLES ...... 10 ABSTRACT ...... 11

CHAPTER 1 INTRODUCTION ...... 13 1.1 Many Planets, Few Photons – the Importance of Direct Detection . 14 1.2 The Difficulty of Direct Detection ...... 16 1.3 What this Thesis Contains ...... 21

CHAPTER 2 AN IMAGING SURVEY FOR EXTRASOLAR PLANETS AROUND 45 CLOSE, YOUNG WITH SDI AT THE VLT AND MMT ...... 24 2.1 Introduction ...... 24 2.2 The Simultaneous Differential Imagers at the VLT and MMT . . . . 27 2.2.1 Hardware Considerations ...... 27 2.2.2 Discoveries with the SDI Cameras ...... 28 2.2.3 Observational Techniques and Data Reduction ...... 28 2.3 The SDI Survey ...... 34 2.3.1 Survey Design / Target Selection ...... 34 2.3.2 The Performance of the SDI Filters as Spectral Indices . . . . 36 2.3.3 Contrast Limits and Minimum Detectable Separation 39 2.3.4 Survey Completeness ...... 53 2.3.5 Sensitivity Case Study: AB Dor with Simulated Planets . . . 55 2.3.6 Comparison with Other Direct Detection Methods ...... 56 2.3.7 New and Confirmed Close Binary Stars ...... 57 2.3.8 Candidate Identification / Elimination ...... 58 2.3.9 Planet Detectability ...... 58 2.4 Conclusions ...... 62

CHAPTER 3 DISCOVERY OF A VERY NEARBY TO THE : A METHANE RICH BROWN DWARF COMPANION TO THE LOW SCR 1845-6357 ...... 121 3.1 Introduction ...... 121 3.2 Observations and Data Reduction ...... 122 3.3 Results and Discussion ...... 123 3.3.1 Spectral Type ...... 124 3.3.2 H ...... 125 3.3.3 Likelihood of Being a Bound Companion and T Dwarf Num- ber Densities ...... 126 3.3.4 Mass Estimate for SCR 1845B ...... 128 7

TABLE OF CONTENTS — Continued 3.4 Conclusions ...... 128

CHAPTER 4 HIGH RESOLUTION MID - INFRARED IMAGING OF THE AGB STAR RV BOO WITH THE STEWARD OBSERVATORY ADAPTIVE OPTICS SYS- TEM ...... 132 4.1 Introduction / Application to Planet-finding Science ...... 132 4.2 Circumstellar Structure around Stars . . 132 4.3 Observations and Data Reduction ...... 135 4.4 Followup Observations ...... 140 4.5 Analysis ...... 141 4.6 Discussion ...... 145 4.7 Conclusions ...... 148

CHAPTER 5 PRELIMINARY RESULTS OF A MULTI-WAVELENGTH DIFFEREN- TIAL IMAGING EXPERIMENT FOR THE HIGH CONTRAST IMAGING TESTBED158 5.1 Introduction / Motivation ...... 158 5.2 Data Acquisition and Reduction ...... 161 5.3 Analysis ...... 162 5.4 Conclusions ...... 165

CHAPTER 6 CONCLUSIONS AND FUTURE DIRECTIONS ...... 182 6.1 Conclusions from the SDI Imaging Survey for Extrasolar Planets and Ramifications for Future Planet Imaging Surveys ...... 182 6.2 Development of Technology for Future Planet Searches ...... 184 6.2.1 High Strehls ...... 184 6.2.2 High Strehls and High Contrasts ...... 185

REFERENCES ...... 188 8

LIST OF FIGURES

1.1 Schematic of a Typical AO System ...... 19

2.1 SDI filters ...... 32 2.2 Raw VLT SDI data ...... 33 2.3 Reduced VLT SDI data ...... 35 2.4 Age vs. Distance ...... 37 2.5 SDI methane spectral indices for the T dwarfs SCR 1845B, Gl 229B, Ind Ba, and Ind Bb ...... 40 2.6 Comparison of Contrast Curves generated in 3 different manners for a set of 6 typical program stars ...... 44 2.7 Sensitivity curve for DX ...... 45 2.8 Sensitivity vs. Separation ...... 46 2.9 Sensitivity vs. Separation ...... 47 2.10 Sensitivity vs. Separation ...... 48 2.11 Contrasts for VLT survey objects with H < 4.5 ...... 64 2.12 Contrasts for VLT survey objects with 5.5 > H > 4.5 ...... 65 2.13 Contrasts for VLT survey objects with 6.5 > H > 5.5 ...... 66 2.14 Contrasts for VLT survey objects with 7.5 > H >6.5 ...... 67 2.15 Contrasts for VLT survey objects with H >7.5 ...... 68 2.16 Contrasts for MMT survey objects observed in May 2005 ...... 69 2.17 Contrasts for MMT survey objects observed in February 2006 . . . 70 2.18 FWHM vs. Contrast at 0.5” ...... 71 2.19 Minimum Separations ...... 72 2.20 Contour Plots ...... 73 2.21 50% completeness plot ...... 82 2.22 Reduced VLT SDI data ...... 83 2.23 Maximum Achievable Planet Contrast vs. Separation ...... 84 2.24 Maximum achievable H band planet contrast vs. separation . . . . 85 2.25 Minimum Detectable Planet Mass vs. Separation ...... 86 2.26 Comparison with other direct detection methods ...... 87 2.27 Minimum Detectable Mass vs. Separation ...... 88 2.28 Expected number of planets detected ...... 97

3.1 An SDI image of SCR 1845 ...... 130 3.2 Images of SCR 1845 using the SDI device and reduced using a cus- tom SDI pipeline ...... 131

4.1 9.8, 11.7, and 18 m images of the PSF stars AC Her, UMa, and Her as observed at the MMT ...... 150 4.2 The 11.7 m PSF of AC Her before and after PSF subtraction . . . . 151 9

4.3 AO Images of RV Boo, UMa, Her, and AC Her at 9.8 m . . . . 152 4.4 Position angle of the semi-major axis vs. time (after first observa- tion) for deconvolved RV Boo and UMa nod images ...... 153 4.5 Eccentricity vs. PSF FWHM for RV Boo, UMa, Her, and AC Her images ...... 154 4.6 Deconvolved image of RV Boo ...... 155 4.7 IR emission and model fit to the RV Boo disk ...... 156 4.8 Comparison of RV Boo to the best fit dust disk model ...... 157

5.1 Gallery of single wavelength images and single differences with a nominal contrast level of 106...... 168 5.2 Gallery of single wavelength images and single differences with a nominal contrast level of 107...... 169 5.3 Gallery of single wavelength images and single differences with a nominal contrast level of 108...... 170 5.4 Gallery of single wavelength images and single differences with a nominal contrast level of 109...... 171 5.5 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 106. . . 172 5.6 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 107. . . 173 5.7 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 108. . . 174 5.8 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 109. . . 175 5.9 Boxes used for speckle RMS calculation (Table 2) and trajectories for contrast plots ...... 176 5.10 Contrast as a function of wavelength ( in nm from 798.9 nm) inside and outside of the dark hole...... 177 5.11 Contrast plots inside and outside of the dark hole at a nominal contrast level of 106 ...... 178 5.12 Contrast plots inside and outside of the dark hole at a nominal contrast level of 107 ...... 179 5.13 Contrast plots inside and outside of the dark hole at a nominal contrast level of 108 ...... 180 5.14 Contrast plots inside and outside of the dark hole at a nominal contrast level of 109 ...... 181 10

LIST OF TABLES

2.1 Properties of SDI Survey Stars ...... 98 2.1 Properties of SDI Survey Stars ...... 99 2.1 Properties of SDI Survey Stars ...... 100 2.1 Properties of SDI Survey Stars ...... 101 2.2 VLT SDI Observation Log ...... 102 2.2 VLT SDI Observation Log ...... 103 2.2 VLT SDI Observation Log ...... 104 2.2 VLT SDI Observation Log ...... 105 2.2 VLT SDI Observation Log ...... 106 2.2 VLT SDI Observation Log ...... 107 2.3 MMT SDI Observation Log ...... 108 2.4 Magnitude Offsets ...... 109 2.5 Limiting H mag (5) at 0.5” ...... 110 2.5 Limiting H mag (5) at 0.5” ...... 111 2.5 Limiting H mag (5) at 0.5” ...... 112 2.6 Limiting H mag (5) at 1.0” ...... 113 2.6 Limiting H mag (5) at 1.0” ...... 114 2.6 Limiting H mag (5) at 1.0” ...... 115 2.7 Star/Planet Projected Minimum Detectable Separations for 5 and 10 MJup Planets ...... 116 2.7 Star/Planet Projected Minimum Detectable Separations for 5 and 10 MJup Planets ...... 117 2.7 Star/Planet Projected Minimum Detectable Separations for 5 and 10 MJup Planets ...... 118 2.7 Star/Planet Projected Minimum Detectable Separations for 5 and 10 MJup Planets ...... 119 2.8 Binary Properties ...... 120

3.1 SCR 1845 Photometry ...... 129

4.1 FWHMs for RV Boo and PSF stars ...... 141

5.1 Filter Wavelengths and Bandwidths ...... 166 5.2 Speckle RMS in Right-Side Dark Hole and Left-Side Comparison Region ...... 167 11

ABSTRACT

We present the results of a survey of 45 young (<250 Myr), close (<50 pc) stars with the Simultaneous Differential Imager (SDI) implemented at the VLT and the MMT for the direct detection of extrasolar planets. Our SDI devices use a double Wollaston prism and a quad filter to take images simultaneously at three wavelengths surrounding the 1.62 m methane absorption bandhead found in the spectrum of cool brown dwarfs and extrasolar giant planets. By performing a difference of adaptive optics corrected images in these filters, speckle noise from the primary star can be significantly attenuated, resulting in photon (and flat- field) noise limited data. In our VLT data, we achieved H band contrasts > 10 mag (5) at a separation of 0.5” from the primary star on 45% of our targets and H band contrasts of > 9 mag at a separation of 0.5” on 80% of our targets. With this degree of attenuation, we should be able to image (5 detection) a 7 MJup planet 15 AU from a 70 Myr K1 star at 15 pc or a 7.8 MJup planet at 2 AU from a 12 Myr M star at 10 pc. Using the capabilities of the unique SDI device, we also discovered a methane-rich substellar companion to SCR 1845-6357 (a recently discovered (Hambly et al., 2004) M8.5 star just 3.85 pc from the Sun (Henry et al., 2006)) at a separation of 4.5 AU (1.170” 0.003” on the sky) and fainter by 3.57 0.057 mag in the 1.575 m SDI filter. We also present high resolution ( 0.1”), very high Strehl ratio (0.97 0.03) mid-infrared (IR) adaptive optics (AO) images of the AGB star RV Boo utilizing the MMT adaptive secondary AO system. RV Boo was observed at a number of wavelengths over two epochs and appeared slightly extended at all wavelengths. With such high Strehls we can achieve super-resolutions of 0.1” by deconvolving RV Boo with a point-spread function (PSF) derived from an unresolved star. 12

SDI on ground based telescopes provides significant speckle attenuations down to star-planet contrasts of 1-3 104. To test the classical SDI technique at con- trasts of 1069, we implemented a similar multiwavelength differential imaging scheme for the JPL High Contrast Imaging Testbed. 13

CHAPTER 1

INTRODUCTION

All philosophy is based on two things only: curiosity and poor eyesight; if you had better eyesight, you could see perfectly well whether or not these stars are solar systems, and if you were less curious, you wouldn’t care about knowing, which amounts to the same thing. The trouble is, we want to know more than we can see. – Bernard le Bovier de Fontanelle, from Conversations on the Plurality of Worlds, 1686

It’s very easy to find a planet. Just look under your feet. Perhaps this is why the search for planets around other stars holds so much immediacy and resonance for both the general public and astronomers – we all have much more experience with planets than other astronomical objects, such as nebulae or neutron stars, because we live on one. Planets are home. So it’s not surprising that we want to find more of them. Finding planets tells us some- thing about our own origins and also gives us clues regarding the frequency of inhabitable worlds in the universe. Now if you want to find a second planet, it gets a bit trickier. Maybe you can go out in the evening and check out right after the sun sets. Maybe or is up. So you can directly detect light from five more planets that way and, if you go watch them night after night (as the ancients did), you can see them wander across the sky and past apparently fixed stars. If you’d like to find more planets, you’re going to need a telescope. William Herschel discovered telescopically by chance in 1781. John Couch Adams 14 and Urbain Leverrier posited the existence of from gravitational pertur- bations in the of Uranus. Using Adams’ and Leverrier’s position prediction as a guide, Johann Gottfried Galle discovered Neptune on September 23, 1846. Now, an observer with an 8 in telescope and an ephemeris can find Uranus and Neptune fairly easily – but the initial discovery of these planets was quite a feat! With the discovery of new planets within our own , 18th and 19th century thinkers posited the existence of many other similar planets orbiting around the innumerable stars in the sky. However, the discovery of these worlds was forced to wait over a century, since the difficulty of finding planets in our own solar system pales in comparison to the difficulty of finding planets outside of our solar system. In our solar system we have the advantage of being able to look away from our sun at objects that, while they are faint, are relatively close. Looking for planets around other stars is much more difficult – a planet around another star is intrinsically faint and would be lost in the glare of the star it was orbiting. Directly detecting an extrasolar planet around even the closest star () is akin to being able to resolve a moth near a spotlight in San Diego from Boston (Planetquest Website, planetquest.jpl.nasa.gov). Yet, within the last 15 years, over 200 planets orbiting other stars have been detected indirectly.

1.1 Many Planets, Few Photons – the Importance of Direct Detection

Only relatively recently has the existence of extrasolar planets transitioned from speculation to reality. In October 1995, Michel Major and Didier Queloz an- nounced the discovery of 51 Peg b, the first extrasolar planet. Since then, the tally of extrasolar planets has grown to over 200. The planets discovered so far are only the tip of the iceberg – the full census of planets has only just begun, and 15 will eventually tell us how common worlds like the actually are. Almost all known extrasolar planets have been discovered indirectly. The vast majority of known extrasolar planets were discovered using the technique. In the radial velocity technique (hereafter referred to as RV), one spec- troscopically measures small changes in the radial velocity of the star from the pull of the planet. A number of planets have also been discovered using the tran- sit method, i.e. measuring the slight eclipse of the primary star caused when an extrasolar planet transits in front of it. A few planets have also been discovered by microlensing – where the of an unseen planet and star bends the light from a distant star, causing a temporary increase in the brightness of the distant star. Significant effort has also gone into planet detection through – measuring the slight physical wobble of a star on the sky as it is tugged by a planetary companion. However, while these indirect methods allow estimates of the mass and or- bit of extrasolar planets, they do not provide any information on other physical properties of planets such as composition or . Thus, they are inherently limited. Obtaining actual photons from planets (and eventually spectra) is vital in order to understand the physical properties of planets – to transition between the era of planet discovery into that of planet characteriza- tion. Currently, a spectrum has been obtained (somewhat marginally) for only one planet – HD 209458b (Richardson et al. 2007). This spectrum was attained by observing the spectrum during the secondary eclipse, where the planet tran- sits in front of the star, then subtracting the spectrum during the primary eclipse (where the planet transits behind the star) to obtain the spectrum of the planet by itself. However, this technique is limited to very hot high-mass planets very close to their parent star – a particular extreme subset of planets not indicative of the 16 properties of massive in general. Direct detection followed by direct spectroscopy will be an extremely powerful technique to study the properties of extrasolar planets. Additionally, to gain a better understanding of how planets form, as well as the distribution of earthlike planets, requires a full census of the extrasolar planet distribution as a function of radius. Currently, we are developing this cen- sus from the short radius end outward – the radial velocity technique requires observations on timescales similar to the of a planet to discover that planet (optimally sampling over two full orbital periods). Therefore, over a decade after the first discoveries of extrasolar planets, we are only now beginning to discover planets at similar radii to Jupiter in our own solar system. Indeed, the vast majority of known extrasolar planets have been discovered within 5 AU of their parent star. Direct detection is more effective at larger radii from the parent star and can discover planets in a radius regime which cannot be probed by the radial velocity method in a reasonable amount of time. One important trend to note is that the majority of planets so far detected by all techniques are extremes – extremely massive, extremely close to their star, or around an extremely low mass star (or brown dwarf even). This makes a com- plete accounting of the distribution of extrasolar planets very difficult. The task for the next generation of planet-finding is to push on to discover more “normal” planets – lower mass, further from their stars. Direct detection lets us probe fur- ther out, complementary to the close in regime probed by RV and transit searches.

1.2 The Difficulty of Direct Detection

Currently, very few extrasolar planet candidates have been imaged directly (for in-

stance, 2MASS 1207b ( 8 3 MJup), Oph 1622B ( 13 5 MJup), and CHXR 73 B 17

( 12.5 8 MJup), Chauvin et al., 2005a; Close et al., 2007a; Luhman et al., 2006; Brandeker et al., 2006). The few candidates discovered of “planetary mass” <

13 MJup are companions to brown dwarfs and possess properties more similar to young brown dwarfs (separations > 50 AU; log(g) > 0.3) than to giant extrasolar planets orbiting sun-like stars. Based on their large (>50 AU) separations, these objects appear to have formed via a fragmentation process, more similar to brown dwarfs. Certainly, these objects are too massive to have formed within a circum-brown dwarf disk around their primary (as such an ob- ject around a star would have formed within the star’s circumstellar disk). Hence, to date no true images of extrasolar planets (orbiting around a star rather than a brown dwarf) have been obtained. However, this lack of detections is not from lack of effort. Direct detection of extrasolar planets is extremely technically difficult. Only within the last 20 years has the technology become available to even attempt direct detection. Three major issues must be overcome in order to directly detect extrasolar planets. First, sufficient angular resolution is necessary to image planets which (at least from the model of our solar system) likely lie within 40 AU of their parent star. For stars even as close as 20-40 pc from the Earth, we then require subarcsecond angular resolution in order to image nearby planetary companions. Theoretically, with a large telescope (D > 6 meters) and very good seeing (<0.4”), this angular resolution regime should be attainable. However, planets are drastically fainter than their parent star, so even with this level of angular resolution, a planet will be lost in the glare of the parent star. This brings us to the second issue – analogues to Jupiter ( from 1 - 5

8 MJup, ages from 1 - 5 Gyr) will be >10 times fainter than their parent star in the near-IR and >1011 times fainter in the optical (and as noted above, will lie 18 within 1” of the star). However, one can somewhat sidestep the issue of con- trast by focusing on young planets. Young planets (<100 Myr) are still self- luminous and somewhat brighter – 1047 fainter than their primary stars at near-IR wavelengths. Nonetheless, these contrasts are still punishing and impos- sible to achieve from the ground even with the biggest telescopes and best seeing – because of Earth’s atmosphere, which blurs out the point spread function (here- after PSF) of and reduces angular resolution and contrast within 1” of the star – in other words, stars twinkle. The negative influence of the atmosphere can be countered by either: 1) taking your telescope above the atmosphere (i.e. the method) or 2) finding some way to “remove” the atmosphere (or at least its effects) from your data. One way to “remove” the atmosphere is through adaptive optics (hence- forth AO). Very simply put, adaptive optics removes the twinkle from the stars. On an adaptive optics enabled telescope, some amount of the incoming light is split off and sent to a wavefront sensor as opposed to the science detector. The wavefront sensor measures the aberration of the incoming wavefront (typically on order of a few microns) for a point source (i.e. the guide star.) This mea- sured aberration can then be removed using a deformable mirror tuned to the

1 exact opposite shape (with 2 of the power) of the aberrated wavefront. Since the uncorrected atmospheric wavefront changes on timescales of milliseconds the whole wavefront measurement/correction servo loop must run at millisec- ond sampling speeds in a typical AO stystem (typically 500 Hz). A schematic diagram of a typical AO system is presented in Fig. 1.1. Theoretically, a large telescope (D > 6 meters) plus an adaptive optics (AO) system should be able to reach the photon-noise limit at 100 separations from the star with an hour of exposure time and thus attain the very high (>105) contrasts 19

Figure 1.1 Schematic of a Typical AO System. Incoming light is split according to wavelength. Red light is sent to the science camera, while blue light is sent to the wavefront sensor. The wavefront sensor measures the aberration of the incoming wavefront (typically on order of a few microns) for a point source (i.e. the guide star.) This measured aberration is then be removed using a deformable 1 mirror tuned to the exact opposite shape (with 2 of the power) of the aberrated wavefront. Since the uncorrected atmospheric wavefront changes on timescales of milliseconds the whole wavefront measurement/correction servo loop must run at millisecond sampling speeds in a typical AO stystem (typically 500 Hz). Image reprinted from Gemini Observatory press release on June 2, 2003. 20 necessary to image a young extrasolar . Thus, numerous adaptive optics surveys to directly detect extrasolar planets have been completed (for in- stance, Kaisler et al., 2003; Masciadri et al., 2005). These surveys have yielded in- teresting contrast limits but no true extrasolar giant planet candidates. These null results can be directly attributed to issue number three – bright quasi-static speck- les (also known as super speckles) caused by slowly evolving instrumental aber- rations which remain in adaptive optics images even after adaptive optics correc- tion (see for example Racine et al., 1999). These super speckles evolve stochas- tically on relatively long (minute) timescales and also vary somewhat chromati- cally (especially due to out of pupil optics with phase errors), producing corre- lated speckle noise which is very difficult to calibrate and remove (Racine et al., 1999). For purely photon-noise limited data, the signal to noise (S/N) increases as t0.5, where t is the exposure time. Approximately speaking, for speckle-noise lim- ited data, the S/N does not increase with time past a specific speckle-noise floor (limiting AO contrasts often to 103 at 0.5”, Racine et al. 1999; Masciadri et al. 2005). More exactly, S/N does continue to increase with time, but as the speckle noise in successive frames becomes correlated, the N gain becomes considerably slower. Effectively independent speckles then persist for many minutes rather than a small fraction of a second (Racine et al. 1999). This correlated speckle noise is considerably above the photon noise limit and makes planet detection very difficult. Interestingly, space telescopes such as HST also suffer from limit- ing correlated speckle noise due to temperature variations which induce changes in the PSF (known as “breathing”, Bely , 1993; Schneider et al., 2003). Many observatories, including Gemini, Subaru, and the VLT, are currently building dedicated planet-finding AO/ cameras in order to over- come this speckle noise floor (Dohlen et al., 2006; Macintosh et al., 2006; Tamura 21

& Lyu, 2006). A number of instrumental speckle-attenuation methods have been proposed, such as spectral differential imaging (Racine et al., 1999; Marois et al., 2000, 2002, 2005), azimuthal differential imaging (Marois et al., 2006), inte- gral field spectroscopy (Sparks & Ford, 2002; Berton et al., 2006; Thatte et al., 2007), precise wavelength control methods such as those developed at the High Contrast Imaging Testbed (Trauger et al., 2004), focal plane wavefront sensing (Codona & Angel , 2004; Kenworthy et al., 2006), and nulling interferometry (Liu et al. , 2006).

1.3 What this Thesis Contains

In the first half of this work, I discuss recent advances made in direct detection using the Simultaneous Differential Imagers at the Very Large Telescope (here- after VLT) at Cerro Paranal in Chile and the Multiple Mirror Telescope (hereafter MMT) on Mount Hopkins in Arizona. The Simultaneous Differential Imagers at the VLT and MMT, built and commisioned by our team (Lenzen et al., 2004, 2005; Close et al., 2005a), utilize a spectral differential speckle-attenuation technique (pioneered by Racine et al., 1999; Marois et al., 2000, 2002, 2005). These devices exploit a methane absorption feature at 1.62 m (see Fig. 2.1) which is robustly ob- served in substellar objects with spectral type later than T3.5 (Geballe et al., 2002; Burrows et al., 2001). SDI utilizes specialized hardware to image simultaneously inside and outside this methane feature with custom 25 nm filters (see Fig. 2.1). Since the super-speckles are coherent with the starlight and both starlight and speckles have a flat spectrum (see Fig. 2.1) in this narrow wavelength band ( / 1.6%), subtracting the “on” and “off” methane absorption images removes ' the starlight and its speckles, while preserving light from any substellar methane companion to the star. 22

I discuss results from our recently completed extensive 45 star survey with the SDI devices at the VLT and MMT. Survey stars were chosen primarily ac- cording to proximity to the Sun (<50 pc) and youth (<250 Myr, typically <100 Myr). We observed 54 stars total – from this sample, we attained full contrast curves and good age estimates for 45 stars. We observed 47 young (<250 Myr) stars, 3 nearby stars with known RV planets, and 4 very close (<20 pc) older solar analogues. We obtained contrasts of H>10 mag (5 ) at 0.500 for 45% of target objects at the VLT and contrasts of H>9 mag (5 ) at 0.500 for 80% of our tar- gets. The VLT SDI device is fully commissioned and available to the community and the MMT SDI device is a PI instrument with the ARIES camera. In contrast, the dedicated planet-finding instruments such as Sphere and GPI (Dohlen et al., 2006; Macintosh et al., 2006) being built at the VLT and Gemini will not see first until 2011. Thus, as a precursor to planet surveys with these dedicated planet finding cameras, the results from the SDI devices are especially timely and rel- evant, particularly to inform the large Gemini NICI survey starting in 2007 (Liu et al., 2005). I also discuss the discovery of an interesting companion T dwarf benchmark object using our Simultaneous Differential Imager at the VLT. In the second half, I discuss the development of technology and techniques that may someday be used to detect earthlike extrasolar planets. Using the unique adaptive secondary mirror AO system at the 6.5m MMT (Wildi et al. 2003, Brusa et al. 2003), it is possible to achieve nearly perfect (Strehl ratio 0.97 0.03), high resolution ( 0.100) images at mid-IR wavelengths. In the current work, we used this capability to probe asymptotic giant branch star and proto-planetary nebu- lae morphologies on finer scales than ever before possible in the mid-IR. Through deconvolution, the nearly perfect images (Strehl ratio 0.97 0.03) produced with AO at the MMT allow resolutions better than that of the diffraction limit of the 23 telescope. These very high Strehl ratio images also allow us a glimpse into the future of planet-finding, where high-order AO systems such as the planned AO system for GPI, will make such high Strehl ratios routine at near-IR wavelengths as well. We also explored the performance of the SDI differential imaging technique at extremely high contrasts (1069) at the High Contrast Imaging Testbed at JPL. The High Contrast Imaging Testbed tests precursor technologies for the Terres- trial Planetfinder Mission and consists of a coronographic system kept in vacuum and vibration-isolated. We implemented a multi-wavelength differential imaging experiment similar to the SDI technique as implemented on ground-based tele- scopes. For ground based observing, simultaneous imaging in at least two filters is necessary to overcome the stochastic speckle noise floor remaining even af- ter adaptive optics correction. For space-based observing, however, speckles are stable on timescales of hours to days, making simultaneity of imaging unneces- sary. This multi-wavelength differential imaging experiment measures speckle evolution as a function of wavelength and contrast level. We test whether the ground-based simultaneous differential imaging technique can be generalized to a non-simultaneous differential imaging technique for a space mission. 24

CHAPTER 2

AN IMAGING SURVEY FOR EXTRASOLAR PLANETS AROUND 45 CLOSE, YOUNG

STARS WITH SDI AT THE VLT AND MMT

2.1 Introduction

While1 over 200 extrasolar planets have been detected2 over the last 11 years (mostly via the radial velocity technique), very few extrasolar planet candidates have been imaged directly (for instance, 2MASS 1207b ( 8 3 MJup), Oph 1622B ( 13 5 MJup), and CHXR 73 B ( 12.5 8 MJup) Chauvin et al., 2005a; Close et al., 2007a; Luhman et al., 2006; Brandeker et al., 2006). The few candidates dis- covered of “planetary mass” < 13 MJup are companions to brown dwarfs and possess properties more similar to young brown dwarfs (separations > 50 AU; surface gravity log(g) > 0.3) than to giant extrasolar planets orbiting sun-like stars. Based on their large (>50 AU) separations, these objects appear to have formed via a fragmentation process in a manner analogous to the formation of brown dwarfs. Hence, to date no true images of extrasolar planets have been obtained. Theoretically, a large telescope (D > 6 meters) plus an adaptive optics (AO) system should be able to reach the photon-noise limit at 100 separations from the star with an hour of exposure time and thus attain the very high (>105) contrasts necessary to image a young extrasolar giant planet. Thus, numerous adaptive optics surveys to directly detect extrasolar planets have been completed (for in- stance, Kaisler et al., 2003; Masciadri et al., 2005). These surveys have yielded

1This work first appeared as B. Biller, L. Close, E. Masciadri, E. Nielsen, R. Lenzen, W. Brand- ner, D. McCarthy, M. Hartung, S. Kellner, E. Mamajek, T. Henning, D. Miller, M. Kenworthy and C. Kulesa, 2007, The Astrophysical Journal, in press. Reproduced by permission of the AAS. 2http://exoplanet.eu/catalog.php, maintained by Jean Schneider 25 interesting contrast limits but no true extrasolar giant planet candidates. The difficulty in directly imaging extrasolar giant planets can be attributed to the unfortunate fact that bright quasi-static speckles (also known as super speckles) caused by slowly evolving instrumental aberrations remain in adaptive optics images even after adaptive optics correction (see for example Racine et al., 1999). These super speckles evolve stochastically on relatively long (minute) timescales and also vary somewhat chromatically, producing correlated speckle noise which is very difficult to calibrate and remove (Racine et al., 1999). For photon-noise limited data, the signal to noise S/N increases as t0.5, where t is the exposure time. Approximately speaking, for speckle-noise limited data, the S/N does not increase with time past a specific speckle-noise floor (limiting AO con- trasts often to 103 at 0.5”, Racine et al. 1999; Masciadri et al. 2005). More exactly, S/N does continue to increase with time, but as the speckle noise in successive frames becomes correlated, the sqrt(N) gain becomes considerably slower. Effec- tively independent exposures then have durations of many minutes rather than a small fraction of a second (Racine et al. 1999). This correlated speckle noise is considerably above the photon noise limit and makes planet detection very difficult. Interestingly, space telescopes such as HST also suffer from limiting correlated speckle noise due to temperature variations which induce changes in the PSF (known as “breathing”, Bely , 1993; Schneider et al., 2003). Many observatories, including Gemini, Subaru, and the VLT, are currently building dedicated planet-finding AO/coronagraph cameras in order to over- come this speckle noise floor (Dohlen et al., 2006; Macintosh et al., 2006; Tamura & Lyu, 2006). A number of instrumental speckle-attenuation methods have been proposed, such as spectral differential imaging (Racine et al., 1999; Marois et al., 2000, 2002, 2005), azimuthal differential imaging (Marois et al., 2006), integral 26

field spectroscopy (Sparks & Ford, 2002; Berton et al., 2006; Thatte et al., 2007), precise wavelength control methods such as those developed at the High Con- trast Imaging Testbed (Trauger et al., 2004; Trauger & Traub, 2007), focal plane wavefront sensing (Codona & Angel , 2004; Kenworthy et al., 2006), and nulling interferometry (Liu et al. , 2006). The Simultaneous Differential Imagers at the VLT and MMT, built and commi- sioned by our team (Lenzen et al., 2004, 2005; Close et al., 2005a), utilizes a spec- tral differential speckle-attenuation technique (pioneered by Racine et al., 1999; Marois et al., 2000, 2002, 2005). It exploits a methane absorption feature at 1.62 m (see Fig. 2.1) which is robustly observed in substellar objects with spectral type later than T3.5 (Geballe et al., 2002; Burrows et al., 2001). SDI utilizes specialized hardware to image simultaneously inside and outside this methane feature with custom 25 nm filters (see Fig. 2.1). Since the super-speckles are coherent with the starlight and both starlight and speckles have a flat spectrum (see Fig. 2.1) in this narrow wavelength band ( / 1.6%), subtracting the “on” and “off” methane ' absorption images removes the starlight and its speckles, while preserving light from any substellar methane companion to the star. We have completed a 45 star survey with the SDI device at the VLT and MMT. Survey stars were chosen primarily according to proximity to the Sun (<50 pc) and youth (<300 Myr, typically <100 Myr). We observed 54 stars total – from this sample, we attained full contrast curves and good age estimates for 45 stars. We observed 47 young (<250 Myr) stars, 3 nearby stars with known RV planets, and 4 very close (<20 pc) older solar analogues. We obtained contrasts of H>10 mag (5 ) at 0.500 for 45% of target objects at the VLT and contrasts of H>9 mag (5 ) at 0.500 for 80% of our targets. The VLT SDI device is fully commissioned and available to the community and the MMT SDI device is a PI instrument with 27 the ARIES camera. In contrast, the dedicated planet-finding instruments such as Sphere and GPI (Dohlen et al., 2006; Macintosh et al., 2006) being built at the VLT and Gemini will not see first light for several years. Thus, as a precursor to planet surveys with these dedicated planet finding cameras, the results from the SDI devices are especially timely and relevant, particularly to inform the large Gemini NICI survey starting in 2007 (Liu et al., 2005)

2.2 The Simultaneous Differential Imagers at the VLT and MMT

The VLT Simultaneous Differential Imager (henceforth SDI) was built at the Uni- versity of Arizona by L. Close and installed in a special f/40 camera relay for the VLT AO camera CONICA built by R. Lenzen at the Max Planck Institute for Astronomy, Heidelberg. These were both installed at the VLT in August 2003. The MMT SDI was also built at the University of Arizona. In February 2004, it was installed in the ARIES f/30 camera built by D. McCarthy. Both devices are available to the observing communities of their respective telescopes.

2.2.1 Hardware Considerations

The SDI device consists of a custom double Wollaston, which splits the incoming AO beam into four identical beams (utilizing calcite birefringence to minimize non-common path error – adding only <10 nm rms of differential non-common path errors per the first few Zernikes modes – Lenzen et al. 2004a). Each beam then passes through a narrowband filter with a central wavelength either on or off methane absorption. Three different filters were used; all filters were placed in different quadrants on the same substrate. SDI filters for the VLT and MMT were manufactured by Barr Associates. Filter wavelengths were chosen on and off the methane absorption feature at 1.62 m and were spaced closely (every 0.025 m) in order to limit residuals due to speckle and calcite chromatism. We used four 28

filters F1, F2, F3a, and F3b with measured cold central wavelengths F1 1.575 m, F2 1.600 m, and F3a F3b 1.625 m. The filters are approximately 0.025 m in bandwidth (1.6%). The SDI filter transmission curves overlaid on a theoretical young planet spectrum (private communication, D. Sudarsky) are presented in Fig. 2.1.

2.2.2 Discoveries with the SDI Cameras

The SDI device has already produced a number of important scientific results: the discovery of the important calibrator object AB Dor C (Close et al., 2005b) which is the tightest (0.16”) low mass (0.090 0.05 M, 100 fainter) companion detected by direct imaging, the most detailed methane surface maps of Titan from the pre-Cassini era (Hartung et al., 2004), the discovery of Ind Ba and Bb, the nearest binary brown dwarf (McCaughrean et al., 2004), the discovery of SCR 1845-6357B, a very close (3.85 pc) T6 brown dwarf (Biller et al., 2006b; Kasper et al., 2007), and evidence of orbital motion for Gl 86B, the first known companion to an host star (Mugrauer & Neuhauser,¨ 2005). In fact, the SDI device discovered all known brown dwarfs within 5 pc of the Sun. It has also set the best upper limit on the of the older ( 1 Gyr) extrasolar planet around Eri.

2.2.3 Observational Techniques and Data Reduction

To ensure the highest possible signal to noise ratio and to maximize SDI speckle attenuation, a complex data acquisition procedure was followed for each star. For each object observed, we saturated the inner 0.1” of the star, thus providing a wide dynamic range and contrast down into the halo. Base exposure times (DIT) range from 0.3 to 20 s (typically this was > 2s to allow Fowler sampling at the VLT), depending on the H magnitude of the observed star. A number of 29 exposures (NDIT) with the base exposure time are then coadded in hardware to produce a standard 2 minute long base datum. An example raw datum is presented in Fig. 2.2 3. Base datum are then taken at a grid of dither positions (4 0.5” spacings with the MMT, 5 0.5” spacings with the VLT). This dither pattern is then repeated at typically two telescope “roll angles” (where a “roll angle” refers to a different field derotator position / position angle (henceforth PA) settings). A subtraction of data taken at different roll angles further attenuates super-speckle residuals (since the weak residual speckles after SDI subtraction are instrumental features in the SDI optics which do not shift with a change in roll angle) while producing a very important signature “jump” in position for any physical companion (since a physical companion will appear to shift by the roll angle difference between datasets). For a space telescope such as Hubble (where the entire telescope can be rolled), a companion detected at the 5 level in two different roll angles would be detected at the 7 level (a S/N gain of √2) across the entire dataset (assuming roughly Gaussian statistics). This method is somewhat less effective with ground based telescopes where field rotation is provided by the field derotator rather than rolling the entire telescope (thus, super speckles from the telescope optics can appear to rotate by the roll angle as well). Nonetheless, observing at two roll angles provides us with two independent detections of a substellar companion at different locations on the detector, thus allowing us to rule out a “false positive” detection at an extremely high level of confidence – indeed, the only three faint companions ( Ind Bb, SCR 1845-6357B, and AB Dor C) ever detected with 5 3As with all our survey data, this was taken with the original SDI double Wollaston prism. In February 2007, the original prism was replaced with a next generation prism which is cut in such a way that each subimage now subtends a whole quadrant of the detector chip. The new prism is also fabricated from YV04, a material which produces smaller chromatic errors at 1.6m than the original calcite. 30 using SDI in more than one roll angle have all proven to be real. A typical observing block at the VLT then consists of the following series of : 1) 10 minute long dither pattern taken with a roll angle of 0 degrees. 2) 10 minute long dither pattern taken with a roll angle of 33 degrees. 3) 10 minute long dither pattern taken with a roll angle of 33 degrees. 4) 10 minute long dither pattern taken with a roll angle of 0 degrees. A custom template was developed at the VLT to automate this process in each observation block (hereafter OB). Each base datum was reduced using a custom IDL pipeline (described in detail in Biller et al. (2006a) and Biller et al. (2006c)). This pipeline performs sky-subtraction, flat-fielding, and bad pixel removal, extracts a square aperture around each separate filter image, scales the platescale of each filter image so that the speckles in each filter fall at the same radii despite chromatic differences, scales the flux in each image to remove any quantum efficiency differences be- tween the images, and filters out very low (>15 pixels) spatial frequencies by unsharp masking each image. Each filter image is then initially aligned to a ref- erence image to within 0.25 pixels using a custom shift and subtract algorithm (Biller et al. (2006a,c)). One master reference image is used for each 40 minute long dataset. After each of the filter images has been aligned to the reference im- age, we calculate two differences which are sensitive to substellar companions of

spectral types T (Teff < 1200 K) and “Y” (Teff < 600 K). The first is optimal for T spectral types:

Difference1 = F 1(1.575 m) F 3a(1.625 m) (2.1) The second is optimal for Y spectral types:

Difference2 = F 2(1.6 m) F 3a(1.625 m) (2.2) 31

An additional alignment is performed before the SDI subtraction; using the F1 image as our reference image, we align images F1 and F3a to within 0.05 pixels. A similar alignment is performed with images F2 and F3a, using the F2 image as the reference image. These differences are also somewhat sensitive to hotter substellar companions (L and early T spectral types), due to the fact that the platescale in each filter image has been scaled to a reference platescale to align the Airy patterns in each image. A real object (as opposed to a speckle) will not scale with the Airy pattern and thus, after scaling, will appear at a slightly different radius in each filter image. Subtracting images in different filters will then produce a characteristic dark-light radial pattern for a real object. This effect obviously scales with radius – at the VLT, an object at 0.5” will be offset by less than 1 pixel between filters, while an object at 1.5” will be offset by 3 pixels, producing a very noticeable pattern. Thus, the SDI subtractions have a limited sensitivity to bright L and early T companions. We note that AB Dor C (H 5 mag) was detected at 0.15” (February 2004, Close et al. 2005) and 0.2” (September 2004, Nielsen et al. 2005) separations from AB Dor A even though AB Dor C has no methane absorption features (as is expected from its M5.5 spectral type, Close et al. 2007b.) We additionally calculate one further non-differenced combination sensitive to M, L, and early T companions:

Broadband = F 1(1.575m) + F 2(1.6m) + F 3(1.625m) (2.3)

After each datum is pipelined the data are further processed in IRAF. For each 10 minute long dither pattern, all three combinations described above and the four reduced filter images are median combined. Each 10 minute dataset is then differenced with the following 10 minute dataset (taken at a different position 32 angle). All roll-angle differenced images for each target object observation are then median combined to produce the final data product. A fully reduced 30 minute dataset of AB Dor A (70 Myr K1V star at a dis- tance of 14.98 pc, V=6.88) from the VLT SDI device is presented in Fig. 2.3. Sim- ulated planets have been added at separations of 0.55, 0.85, and 1.35” from the primary, with F1(1.575m) = 10 mag (attenuation in magnitudes in the 1.575 m F1 filter) fainter than the primary. For details and further discussion of these planet simulations see Section 3.4.

2.3 The SDI Survey

2.3.1 Survey Design / Target Selection

Survey objects were selected primarily on the basis of youth and proximity. With a number of exceptions, our 54 observed survey objects are within 50 pc of the Sun and less than 250 Myr in age. (The nine exceptions include three some- what older stars with known radial velocity planets, two more distant (<150 pc) stars with extreme youth indicators, and four older nearby young solar ana- logues which were initially misclassified as young objects.) Distances were ob- tained for 48 of our objects from Hipparcos parallax measurements (parallaxes of >0.02”, corresponding to distances <50 pc, Perryman et al., 1997). Stars were age-selected according to two methods: 1) if possible, according to young clus- ter membership (and adopting the established age for that cluster) for clusters with well established ages such as the Beta Pic, TW Hya, AB Dor and Tuc-Hor moving groups or 2) according to other age indicators including the strength of spectral age indicators (for instance, the Li 6707, the Calcium H and K lines, and H emission) as well as from X-ray emission, variability, and rotational speed. As moving group ages are generally more robust than measurements for individual 33

Figure 2.1 SDI filter transmission curves overlaid on the theoretical spectrum (pri- vate communication, D. Sudarsky) of a young extrasolar planet (30 Myr, 3 MJup). Filters 1 and 2 sample off the 1.62 m CH4 absorption feature, while filter 3 sam- ples within the absorption feature. In contrast, the spectrum of the K2V star Eri (Meyer et al. 1998) is flat across the whole wavelength band. Subtracting images taken in filters “on” and “off” the methane absorption feature will remove the star and speckle noise (which is coherent with the starlight) while preserving any light from giant planet companions. (Details of the complex SDI data pipeline are provided in Section 2.3.) 34

Figure 2.2 Two minutes of raw SDI data from NACO SDI’s 1024 1024 Aladdin array in the VLT CONICA AO camera (Lenzen et al. 2004). A number of elec- tronic ghosts are apparent outside the four square filter apertures (each aperture is rotated by 30); indeed, filter apertures were specifically selected to exclude these ghosts. Note that this is an image of the original Aladdin array; the current SDI array has far fewer bad pixels. 35

Figure 2.3 Left: A complete reduced dataset (28 minutes of data at a series of ro- tator angles (“roll angles”) – 0, 33, 33, 0) from the VLT SDI device. Simulated planets have been added at separations of 0.55, 0.85, and 1.35” from the primary, with F1(1.575m) = 10 mag (star-planet contrast in magnitudes) fainter than the primary. These planets are scaled from unsaturated images of the example star (AB Dor A) taken right before the example dataset (and have fluxes and pho- ton noise in each filter appropriate for a T6 effective temperature). Past 0.7”, the simulated planets are detected in both roll angles with S/N > 10. Observing at two different roll angles produces two independent detections, and hence makes the chance of detecting a “false positive” almost null. Right: Standard AO data reduction of the same dataset. Filter images have been coadded (rather than sub- tracted), flat-fielded, sky-subtracted, and unsharp-masked. Simulated planets have been added with the same properties and at the same separations as before. None of the simulated planets are clearly detected in the standard AO reduction. Additionally, many more bright super speckles remain in the field. 36 stars, we expect the ages of stars in these associations, on average, to have greater accuracy. Our survey covers stars in the Beta Pic, TW Hya, AB Dor, IC 2391, and Tucanae/Horologium moving groups. We select targets stars based on two overlapping criteria: 1) stars within 25 pc and younger than 250 Myr, and 2) stars within 50 pc and younger than 40 Myr (see Fig. 2.4). Our original list has been modified according to the amount of allocated time at the telescope, the unavailability of GTO targets, as well as severe weather constraints for the MMT portion of our survey. At the VLT, our observing runs spanned the months of August through February over 2004 and 2005. Thus, due to the spacing of observing runs, in the south, the survey is close to complete from 17 - 13 hours RA. At the MMT, we had two observing runs, one in May 2005 and one in February 2006. Thus, in the north, the survey is complete for the RA range 11 - 21 hours. Survey objects are presented in Table 2.1. A detailed table of observations is presented in Table 2.2. Survey objects are plotted as a function of distance and age in Fig. 2.4. Our “median” survey object is a K star with an age of 30 Myr and at a distance of 25 pc.

2.3.2 The Performance of the SDI Filters as Spectral Indices

It is important to carefully consider the expected strength of the 1.62 m methane absorption break utilized by the SDI device. The stronger the break strength, the more companion light is preserved after SDI filter subtraction. For a candi- date object with a weak break strength, SDI subtraction may effectively attenuate the candidate object itself, rendering it undetectable (although, at separations > 0.15”, a bright object may still be detectable due to the characteristic dark-light radial pattern produced by any real object after pipelining, see Section 2.2.) To determine the methane break strength expected for a candidate object (and 37

Figure 2.4 Age vs. distance for survey stars within 50 pc and younger than 250 Gyr. Spectral types are delineated by plot symbols. Objects were selected ac- cording to youth and proximity to the Sun. 45 of our survey objects are within 50 pc of the Sun and less than 250 Myr in age. Of the remaining objects, two are very young (<10 Myr), somewhat more distant (<150 pc) objects, three are nearby stars with known RV planets, and four are nearby solar analogues (<20 pc) that were initially misclassified as young. We selected targets according to two overlapping criteria (shown on plot as solid black lines) 1) stars within 25 pc and younger than 250 Myr and 2) stars within 50 pc and younger than 40 Myr. Stars were age-selected according to association membership, or, in the case of unassociated stars, age indicators such as the strength of the Li 6707 A˚ line, Cal- cium H and K lines, H emission, X-ray emission, etc. Distances were obtained from Hipparcos parallax measurements (parallaxes of >0.02”). Our “median” survey object is a K star with an age of 30 Myr and at a distance of 25 pc. 38 thus, the expected performance of SDI for that candidate), we define an SDI methane spectral index calculated from our SDI F1(1.575 m) and F3(1.625 m) filter images (similar to the methane spectral index defined by Geballe et al., 2002).

2=1.5875m F 1 SF 1()d index( ) = R1=1.5625m (2.4) F 3 4=1.6375m S F 3()d R3=1.6125m Each SDI filter was manufactured by Barr Associates to have a precise band- width of 0.025 m, so the wavelength intervals (2 - 1 = = 4 - 3) in the numerator and denominator have the same length for the SDI methane index. We calculated SDI spectral indices for the four brown dwarfs which have been observed with SDI – the T6 Gl 229B (Nakajima et al., 1995), the T5.5 SCR 1845B (Biller et al., 2006b; Kasper et al., 2007) and Ind Ba-Bb (T6 + T1) (McCaughrean et al., 2004). Since we only possess SDI data on a limited number of T dwarfs, we calculated the same SDI spectral indices from spectra of 56 L dwarfs and 35 T dwarfs (Knapp et al., 2004) in order to evaluate the performance of the SDI for a wide range of L and T dwarf objects. Spectra for these objects were obtained from Sandy Leggett’s L and T dwarf archive 4. In order to make an accurate compar- ison, SDI filter transmission curves were convolved into these calculations (see Fig. 2.1). Since we have full spectral data for these objects, we also calculated the 1.62 m methane spectral index defined by Geballe et al. (2002), which was found to be similar to our SDI methane spectral indices. SDI methane spectral indices are plotted for both the M9 and T6 components of SCR 1845, the T dwarfs Gl 229B, Ind Ba, Ind Bb, and 94 other L and T dwarfs in Fig. 2.5. Geballe et al. (2002) note that Gl 229B has an anomalously high methane index for its spectral type and assign a large uncertainty to Gl 229B’s spectral type – T6 1 – which is 4http://www.jach.hawaii.edu/ skl/LTdata.html 39 also reflected in its anomalously large SDI spectral index compared to other T6 dwarfs. From this analysis, we conclude that the SDI device can effectively detect objects with spectral type later than T3. Since T dwarfs with spectral type earlier than T3 are relatively uncommon compared to later T dwarfs, the SDI device can effectively detect the full range of extrasolar giant planet / brown dwarf spectral types of interest. According to the models of Burrows et al. 2003, Baraffe et al.

2003, and Marley et al. 2006, planets >10 Myr old should possess Teff < 800 K and have spectral type of T8 or greater.

2.3.3 Contrast Limits and Minimum Detectable Planet Separation

To determine the range of possible star-planet contrasts achieved in our survey, we generated contrast curves as a function of radius for every survey star. We tested three different methods of generating contrast curves: 1) translating a 6 6 pixel (0.1” 0.1”) box along a particular radial trajectory away from the center of the star image (typical PSF FWHM was 3-5 pixels) then calculating the stan- dard deviation in the box at each point along this trajectory, 2) averaging con- trast curves generated along four such trajectories, and 3) calculating the stan- dard deviation within annular regions six pixels in width centered on the pri- mary PSF (spider diffraction spikes were not masked out in this case because they are already well removed by the spectral difference). Contrast curves gen- erated in these three manners are presented for a set of six typical program stars (AB Dor, DX Leo, GJ 182, AB Pic, GJ 799A, and GJ 799B) in Fig. 2.6. In general, all three methods produce remarkably similar contrast curves and are equally suitable for characterizing the noise properties of an observation. However, we choose to utilize the single trajectory method because it best simulates the par- ticular signal to noise issues encountered when searching for faint companions among super-speckles of similar intensity and FWHM (since it preserves pixel to 40

Figure 2.5 SDI methane spectral indices for the T dwarfs SCR 1845B, Gl 229B, Ind Ba, and Ind Bb (from Biller et al. 2006b). As a comparison, SDI methane spectral indices calculated from spectra for 94 L and T dwarfs (spectra from Knapp et al., 2004) are overplotted. SCR 1845B, Gl 229B, and Ind Bb show strong methane indices, whereas Ind Bb (T1) is relatively constant in flux across the SDI filters and has a much lower methane index. Geballe et al. (2002) note that Gl 229B has an anomalously high methane index for its spectral type. While Geballe et al. (2002) find an overall spectral type of T6 1 for Gl 229B, they assign Gl 229B a spectral type of T7 based on the methane index (which we adopt here). 41 pixel noise variations due to super-speckles). The annular method averages out speckle noise properties azimuthally. This produces somewhat unrealistic results in the case of a faint companion search where one is concerned only with the speckle structure within the local area of a candidate faint companion – speckle structure on the other side of the image is unimportant. In addition, we have tried to choose a very “typical” trajectory per star – ideally, trajectory to trajec- tory variations will average out across the entire survey. Contrast curves for each program star were calculated along a trajectory 45 from the image x axis in the first quadrant. The 45 angle was selected as one of many possible representative trajectories which were unaffected by instrumental effects such as spider arms, vibrations along azimuth or altitude mounts, etc. At each point along this trajectory, the standard deviation was calculated (except for the PSF contrast curve, for which the mean was calculated). A fully labeled example contrast curve for the star DX Leo is presented in Fig. 2.7. Contrast curves were generated for a number of cases for each object. First, a contrast curve was generated for the full reduced and differenced SDI data (labeled SDI data curve) (F1(1.575 m) - F3a(1.625 m) for two roll angles). A PSF contrast curve curve was generated from a median combination of all the F1(1.575 m) filter images for each dataset weighted according to the number of exposures, dithers, and roll angles in the dataset. To recreate the equivalent obser- vation without using the SDI technique (and thus characterize the performance of SDI compared to conventional AO techniques), an “optimized conventional AO” curve was generated by combining images from all three filters at each roll angle:

Broadband = F 1(1.575m) + F 2(1.6m) + F 3(1.625m) (2.5) 42 then unsharp masking (with a Gaussian of width 4 that of the PSF) to remove low spatial frequencies, and subtracting the “Broadband” combinations at differ- ent roll angles from each other. To characterize the noise level in each observation, we calculated an SDI noise curve, which is a combination of photon-noise, flat-field noise, and read noise. Per exposure:

2 2 2 SDI =qphoton + flatfield + readnoise (2.6)

Photon-noise was calculated as:

photon = √nelectrons (2.7)

Readout noise for the CONICA detector at the VLT in Fowler sampling mode is 1.3 ADU (analog-to-digital unit). The gain for the latest CONICA detector in the Fowler sampling mode is 12.1 electrons/ADU so readnoise = 15.73 electrons. NACO and ARIES flat fields were found to be accurate to about 1%, so flat- field noise was estimated as:

flatfield = nelectrons (2.8) where =0.01. The total noise for a full observation (4-5 dithers, 2-4 roll angles, excluding speckle noise) was then calculated by weighting the SDI noise per ex- posure by the number of exposures (NDIT number of dithers number of roll angles):

SDI fullobs = SDI NDIT (number of dithers) (number of roll angles) q (2.9) 43

The PSF curve for a full observation was similarly weighted:

P SF = (medianP SF ) NDIT (number of dithers) (number of roll angles) (2.10) For the sample curve shown in Fig. 2.7, the SDI data is “flat-field” limited within 0.5” of the star. From 0.5” onwards, the SDI data is photon noise limited, approaching the read-noise limit at separations > 2”. We converted our contrasts in electrons to contrasts in magnitudes in the F1(1.625 m) filter. Contrast plots in mag are presented for all non-binary sur- vey objects in Figs. 2.11 to 2.17 according to the H magnitude of the primary for the VLT and according to observing run for the MMT. For every observa- tion which possesses an unsaturated acquisition image (typically 10 0.1 s im- ages taken over 30 s), the stellar peak in the unsaturated acquisition image was used to scale the saturated stellar peak in the saturated data images and thus at- tain accurate contrasts in magnitudes. For observations lacking an unsaturated acquisition image, contrast curves for other stars which had similar peaks, read noise values, and shape to the contrast curve in question were selected from the library of contrast plots in electron units. The peaks utilized for these matching contrast curves were then used to scale the observation missing an acquisition image. A peak of 2.2 105 ADU was adopted for Eri (Kellner et al. 2007, Janson et al. 2007) and Ind A (Geißler et al. 2007). We present contrast curves for 48 stars in this paper; the remaining six survey stars were either very close binaries, making it difficult to generate a contrast curve, or had particularly low quality datasets. For the VLT data, attainable contrast depends on primary star H magnitude as well as seeing FWHM and Strehl ratio during the observation. For the brightest 44

Figure 2.6 Comparison of Contrast Curves generated in 3 different manners for a set of 6 typical program stars (upper left: AB Dor, upper right: DX Leo, middle left: GJ 182, middle right: AB Pic, lower left: GJ 799A, lower right: GJ 799B). Contrast curves were generated by: 1) translating a 6 6 pixel (0.1” 0.1”) box along a particular radial trajectory away from the center of the star image (typical PSF FWHM was 3-5 pixels) then calculating the standard deviation in the box at each point along this trajectory, 2) averaging contrast curves generated along four such trajectories, and 3) calculating the standard deviation within annular regions 6 pixels in width centered on the primary PSF (spider diffraction spikes were not masked out in this case because they are already well removed by the spectral difference). In general, all three methods produce remarkably similar contrast curves and are equally suitable for characterizing the noise properties of an observation. Since it preserves pixel to pixel contrast variations due to speckle noise, the single trajectory method better simulates the S/N issues encountered in searching for faint companions. 45

Figure 2.7 Sensitivity curve for DX Leo (18 pc, K0V, 115 Myr, V=7.05, H=5.242). This is 28 minutes of VLT SDI data. The CONICA PSF curve is the median combination of all the F1(1.575 m) filter images for this dataset (with a gain correction applied which accounted for the number of exposures, dithers, and roll angles). The “optimized conventional AO” curve was generated by averag- ing images from all three filters at each roll angle, unsharp masking to remove low spatial frequencies, then subtracting the combinations at different roll an- gles from each other. The “measured SDI” data curve is the full reduced and differenced SDI data for this object (F1(1.575 m) - F3a(1.625 m) for two roll an- gles). The “theoretical SDI noise” curve is calculated from photon noise (long dashed green curve), flat-field noise (short dashed black curve), and read noise (solid black line) added in quadrature. Within 0.5”, the SDI data is “flat-field” noise limited. (In reality, we are limited by super speckle residuals within this radius. Our flat fields are accurate to the 1% level, but the speckle residu- als <0.5” vary more than this and thus dominate the SDI noise.) From 0.5” onwards, the SDI data is photon-noise limited, asymptotically approaching the read-noise limit at separations > 2”. For a complete set of sensitivity curves, see: http://coatlicue.as.arizona.edu/ bbiller/SDI.html. 46

Figure 2.8 Contrast Plots for Sample Stars with H < 5.5. The CONICA PSF curve is the median combination of all the F1(1.575 m) filter images for each dataset. The optimized conventional AO curve was generated by combining images from all three filters at each roll angle, unsharp masking to remove low spatial fre- quencies, then subtracting the combinations at different roll angles from each other. The SDI data contrast curve is generated from the full reduced and differ- enced SDI data for each object (F1(1.575 m) - F3a(1.625 m) for two roll angles). The SDI noise curve is a combination of photon noise (dot-dashed curve), flat- field noise (green curve), and read noise (solid line). Within 0.5”, the SDI data is flat-field noise limited. From 0.5” onwards, the SDI data is photon noise limited, approaching the read-noise limit at separations > 2”. 47

Figure 2.9 Contrast Plots for Sample Stars with 5.5 < H < 6.5. The CONICA PSF curve is the median combination of all the F1(1.575 m) filter images for each dataset. The optimized conventional AO curve was generated by combin- ing images from all three filters at each roll angle, unsharp masking to remove low spatial frequencies, then subtracting the combinations at different roll angles from each other. The SDI data contrast curve is generated from the full reduced and differenced SDI data for each object (F1(1.575 m) - F3a(1.625 m) for two roll angles). The SDI noise curve is a combination of photon noise (dot-dashed curve), flat-field noise (green curve), and read noise (solid line). Within 0.5”, the SDI data is flat-field noise limited. From 0.5” onwards, the SDI data is photon noise limited, approaching the read-noise limit at separations > 2”. 48

Figure 2.10 Contrast Plots for Sample Stars with H > 6.5. The CONICA PSF curve is the median combination of all the F1(1.575 m) filter images for each dataset. The optimized conventional AO curve was generated by combining images from all three filters at each roll angle, unsharp masking to remove low spatial fre- quencies, then subtracting the combinations at different roll angles from each other. The SDI data contrast curve is generated from the full reduced and differ- enced SDI data for each object (F1(1.575 m) - F3a(1.625 m) for two roll angles). The SDI noise curve is a combination of photon noise (dot-dashed curve), flat- field noise (green curve), and read noise (solid line). Within 0.5”, the SDI data is flat-field noise limited. From 0.5” onwards, the SDI data is photon noise limited, approaching the read-noise limit at separations > 2”. 49 stars in the survey (H<4.5), we attain 5 contrasts of F1 12 mag at separations of >1” from the star. For the faintest survey stars, we only attain 5 contrasts of F1 10 mag >1” from the star. However, considerable spread in attained contrast is observed in each H magnitude bin – most likely due to variations in observing conditions (seeing, Strehl ratio, etc.) across multiple observations. To quantify the effect of seeing on attainable contrast, in Fig. 2.18 we plot the seeing FWHM (averaged over the observation – the error bars on seeing are the seeing variations as measured by the standard deviation of the seeing over each obser- vation) vs. attained 5 contrast at 0.500 for 10 of the stars presented in Fig. 2.12 with H magnitudes between 4.5 – 5.5. For this sample of stars with similar H magnitudes, achievable contrast is roughly inversely proportional to the seeing FWHM. A fair amount of scatter is apparent in this plot and is due in part to seeing variations over the course of each observation. Seeing FWHM can vary considerably over the 20-40 minute timescale of a typical SDI observation, affect- ing the AO system performance and thus the achievable contrast. However, higher attained contrast does not necessarily translate across the board to a lower minimum detectable planet mass. Although one might be able to attain a very high contrast (5 contrast >11 mag at 1” limited by photon noise) for a bright young A star, one would have more luck searching for low luminos- ity planets around an intrinsically faint young M star (5 contrast 9 mag at 1” limited by read noise), since the inherent contrast difference expected between star and planet is considerably smaller. We obtained contrasts of H>10 mag (5 ) at 0.500 for 45% of target objects at the VLT and contrasts of H>9 mag (5 ) at 0.500 for 80% of our targets. This is more a statement on the spectral types in our sample than a performance related issue. In general, the MMT SDI device performed at a slightly lower contrast level 50 than the VLT SDI device – attaining 5 contrasts 0.5-1 magnitude less than those achieved at the VLT for similar separation and primary star H magnitude. The lesser performance of the MMT system can be attributed to two factors. First, the diameter of the MMT is 6.5m versus the VLT, which has an 8.2 m diameter – resulting in a considerable decrease in sensitivity. Additionally, the seeing sam- pled by the MMTAO system was not as stable as for the NACO AO system – Strehl ratios often changed dramatically over an observation, limiting the attain- able contrast. However, the MMT SDI results still probe a higher contrast regime at separations <1” than is possible with standard AO techniques. In order to determine what objects realistically can be detected for our survey stars, we must convert between our instrumental F1(1.625 m) filter magnitudes and H band magnitudes and then compare the H magnitudes to those expected from models of young planets (such as the COND models of Baraffe et al. 2003). To accomplish this, the spectra of both the primary and secondary components of each target must be taken into account. To convert from our F1 filter magnitudes into calibrated H band magnitudes we must calculate the H band magnitude offsets for both the primary star and a potential methane companion (OffsetA and OffsetB respectively):

H = HA HB (2.11)

H = (OffsetB +F 1B) (OffsetA +F 1A) = (OffsetB OffsetA)+F 1 (2.12)

For primary stars with spectral types F-K, we assume that the star has very little chromatic variation within the middle of the H band, so OffsetA is zero (see Fig. 2.1). For lower mass M stars, which are very red, the magnitude offset is not 51 negligible. To take an extreme example, a very low mass M8 primary will have

a magnitude offset of OffsetA=-0.12 0.08 mag (calculated using the spectrum of the M8 star VB10, an H transmission curve, and our F1 filter transmission curve).

The latest stars in our survey have spectral type M0- M5, so OffsetA will be <0.1 mag for these cases. Any T3 or later companion to one of our survey stars will be blue compared to the primary and will appear “brighter” in the F1 filter than in the H band (in other words, it will have a higher “flux” in the F1 filter (number of photons

per unit bandwidth) – see Fig. 2.1) – so OffsetB will definitely be non-negligible.

We calculated OffsetB for 18 objects with spectral types of T4.5-T8 (spectra from Knapp et al., 2004), then averaged together by spectral type to derive an aver-

age offset for each spectral type. For a T5 companion, OffsetT = 0.5 0.05 mag, 5 for a T6 companion, OffsetT = 0.6 0.07 mag, and for a T8 companion, OffsetT 6 8 = 0.87 0.04 mag. Magnitude offsets are presented in tabular form in Table 2.4. While we do not convert our full F1 contrast plots to H contrast plots, for ev- ery survey star we calculate limiting H contrasts (5 values), at 0.5” and 1.0”, equivalent separation in AU, apparent H magnitude, and absolute H magnitude for a T8 spectral type companion (since extrasolar planets are expected to have spectral type > T8, Burrows et al. 2003, Baraffe et al. 2003). These results are presented in Tables 2.5 and 2.6. However, it is difficult to translate our absolute H magnitudes into model planet masses since we have assumed a T8 spectral type in our conversion between F1 and H contrasts – but a companion which ac- tually has the limiting absolute H magnitude we find (combined with the known age and distance of the system) may have a very different spectral type. Since we cannot translate our H magnitudes directly into planetary mass com- panions, we followed the analysis of Masciadri et al. (2005) and translated the- 52 oretical planet models (using the COND models of Baraffe et al. 2003) into H magnitudes then determined the minimum separation at which such a compan- ion could be detected (at the 5 level) in our survey. The minimum separation at

which a 5 MJup or a 10 MJup companion could be detected for each of our survey stars is shown in Table 2.7. Using the Baraffe et al. (2003) COND models, for our

top 15 stars, we detect no 5 MJup planets at separations larger than 24.3 AU and

no 10 MJup planets at separations larger than 9.25 AU. While these numbers are comparable to those found in Masciadri et al. (2005), our current survey actually attains higher contrasts on a case by case basis than Masciadri et al. (2005). Our median survey object has an age of 30 Myr and a K spectral type whereas the me- dian survey object of Masciadri et al. (2005) has a considerably younger age of 12 Myr and an M spectral type – the star-planet contrast is less at younger ages and for later spectral types. Thus one would expect a younger object to have a lower minimum separation at a given attained contrast than a similar but older object and similarly a later spectral type object to have a lower minimum separation at a given attained contrast than a earlier spectral type object with a similar age. For the 10 objects in common between the surveys, our survey attains lower mini- mum separations for 6 out of 10 objects and comparable separations for 2 others (we note also that the objects for which we did not attain lower separations were particularly low quality SDI datasets). Minimum detectable separations for a 5

MJup object for the 10 objects in common are plotted in Fig. 2.19 (using the ages adopted by Masciadri et al. 2005). Our survey is generally more sensitive than Masciadri et al. (2005) on shared stars because the SDI technique allows us to achieve higher contrasts closer to the star (separations of 0.3” - 1.0”) compared to the deep broad-band imaging technique of Masciadri et al. (2005), thus allowing us to potentially detect companions at tighter separations. We also shared four 53 survey objects in common with Lowrance et al. (2005) and one object ( Eri) in common with Luhman and Jayawardhana (2002). In all of these cases, our limit- ing contrasts at 0.5” (H 10-11 mag) are considerably higher than those attained in these previous surveys (H 6.5-7.6 mag), thus we are sensitive to planets at much smaller separations with SDI.

2.3.4 Survey Completeness

One would not expect a planet to be detectable at all phases of its orbit – thus, to really understand the types of planets to which we are sensitive, we must take orbital motion into account and translate separations on the sky into orbital semi- major axes (a). To this end, we generated contour plots of fractional completeness as a function of mass and semi-major axis. For every survey star, we simulate 10000 planets for each combination of mass and semi-major axis. Eccentricities are drawn from a distribution of eccentricities consistent with known radial ve- locity planets. Standard distributions were used to randomly compute viewing angle and orbital phase, giving an instantaneous separation between star and planet. We use the distance, age, spectral type, and H-band magnitude of the star, and luminosity as a function of mass, calculated from the Baraffe et al. (2003) COND models, to provide each simulated planet a separation on the sky in arc- seconds, and an H-band flux ratio compared to its parent star. Since some stars were observed at multiple epochs, each of the simulated planets is advanced in its orbit by the time interval between observations, in most cases about one . At each , the planets are compared to the appropriate contrast curve, and any planet detectable in at least one epoch is considered to be detected. We can then determine the percentage of simulated planets detected as a function of mass and semi-major axis for each survey star. Contour plots for our survey stars are pre- sented in Fig. 2.20. Note that we conservatively assume only T-type objects can 54

be detected, hence masses > 10 MJup are not considered for many young targets. The value attached to each contour level defines the completeness of our obser- vation to detecting (at the 5 confidence level) a planet with the given semi-major axis and mass. It is worth noting that the only assumptions necessary for the gen- eration of these plots is the eccentricity distribution of planets and the Baraffe et al.(2003) COND models. We use this method to summarize our survey completeness in Fig. 2.21. Hav- ing computed the completeness for each star to planets at various masses and semi-major axes, we take slices at representative values of the semi-major axis, and present the number of stars in our 54 star survey which are at least 50% com- plete to such a planet. Our survey places the strongest constraints on planets

between 4-8 MJup with semi-major axes between 20-40 AU. With 20 such stars (with 50% or greater completeness in this mass/semi-major axis range) surveyed without a detection of a planet, a simple way of interpreting our results (though without statistical rigor) is that we would expect the frequency of such planets to be of order 10% or less. A more rigorous analysis is undertaken in Nielsen et al. submitted. The evolutionary models of Baraffe et al. (2003); Burrows et al. (2003) utilize a “hot start” initial condition which, while appropriate for brown dwarfs, is possi- bly significantly different from the actual initial origins of planets. The Burrows et al. (2003) and Baraffe et al. (2003) models begin with a high-temperature, high- entropy hydrogen-helium sphere which is allowed to radiate and cool over time. In contrast, a planet forms when gas accretes onto a rocky core, according to the core- models of Ida and Lin (2005) and the disk instability models of Boss (2003). Recently, Marley et al. (2006) simulated model planets with more re- alistic (lower entropy) initial conditions. These model planets have significantly 55 lower at young ages (<1 Gyr). Model planets also converge to the

“hot start” evolutionary tracks at different times according to mass – a 1 MJup model converges to traditional tracks by 20 Myr, while a 10 MJup requires up to 1 Gyr to match traditional tracks. Currently, H band magnitudes for these models are not yet available, but will be available in Fall 2007 (private communication, J. Fortney). When H band magnitudes are available, we will repeat this analysis using these new models.

2.3.5 Sensitivity Case Study: AB Dor with Simulated Planets

Since our survey data are highly saturated in the core of the images, it is diffi- cult to place simulated objects in our data with a high degree of positional accu- racy, as there is no external reference for position between data taken at different dithers and roll angles. However, as part of the SDI survey observations, our team discovered a close-in (0.15600) companion (hereafter AB Dor C) to the young star AB Dor (Close et al., 2005b). While this companion is a very low mass M star (0.090 0.005 MSun, M5.5 1, Close et al., 2005b, 2007b) and hence, does not possess methane absorption features, it is still clearly detected in our SDI data. In our second AB Dor dataset where AB Dor C is separated from its primary by 0.2” (Nielsen et al. 2005), the AB Dor C source can be used to our advantage as a reference position from which to offset – allowing us to add simulated planets into this dataset with highly accurate positions and relative fluxes independent of our “pipeline” calculated centroids. Simulated planets were produced by scaling 10 0.1 s unsaturated images of AB Dor A taken right before the example dataset. Planets were simulated with F1(1.575m) = 9, 10, 11, and 12 mag and with methane break strengths appro- priate for T5, T6, and T8 spectral types. Methane break strengths were calculated using the methane spectral index defined in Section 2.3.2. Photon noise and zero 56 points appropriate for each object were added using the IRAF artdata/mkobject tool. The photometric zero point was calculated from AB Dor C. A fully reduced 28 minute dataset of AB Dor A (70 Myr K1V at a distance of 14.98 pc, V=6.88) from the VLT SDI device is presented in Fig. 2.22 with simulated planets added at separations of 0.4”, 0.6”, 0.8”, 1.0”, 1.2”, 1.4”, 1.6”, 1.8”, 2.0”, and 2.2” from the primary (F1(1.575m) = 9, 10, 11, and 12 mag and spectral type T8). Past 0.7”, the F1(1.575m) = 10 simulated planets are detected with S/N > 10. The 2.2” object falls off the edge of the aperture in several dithers and thus ap- pears somewhat attenuated compared to the other simulated objects. Maximum achievable companion contrast at the 5 level as a function of distance from the star is plotted in Fig. 2.23. The residual noise curve for this star (see section 2.3.3) is also overplotted. Contrast curves (5) calculated with both techniques agree well with each other. Using the magnitude offsets developed in section 2.3.4, we convert our F1(1.575m) contrasts into H for each spectral type. We adopt

OffsetA = 0 mag, OffsetB = 0.5 mag for a T5 object, OffsetB = 0.6 mag for a T6 object, and OffsetB = 0.87 mag for a T8 object. H vs. separation in arcsec is presented in Fig. 2.24. F1 contrasts were translated into planet masses using the 100 Myr COND models of Baraffe et al. (2003). According to the 100 Myr old model, objects with

mass 10 MJup will have Teff < 900 K – these objects are reliably of spectral types later than T7 (temperature scale from Burgasser et al. 2006). Thus, we adopt the T8 spectral type curve for this analysis. AB Dor has a likely age of 50-70 Myr (Nielsen et al. 2005, Close et al. 2007b) – we interpolate the COND models of Baraffe et al. (2003) to derive masses at these ages as well. The minimum de- tectable planet mass as a function of distance from the star is plotted in Fig. 2.25.

Adopting an age of 70 Myr for AB Dor A, we can detect a 7 MJup planet 12 AU 57 from the star. However, as noted above, the Baraffe et al. (2003) models utilize a hot start initial condition which may be inappropriate for a young planet. The Marley et al. (2006) models utilize more appropriate initial conditions and when H band magnitudes become available for these models, we will repeat this anal- ysis.

2.3.6 Comparison with Other Direct Detection Methods

We believe that our SDI images are the highest contrast astronomical images ever made from ground or space for methane rich companions 1” from their star. To substantiate this claim, we compare our SDI contrast curves with those pro- duced using a variety of other competing methods (Azimuthal Differential Imag- ing (ADI), Marois et al. (2006), Lyot Coronagraph, Hinkley et al. (2007), HST NICMOS, Schneider et al. (2003), K-band Keck AO, Schneider et al. (2003), and NACO deep imaging in the Ks band, Masciadri et al. (2005)). Comparison con- trast curves are presented in Fig. 2.26. Apart from the Lyot and NICMOS curves, all curves are from 8m class telescopes. For ease of comparison, we convert our F1=1.575 m SDI contrast curve into the equivalent H contrast appropriate for T8 and Y spectral type companions. For methanated companions, SDI provides improved contrast by 1-4 mag within 100 as compared to other methods.

2.3.7 New and Confirmed Close Binary Stars

A number of close binary stars were discovered or confirmed during our survey. In Table 2.8, we present separations and position angles measured from unsatu- rated SDI images of these stars acquired before each full SDI dataset was taken. These values are meant as estimates, hence, no error estimate is provided. We dis- covered close stellar candidates to HIP 9141 (0.15” measured SDI separation), AB Dor A (0.16” measured SDI separation, see Close et al. 2005a), HD 48189A (0.14” 58 measured SDI separation), HD 135363 (0.26” measured SDI separation) and CD- 64 1208 (0.18” measured SDI separation). The <0.5” separation between the pri- mary stars and these objects makes it highly improbable that they are background objects. Additionally, we confirmed the close binary RXJ 1243.6-7834 (0.068” mea- sured SDI separation) discovered by Brandner et al. (2000), the visual double LH 98 062 (2.4” measured SDI separation) discovered by Mochnacki et al. (2002), the spectroscopic binary TWA 4 (0.78” measured SDI separation) discovered by Torres et al. (1995) and the close binary EK Dra (0.67” measured SDI separation) discovered by Metchev and Hillenbrand (2004).

2.3.8 Candidate Identification / Elimination

Survey data were examined for planet candidates by eye and also using auto- mated detection algorithms; generally, the human eye proved more effective for detecting candidates. We identified eight very tentative planet candidates at the VLT which passed the following tests: 1) Candidate must appear at the appropriate positions in the full reduced data. (i.e. candidate image position must jump by the appropriate roll angle.) 2) Candidate must appear (at least marginally) at the appropriate position in each of the separate roll angle images 3) Candidates detected in the F1(1.575 m) - F3a(1.625 m) difference should also be detected in the F2(1.6 m) - F3a(1.625 m) difference as well. A candidate detected at the 3 level in one image which passes all of these tests is essentially detected at the 6 level across the entire dataset (where two factors of √2 has been added in because the object is seen in multiple roll angles and spectral differences). These eight extremely tentative (<2) candidates are noted in the comments column of Table 2.1, with the predicted mass (from the COND models of Baraffe et al. (2003)) and separation had it been real. No can- 59 didates were detected with > 3 in more than one roll angle image. None of the 8 tentative candidates were detected at a second epoch, thus the survey reached a null result for extrasolar planets at the 3 level and certainly at the 5 level analyzed here.

2.3.9 Planet Detectability

To determine what sort of planets we can detect in this survey, we converted our contrast curves in mag units into minimum detectable mass vs. separation (assuming a late T to early Y spectral type for all possible objects and using the COND models of Baraffe et al. (2003)). We calculated minimum detectable mass vs. separation for all stars with contrast curves in Figs. 2.11 to Figs. 2.17; mini- mum detectable mass vs. separation is presented for a set of four typical survey stars (AB Dor, DX Leo, GJ 182, and GJ 799B) in Fig. 2.27. However, to detect an object of any given mass requires that such an object exists around its parent ob- ject! The likelihood of detecting any object at a given radius is a combination of the minimum detectable mass for the parent star at that radius and the likelihood of such an object existing. Therefore it is very important to fully characterize and understand the expected distribution of objects around each survey star. The results of the survey then also constrain the possible distribution of extrasolar planets as a function of radius. To this end, we ran detailed Monte Carlo simulations to characterize the en- semble of planets expected to exist around each star. We conducted a similar simulation to that used to produce the contour plots of Fig. 2.20, as described in Section 2.3.4 (these simulations are described in much more depth in Nielsen et al. 2006). In contrast to the production of the contour plots, we simulated 106 planets instead of 104, and mass and semi-major axis were assigned distributions of their own. The mass and semi-major axis distributions, like the distribution 60 for eccentricity, were produced by considering the population of published ra- dial velocity planets (e.g. Butler et al. 2006), with mass and eccentricity both chosen to fit the histograms from observed planets. Semi-major axis has been ob- served to follow a distribution of dN a1 for radial velocity planets (Wright et da ∝ al. 2005). Orbital inclinations were assigned randomly. Since the radial velocity method has an inherent bias toward close-in planets (which have shorter orbital periods and larger radial velocity amplitudes), we attempted to correct for this by assuming a power-law distribution that is constant in semi-major axis – i.e. dN constant. We considered the results of Fischer and Valenti’s (2005) volume- da ∝ limited sample, and chose an outer limit for the semi-major axis distribution such that, for stars in the range in our sample, each star is expected to host one planet. This was done by integrating the semi-major axis distribution from 0.02 AU (corresponding to HD 41004Bb, the closest-in exoplanet known thus far) to 2.5 AU, the detection limit for the sample of Fischer and Valenti (2005), then noting the fraction of stars with planets in the metallicity range (-0.5 < [Fe/H] < 0.25) of our target stars (4.1%) and choosing an upper cut-off to the distribution when the integral reaches 100%. This gave us a constant probability distribution for semi-major axis between 0.02 and 45 AU that contains the same number of planets found in the <2.5 AU radial velocity survey. The ensemble of simulated planets is shown for our set of four typical stars in Fig. 2.27. Simulated planets which are detected are plotted as blue dots and those that remain undetected are plotted as red dots. In addition to the contrast plot, we also consider a planet “undetectable” when its apparent H magnitude drops below 21 mag (a limit set by our total integration time), or when the planet’s temperature rises above 1400 K (given as a function of age and planet mass by Baraffe et al. 2003). Above this temperature, the strength of the 1.62 m methane 61 break weakens to the point that the SDI method loses effectiveness. Since we assume that each program star possesses exactly one planet that follows the dis- tributions given above, we can assign a detection probability for that star from the percentage of the simulated planets that are detectable at the 5 level. For our 48 program stars (consisting of 40 stars with ages <250 Myr and closer than 50 pc, one 10 Myr old star at a distance of 67 pc, three stars with known RV plan- ets and four nearby solar analogues) which possess contrast curves, the average detection probability is 4.6%, the median detection probability is 3.5%, and the maximum detection probability is 33%. We have chosen to leave the older stars in this sample in our statistics even though their detection probabilities are essen- tially zero. Integrating over the probability distribution of our program stars, in Fig. 2.28 we plot the number of planets we expect to detect as a function of total stars observed, ordering the results so that the best stars (highest detection proba- bilities) are considered first. For the 48 stars in our surveys for which we acquired contrast curves, we expect to detect a total of 2-3 planets (2.73 to be exact) based on the above assumptions. Thus, using Poisson statistics, our survey null detec- tion rules out this exoplanet distribution at the 93% level. It is important to note that this null result shows that this particular combination of assumptions (mass distribution, eccentricity distribution, constant semi-major axis distribution, up- per limit to semi-major axis at 45 AU, assumption that each star has a planet, and the mass-luminosity conversion from the COND models of Baraffe et al. 2003) is ruled out to this confidence level; determining which individual assumptions are incorrect will required data beyond that of the current survey. Here we rule out one possible exoplanet distribution; further exoplanet dis- tributions are considered in Nielsen et al. (submitted). Nielsen et al. (submitted) examines the implications for the distribution of extrasolar planets based on the 62 null results from the SDI survey combined with the VLT NACO adaptive optics deep imaging survey of Masciadri et al. (2005), two of the largest direct imaging surveys published to date. Combining the measured contrast curves from 23 of the stars observed with the VLT NACO adaptive optics system by Masciadri et al. (2005), and 47 of the stars observed with the VLT NACO SDI and MMT SDI devices by Biller et al. (2007) (for a total of 59 unique stars), they consider what distributions of planet masses and semi-major axes can be ruled out by these data, based on Monte Carlo simulations of planet populations. They can set this upper limit with 95% confidence: the fraction of stars with planets with semi-major axis from 20 to 100 AU, and mass >4 MJup, is 20% or less. Also at the 95% confidence

1.16 level, with a distribution of planet mass of dN/dM M between 0.5-13 MJup, we can rule out a power-law distribution for semi-major axis (dN/da a) with index 0 and upper cut-off of 17 AU, and index -0.5 with an upper cut-off of 46 AU. For the distribution suggested by Cumming et al. (2007), a power-law of index =-0.61, they place an upper limit of 73 AU on the semi-major axis distribution, again at the 95% confidence level. In other words, given these assumptions for the semi-major axis distribution, and using the models of Burrows et al. (2003), giant planets are rare past 73 AU. With current observations, they cannot reject the Ida & Lin (2004) models for the masses and semi-major axes of giant planets with better than 50% confidence. In general, we find that even null results from direct imaging surveys are very powerful in constraining the distributions of gi- ant planets at large separations, but more work needs to be done to close the gap between planets that can be detected by direct imaging, and those to which the radial velocity method is sensitive. 63

2.4 Conclusions

We obtained datasets for 54 stars (45 stars were observed in the southern sky at the VLT, 11 stars were observed in the northern sky at the MMT, and 2 stars were observed at both telescopes). In our VLT data, we achieved H band contrasts > 10 mag (5) at a separation of 1.0” from the primary star on 45% of our targets and H band contrasts of > 9 mag at a separation of 0.5” on 80% of our targets.

With these contrasts, we should be able to image (5 detection) a 7 MJup planet

15 AU from the star around a 70 Myr K1 star at 15 pc or a 7.8 MJup planet at 2 AU from a 12 Myr M star at 10 pc. We believe that our SDI images are the highest contrast astronomical images ever made from ground or space for methane rich companions within 1” of their primary star. Eight tentative candidates were identified (none with S/N > 2 in a single image, corresponding to S/N > 4 in all four images together). Had these can- didates been real, they would have possessed separations of 3 - 15.5 AU and

masses of 2-10 MJup. However, none of the candidates were detected in second epoch observations. Thus, we find a null result from our survey. Nonetheless, our result still has serious implications for the distribution of extrasolar planets. In the course of our survey, we also discovered five new close stellar binary systems with measured separations of 0.14” to 0.26”.

For 20 of our survey stars, we attained 50% completeness for 4-8 MJup planets at semi-major axes of 20-40 AU. Thus, our completeness levels are sufficient to significantly test theoretical planet distributions. From our survey null result, we can rule out (at the 93% level) a model planet population using a constant distribution ( dN constant) of planet semi-major axis out to a distance of 45 AU da ∝ (a number of further exoplanet distribution models are considered in Nielsen et al. submitted). Our null detection in this survey sets strong upper limits on the 64 distribution of young massive extrasolar planets >20 AU from their primaries and provides valuable contraints for theories of planet formation and migration. 65

Figure 2.11 5 Contrasts for VLT SDI survey objects with H < 4.5 in the F1(1.575 m) filter. These contrast curves were generated by translating a 6 6 pixel (0.1” 0.1”) box along a particular radial trajectory away from the center of the star and then calculating the standard deviation within that box as a function of radius. Curves were generated from the full reduced and differenced SDI data for each object (F1(1.575 m) - F3a(1.625 m) for two roll angles). 66

Figure 2.12 Same as Fig. 2.11 but for VLT SDI survey objects with 5.5 > H > 4.5. 67

Figure 2.13 Same as Fig. 2.11 but for VLT SDI survey objects with 6.5 > H > 5.5. 68

Figure 2.14 Same as Fig. 2.11 but for VLT SDI survey objects with 7.5 > H > 6.5. 69

Figure 2.15 Same as Fig. 2.11 but for VLT SDI survey objects with H > 7.5. 70

Figure 2.16 5 contrasts for MMT SDI survey objects observed in May 2005. These contrast curves were generated by translating a 6 6 (0.1” 0.1”) pixel box along a particular radial trajectory away from the center of the star and then calculating the standard deviation within that box as a function of radius. Curves were gen- erated from the full reduced and differenced SDI data for each object (F1(1.575 m) - F3a(1.625 m) for two roll angles). 71

Figure 2.17 Same as Fig. 2.16 but for MMT SDI survey objects observed in Febru- ary 2006. 72

Figure 2.18 Seeing FWHM (averaged over each observation) vs. attained 5 con- trast at 0.500 separation from the primary star for 10 of the stars presented in Fig. 2.12 with H magnitudes between 4.5 – 5.5. The error bars on seeing are the seeing variation (as measured by the standard deviation of the seeing) over each observation. For this sample of stars with roughly the same H magnitude, achiev- able contrast varies roughly inversely with the average seeing FWHM. Scatter in this plot is in part due to the fact that seeing FWHM can change considerably over a 20-40 minute long observation. 73

Figure 2.19 Minimum detectable planet separations for a 5 MJup planet for the 10 objects in common between this survey and Masciadri et al. (2005) who used VLT NACO without SDI. For the purpose of comparison, we have adopted ages from Masciadri et al. (2005); we note our preferred age on the figure where our adopted ages differ from Masciadri et al. (2005). We used theoretical 5 MJup planet model (Baraffe et al. 2003) H magnitudes for these 10 cases then determined the mini- mum separation at which such a companion could be detected (at the 5 level) in our survey. For the 10 objects in common between the surveys, our SDI survey attains lower minimum separations for 6 out of 10 objects and comparable sepa- rations for two others (we note also that the objects for which we did not attain lower separations were particularly low quality AO/SDI datasets). 74

Figure 2.20 Planet detection completeness contour plots for our program stars. For a given mass and semi-major axis, 10,000 planets are simulated by our Monte Carlo method, over the expected distributions of eccentricity, orbital phase, and viewing angle. Given the parameters of the target star and the COND models of Baraffe et al. (2003), we determine what fraction of the simulated planets are de- tectable at the 5 level given the contrast plot for that star. The contours show this detection probability across the 100,000 different combinations of mass and semi- major axis considered in this plot. The strong upper limit in mass is set by our conservative <1400 K limit for the methane break required for a robust SDI detec- tion. In these models, we simply do not allow an object with Teff >1400 K to be detected, when in reality SDI can detect such non-methane objects (e.g. AB Dor C, Close et al. 2005b, Nielsen et al. 2005). For a complete set of planet detection com- pleteness contour plots, see: http://coatlicue.as.arizona.edu/ bbiller/SDI.html. 75

Figure 2.20 (continued) 76

Figure 2.20 (continued) 77

Figure 2.20 (continued) 78

Figure 2.20 (continued) 79

Figure 2.20 (continued) 80

Figure 2.20 (continued) 81

Figure 2.20 (continued) 82

Figure 2.20 (continued) 83

Figure 2.21 Our 50% completeness levels. Combining the results of Fig. 2.20, we consider individual values of the semi-major axis across the planetary mass range, and at each combination calculate the total number of stars in our survey (out of a total of 54) where the fraction of such planets, given by the Monte Carlo Simulation, that can be detected at the 5 level is 50% or greater. Clearly, our survey is best able to place constraints on planets between 4 and 8 MJup, and with semi- major axis between 20 and 40 AU. The decrease in sensitivity for masses >8 MJup is due to the fact that such high mass planets are too hot to possess significant methane absorption if they are very young and, thus, are not ideal SDI targets. The slightly higher completeness for 4-8 MJup planets for semi-major axis of 30 AU vs. semi-major axis of 40 AU is due to the small field of view of the SDI device; planets with semi-major axes > 30 AU can fall outside the SDI field in some of these cases. 84

Figure 2.22 A complete reduced dataset of AB Dor A (28 minutes of data at a series of rotator angles – 0, 33, 33, 0) from the VLT SDI device. Simulated planets have been added every 0.2” from the star (0.4”, 0.6”, 0.8”, 1.0”, 1.2”, 1.4”, 1.6”, 1.8”, 2.0”, and 2.2”) with F1(1.575m) = 9 mag (upper left, attenuation in magnitudes in the 1.575 m F1 filter), 10 mag (upper right), 11 mag (lower left) and 12 mag (lower right) fainter than the star. The 0.4” object falls within the inner dark circle (dark circle radius of 0.5”, 0.5”, 0.7”, and 1.3” respectively for the 9, 10, 11, and 12 mag objects); the 2.2” object falls outside the frame aperture in a number of dither images and thus is detected with lower S/N than the other objects. These simulated planets are scaled from unsaturated images of AB Dor A taken right before the example dataset (and have fluxes and photon-noise in each filter appropriate for a T6 object). 85

Figure 2.23 Maximum achievable planet contrast (5 detection) vs. separation for 28 minutes of VLT SDI data for AB Dor A. To determine the maximum achiev- able planet contrast as a function of separation, we inserted and then attempted to retrieve simulated planets with a variety of separations and F1 contrasts ap- propriate for T5, T6, and T8 spectral types. The residual SDI noise curve for AB Dor A is also overplotted; the two curves agree well, giving us confidence in our measured contrast limits. 86

Figure 2.24 Maximum achievable H band planet contrast (5 detection) vs. sepa- ration for 28 minutes of VLT SDI data for AB Dor A. To determine the maximum achievable planet contrast as a function of separation, we inserted and then at- tempted to retrieve simulated planets with a variety of separations and F1 con- trasts appropriate for T5, T6, and T8 spectral types. F1 contrasts were converted to H magnitudes using the magnitude offsets calculated in section 2.3.3. 87

Figure 2.25 Minimum detectable planet mass (5 detection) vs. separation (AU) for 28 minutes of VLT SDI data for AB Dor A. To determine the minimum de- tectable planet mass as a function of separation, we inserted and then attempted to retrieve simulated planets with a variety of separations and F1 contrasts ap- propriate for T5, T6, and T8 spectral types. F1 contrasts were converted to H magnitudes using the magnitude offsets calculated in section 3.3 and were then converted to absolute H magnitudes using the 2MASS apparent H mag- nitude and the Hipparcos distance for each star. Absolute H magnitudes were converted into planet masses using the COND models of Baraffe et al. (2003) and adopting a range of system ages from 50 - 100 Myr. For AB Dor, we should be able to image (5 detection) a 7 MJup planet 12 AU from the star. For a complete set of minimum detectable planet mass vs. separation curves, see: http://coatlicue.as.arizona.edu/ bbiller/SDI.html. 88

Figure 2.26 Comparison of SDI contrast curve with other methods. The Lyot curve is for the 3.6m AEOS telescope (Hinkley et al. 2007) and the NICMOS curve coronagraph curve is from HST (Schneider et al. 2003); otherwise curves are all from 8m class telescopes. We use the LQ Hya contrast curve from Mas- ciadri et al. (2005) because this star (K2, 18 pc vs. K1, 15 pc) is the closest match from that work to AB Dor A (our SDI comparison star.) The SDI contrast curve has been converted from F1=1.575m to H contrasts appropriate for T8 and Y spectral type objects. Inside 0.4”, SDI contrasts are derived from the 1-trajectory SDI contrast plot of AB Dor A; outside of 0.4”, SDI contrasts are derived from our in-depth planet simulation case study of AB Dor A. For methanated compan- ions, SDI provides improved contrast by 1-4 mag within 100 as compared to other methods. Past 100, narrowband imaging becomes less efficient and broad-band techniques (such as ADI; Marois et al. 2006) reach higher contrasts. 89

Figure 2.27 Minimum detectable mass vs. separation for our program stars. We convert our contrast curves in mag units (from Figs. 2.11 to 2.15) into minimum detectable mass vs. separation (in AU) using the models of Baraffe et al. (2003) and the distance to the star. We simulated an ensemble of 106 possible planets per star, assuming distributions for mass, eccentricity, and semi-major axis based on known radial velocity planets. When combined with the properties of the individual target star and its measured contrast curve, we can determine what fraction of these simulated planets we expect to detect at the 5 level (shown above each plot). The ensemble of simulated planets is shown as small dots for each star; simulated planets which are detected with the contrast attained by SDI are plotted in blue and those that remain undetected are plotted in red. Assum- ing each star possesses exactly one planet, we assign a detection probability for that star from the percentage of simulated planets detected. For our 48 program stars with contrast curves, the average detection probability is is 4.6%, the median detection probability is 3.5%, and the maximum detection probability is 33%. For GJ 799B (12 Myr M star at 10 pc), we can detect (at 5) a 7.8 MJup planet at 2 AU. 90

Figure 2.27 (continued) 91

Figure 2.27 (continued) 92

Figure 2.27 (continued) 93

Figure 2.27 (continued) 94

Figure 2.27 (continued) 95

Figure 2.27 (continued) 96

Figure 2.27 (continued) 97

Figure 2.27 (continued) 98

Figure 2.28 Expected number of planets detected. By taking the results of our Monte Carlo simulations, and assuming that each program star possesses exactly one planet, we can assign a detection probability for that star from the percentage of simulated planets detected. By adding these detection fractions for each star, we can compute the expected number of planets detected from our survey. We order the target stars by decreasing detection probability, and plot the total num- ber of planets expected to be detected as a function of the number of stars. Over the entire survey, we expect to detect 2.73 planets. Thus, using Poisson statis- tics, our assumed distribution for the frequency (1 planet per star, hence 100%), semimajor axis distribution (N(a) constant), and luminosities (from the COND ∝ models of Baraffe et al. 2003) of extrasolar planets is excluded at the 93% level by our extrasolar planet survey null result. 99 ) C, ) epoch epoch up up candidate candidate Dor J J 2005 M binary M AB second second 6 6 al. binary planet planet at at binary et 0.15” 7” Comments 2.4” binary AU, AU, Close poss. (14 (4.8 tentative tentative detected detected 0.16” not not very very % % % % % % % % % % % % % % % 0 11 2.0 9.6 2.4 7.2 2.0 0.2 2.0 9.0 3.4 8.7 4.0 8.4 5.7 Detectability d 7 H 6.2 6.6 5.9 6.9 6.9 6.6 6.5 4.8 6.5 7.2 6.4 6.5 5.3 4.2 7.1 c 8 8 6 V 10 10 10 10 7.5 8.5 9.3 8.1 8.3 6.9 9.9 8.1 9.5 yr 2005 Stars b al. mg mg mg LH98 Pic Pic Pic Ref FMS97 et WSH03 Her/L uc T Dor Dor Dor om om uc/Hor uc/Hor uc/Hor uc/Hor om Beta Beta Beta Age fr T T T T fr fr AB AB AB Li Li Li Possible Survey Lowrance b SDI 0.1 0.1 0.03 0.03 0.03 0.03 0.07 0.16 0.07 0.07 0.03 0.012 0.115 0.016 0.012 0.012 of Age(Gyr) f e j n m l m e e p e f h k o e g e/G5V M0 SpT K0V K3V K4V K1V K1V G0V G6V K1III K6/7 M1V M0/1 M2.5e G5IV/V F8/GOV G3V operties a Pr 41 14.9 23.9 26.7 38.5 13.5 19.8 16.8 42.4 33.8 40.5 10.4 26.3 52.4 37.1 49.3 2.1. Distance(pc) 00.7 05.4 06.1 43.4 26.5 18.4 54.9 52.9 41.4 05.0 39.0 11.0 25.5 20.0 19.2 50.7 able 47 54 47 56 15 59 T 26 57 02 54 28 46 15 26 38 28 DEC -65 -47 -72 -21 -63 -12 -57 -03 -52 -57 +01 +20 +23 +01 +46 +05 a 44.8 56.8 28.2 34.8 48.9 18.9 26.1 06.5 44.8 13.5 32.1 25.9 47.1 08.1 08.7 21.2 RA 28 36 18 59 57 41 18 24 33 33 52 41 00 19 28 23 05 05 06 04 01 04 00 03 05 03 02 02 05 06 01 01 Stars get oung ar T Y 9054 8558 HD HD 378 au / / 062 T PerA 9141 1481 23309 30030 17925 Men Pic Dor +05 207.1 182 ERX8 HIP V834 GJ ERX6 V577 HD LH98 UY AB BD Nearby HIP GJ AO HIP HIP 100 2004 ) ) ) epoch epoch epoch up up up candidate candidate candidate J J J 2005 M M M , al. companion, 7 8 second second second 11 et binary binary binary planet planet planet at at at Hillenbrand binary mass & Comments AU, AU, 0.78” 0.14” 0.26” AU, Chauvin tentative tentative tentative detected detected detected (15.5 (2.6 (10.4 planetary not not not Metchev very very very % % % % % % % % % % % % % % % % 33 14 0.4 5.2 5.4 1.8 8.0 3.0 0.0 8.9 0.0 0.8 0.1 9.3 0.8 1.5 Detectability d 6 5 H 8.1 5.8 7.1 5.2 4.7 5.8 5.2 7.5 7.9 4.9 6.6 9.6 4.3 7.8 5.6 6.3 5.4 6.5 c 7 V 14 11 13 14 11 9.1 9.1 6.5 6.2 7.8 7.6 7.2 6.4 9.4 5.6 7.8 8.7 6.8 8.7 04 2001 b right+ yra yra al. NB98 Pic Pic Ref WSH03 WSH03 WSH03 W Dor Dor Dor Dor uc et T AKCWM95 AKCWM95 om om om om AB AB AB AB om Beta Beta Age fr Her/L Her/L fr fr fr TW fr om om Li fr fr Li Li Li Montes Li Li RHK b 0.1 0.03 0.07 0.07 0.21 0.01 0.07 0.01 0.01 0.01 0.07 0.035 0.013 0.115 0.115 0.016 0.002 0.012 0.012 0.0032 Age(Gyr) 2.1—Continued m e q m e r q e q q q q t v g g u w x u able K2 G0 G0 M5 M0 M1 T SpT K2V K2V K4V K0V K0V K1V K5V G0V G5V G1V M4.5 G1.5V G1/G2V G5IV+KOIV/V a 22 30 30 50 14.3 45.5 33.9 40.9 46.7 17.7 18.3 21.7 44.1 21.6 41.6 29.4 25.2 21.7 23.5 24.2 Distance(pc) 15.3 30.0 18.7 50.7 48.0 02.7 06.2 16.0 55.7 39.6 04.7 42.0 28.0 51.4 00.2 00.0 04.0 07.1 09.4 36.0 01 17 59 48 14 12 57 03 20 46 11 48 54 03 32 57 57 13 03 48 DEC -58 -30 -24 -11 -39 -53 -29 -61 -66 -66 -60 -75 -78 +65 +64 +26 +35 +23 +76 +29 a 11.7 12.9 47.6 05.3 00.2 25.6 43.7 30.8 26.9 32.4 28.3 00.4 55.8 56.3 21.4 27.7 25.5 30.9 47.3 56.0 RA 39 19 45 22 39 32 32 15 17 12 43 38 36 07 09 17 17 22 24 31 08 06 15 11 14 09 09 12 10 11 10 06 08 15 18 17 17 06 12 12 45270 get 1978 HD ar / 4 25 22 T Lup 92945 135363 48189A 155555AB 155555C 166435 Leo Hya Pic +23 UMa Dra J1224.8-7503 J1231.9-7848 A A A 417 1 HD EK BD LQ DX KW HD TW HD HD RX AB HD HD RX TW SRX1 TW GJ 101 ) ) ) epoch epoch epoch up candidate candidate candidate up up J 2007 2007 2007, J J M M M al. al. al. second second second 13 binary 3 6 binary et planet planet planet et et at at at Comments 0.18” AU, 0.068” AU, AU, Janson Geißler Kellner (3 (7 tentative tentative tentative detected detected detected (6.2 not not not very very very % % % % % % % % % % % % % % % % 20 12 23 23 6.8 3.5 5.5 0.0 2.6 0.2 0.0 0.0 0.0 0.0 0.0 0.0 Detectability d 6 5 H — 8.7 6.3 5.2 7.2 2.5 4.3 4.8 3.3 6.3 5.6 8.7 1.9 5.3 2.3 c 7 V 13 10 13 11 12 13 8.2 5.2 4.8 8.8 4.8 7.7 6.5 7.5 4.7 04 06 06 03 1999 2006 b al. right+ al. Gray+06 Pic Pic Pic Pic Pic Pic Gray+ Gray+ Gray+ Ref et RGP93, SUZT06 WSH03, W Dor et Hydra AKCWM95 om om om om om AB om om om Beta Beta Beta Beta Beta Beta Age fr fr fr fr fr TW fr fr fr om fr Li Li Li Benedict RHK Li RHK RHK RHK Lachaume RHK b 1 0.2 0.8 2.5 4.2 2.0 1.3 0.16 0.07 0.01 0.63 0.012 0.012 0.012 0.012 0.012 0.012 0.008 Age(Gyr) v e n n n n n q v v i g e s u e x m q K0 K7 G0 M0 M0 SpT M4e K2V K3V G0V G3V K5V M0V M4.5e M4.5e K4.5V A5IV/V G1/G2V a 2.1—Continued 20 30 150 3.22 3.63 66.7 9.94 10.2 10.2 32.6 3.48 16.6 30.3 20.5 29.2 22.1 23.6 able Distance(pc) T 52.0 57.9 47.5 54.6 29.7 09.5 43.0 27.1 09.0 06.6 10.9 00.5 44.6 08.4 16.5 01.0 07.8 21 44 44 14 27 47 23 20 26 26 48 59 51 03 52 15 34 DEC -09 -56 -45 -31 -32 -32 -37 -34 -64 -39 -64 -33 -78 +20 +38 +09 +04 a 55.8 21.7 26.5 09.5 51.1 51.2 47.1 53.9 03.2 00.6 29.8 57.3 37.0 10.7 26.9 57.8 36.7 RA 32 03 13 45 41 41 23 06 12 36 21 45 45 56 45 44 43 03 22 11 20 20 20 09 21 13 14 19 19 18 23 18 22 12 planets Stars V R oung get Analogues Y ar T known Solar Distant with e 1208 14 A 112312A 1243.6-7834 81040 201091 172555A 114613 128311 181321 186704 224228 A 799A 799B 803 Ind Eri HD GJ GJ RXJ GJ HD CD-64 HD Mor TW Stars Nearby HD HD HD HD HIP HD 102 be our fect average af would an it or Comments hence stars, ages, and significantly these stars, RHK % of not 0.0 K5 ages 1995, does Detectability two subtraction), calcium the age d these for H Mundt 5.3 ages, the of in and c or case values V 7.7 photospheric lithium err the the 1995, ilson in W ecise ichmann, 06 either 2006, 1998, specific pr e W but al. , Mt. b ar 1988, the et of Sciortino e Gray+ adopt Ref so specifically ages earlier Brandner Covino, we om and Age fr or Beuzit, 1984; stars, and K1 es, 1978, al. calibration RHK 2001, Micela, Smith-Moor Schmitt, orr et otherwise, these , T Although , The types 1967, b Houk and 1982, 1988, itt o: 1 ound Neuhauser ages. Cornide W (2004), (Noyes Barbera, ar Zucker 6.3 Krautter Houk , spectral Houk 1991, and t: and Age(Gyr) , Cowley Song 2.1—Continued NB98: v: o, useful detect eiß Udry with Favata, Alcala, n to 2001, determination. and l: and e Jahr 1998, Castr 1995, yield Hiltner stars well-defined , SpT age K5V to De able faint Hu Sozetti, 1960, and Houk for T for Becklin f: oa, been too a Mundt al.(1997) and e ely Cowley ages ages Pagel AKCWM95: r et and Li Zuckerman not ar i: Gliesse and unlikely 1975, to 1995, SUZT06: is n: and om has Catalog have 15.4 fr R’HK ences: 1998, LH98: 1993, 2003 planets which Ca stars Distance(pc) 2001, (2003) mag) stars Cowley efer Perryman Hu Schwartz, Rodono r ichmann, , Leaton on Draper K5 1997, al. W 1.1 ebb k: Fernandez-Figuer and and age ebb, and et W these for Hubrig , W Li 06.4 survey simulated ement Henry Pallavicini 2002, memberships h: and that and 01 Houk (B-V Cutri determined DEC x: cos Covino, all Sciortino Galvez, , e: calcium elation and service -30 r Kuerster 2004, measur only and stars cluster 1979, simply and key: Bessell, age Hippar Survey almost Schmitt, K5 Zuckerman, Schmitt, eiß Song have , a s: Simbad vs. with R’HK Gratton, 15.2 the state ence for Mathioudakis Song, Micela, Jahr RA esult, we to and 29 lithium r the om Pallavicini, efer 2MASS CDS r stars a 1951, fr 22 and R’HK Lopez-Santiago, ichmann, for and R’HK Krautter a the As the for only W Randich, to Favata, . general, esults. accurate om om r get e Gyr 862 ages fr Spectral In derived onyms fr have Zuckerman, Gliesse Alcala, Cutispoto, both. Montes, Zuckerman 1 a b c d 1 Bidelman ar Christian T GJ of Acr m: FMS97: g: j: we S-index p: q: applying r: > mor u: final RGP93: w: WSH03: 103

Table 2.2. VLT SDI Observation Log

Object Date DIT NDIT Total Exp (minutes)

HIP 1481 2004-11-15 14 6 56 2005-11-24 16 5 26.7 2005-11-25 16 5 26.7 2005-11-27 16 5 26.7 ERX6 / HD 8558 2004-11-14 22 4 29.3 2004-11-16 22 4 58.7 ERX8 / HD 9054 2004-11-17 22 4 58.7 HIP9141 2004-09-27 14 6 56 BD +05 378 2005-02-01 32 3 25.6 HD 17925 2003-08-14 7.5 16 40 2003-08-16 4 30 40 2003-08-17 4 30 20 2004-02-02 1 120 20 2004-11-16 4.1 17 46.5 2004-11-17 4.1 17 46.5 LH98 062 2004-02-03 14 9 21 Eri 2004-09-19 0.6 160 64 V834 Tau 2005-01-25 10 9 24 2005-02-01 10 9 24 GJ 182 2004-02-02 7 17 39.7 104

Table 2.2—Continued

Object Date DIT NDIT Total Exp (minutes)

2005-11-22 20 4 26.7 2005-11-24 20 4 26.7 2005-11-27 20 4 26.7 HIP23309 2005-01-30 24 4 25.6 2005-01-31 24 4 51.2 AB Dor 2004-02-02 5 24 20 2004-09-28 12 7 28 2004-11-16 10.4 8 27.7 GJ 207.1 2005-01-27 32 3 25.6 UY Pic 2004-11-16 14 6 28 2004-11-17 14 6 56 AO Men 2004-02-03 14 9 21 2005-11-15 30 1 17.5 2005-11-24 30 1 10 AB Pic 2004-11-14 20 4 26.7 2004-11-15 20 4 26.7 2005-11-22 20 4 13.3 2005-11-25 20 4 53.3 SRX1 / HD 45270 2004-11-18 12 7 28 2004-11-19 12 7 28 105

Table 2.2—Continued

Object Date DIT NDIT Total Exp (minutes)

HD 48189A 2004-11-17 6.5 11 23.8 2004-11-18 6.5 11 23.8 BD +23 1978 2005-01-27 24 4 25.6 2005-01-28 24 4 25.6 LQ Hya 2004-02-02 5 24 40 2004-12-08 14 6 28 2004-12-14 14 6 28 DX Leo 2004-02-05 3 38 19 2005-12-04 14 6 28 2005-12-19 14 6 28 TWA 22 2005-01-25 32 1 48.5 HD92945 2004-02-05 5 24 60 TWA 14 2005-01-28 32 3 25.6 2005-01-29 32 3 25.6 TWA 4 2004-02-02 7 17 9.92 TWA25 2005-01-28 32 3 25.6 RX J1224.8-7503 2004-02-02 40 3 20 2005-01-16 30 3 60 2005-01-27 30 3 120 RX J1231.9-7848 2004-02-05 20 6 20 106

Table 2.2—Continued

Object Date DIT NDIT Total Exp (minutes)

RXJ 1243.6-7834 2004-02-02 5 24 40 HD 114613 2004-02-02 1 120 40 KW Lup 2004-09-15 22 4 14.7 2004-09-16 24 4 22.7 2004-09-17 24 4 24 HD155555AB 2003-08-14 7.5 16 10 2003-08-15 7.5 16 20 2003-08-16 7.5 16 10 2003-08-17 7.5 16 10 2004-09-16 10 9 30 2004-09-18 14 6 28 HD155555C 2003-08-14 30 4 40 2003-08-16 30 4 40 HD172555A 2003-08-17 5 24 20 2004-09-17 5 15 25 2004-09-18 5 15 6.25 2004-09-19 5 15 18.8 CD-64 1208 2003-08-17 20 6 40 2004-09-16 15 6 30 HD181321 2003-08-15 7.5 16 40 107

Table 2.2—Continued

Object Date DIT NDIT Total Exp (minutes)

2004-09-18 11 8 29.3 GJ799B 2003-08-16 20 6 40 2003-08-17 20 6 30 2004-09-19 15 6 30 GJ799A 2003-08-16 20 6 40 2004-09-16 10 9 30 2004-09-19 15 6 30 GJ803 2003-08-14 7.5 18 56.2 2003-08-15 10 12 40 2003-08-17 7.5 16 40 2004-09-17 6 15 30 2004-09-18 10 9 30 Ind A 2004-09-18 0.5 192 48 GJ 862 2003-08-15 10 12 40 2003-08-16 10 12 40 2004-09-19 13 7 48.2 HIP112312A 2004-09-19 25 4 66.7 108

Table 2.2—Continued

Object Date DIT NDIT Total Exp (minutes)

HD224228 2003-08-16 10 12 40 2003-08-17 20 6 40 2004-10-08 14 6 28 2004-10-20 21 4 28 109

Table 2.3. MMT SDI Observation Log

Object Date DIT NDIT Total Exp (minutes)

V577 PerA 2006-02-12 20 7 37.3 2006-02-13 21.5 7 40.1 HIP30030 2006-02-12 30 5 30

1 UMa 2006-02-13 5.8 13 40.2 HD 81040 2006-02-12 11.7 13 40.3 LQ Hya 2006-02-12 8 19 40.5 DX Leo 2005-05-01 10 13 34.7 GJ 417 2005-04-30 7 17 31.7 HD 128311 2006-02-12 4 19 60.8 EK Dra 2005-05-01 20 7 37.3 HD 135363 2005-05-01 30 5 40 HD 166435 2005-04-30 7 17 31.7 2005-05-01 7 17 31.7 HD 186704 2005-05-01 10 13 17.3 HD 201091 2005-04-30 20 7 37.33 110

Table 2.4. Magnitude Offsets

Spectral Type Magnitude Offset

M8 -0.12 0.08 T5 0.5 0.05 T6 0.6 0.07 T8 0.87 0.04 111

Table 2.5. Limiting H mag (5) at 0.5”

Object F1 Separation(AU) H (T8 SpT) mH MH

Eri 9.4 0.12 1.61 10.3 12.2 14.7 Ind A 10.6 0.12 1.81 11.5 13.8 16 HD 201091 8.08 0.52 1.74 8.95 11.5 13.8 HD 114613 6.13 0.26 10.2 7 10.3 8.74 HD 17925 9.69 0.14 5.19 10.6 14.8 14.7 HD172555A 9.14 0.12 15 10 14.3 11.9 UMa 8.04 0.15 7.14 8.91 13.2 12.4 1 HD 48189A 8.54 0.052 10.8 9.41 14.2 12.5 GJ803 9.54 0.091 4.97 10.4 15.2 15.2 AB Dor 9.04 0.019 7.47 9.91 14.8 13.9 HD155555AB 5.87 0.14 15 6.74 11.6 9.21 GJ 417 7.79 0.23 10.9 8.66 13.7 12 HD181321 7.42 0.13 10 8.29 13.3 11.8 SRX1 9.95 0.079 11.7 10.8 16 14.1 GJ799A 7.48 0.082 5.11 8.35 13.6 13.6 DX Leo 8.24 0.19 8.87 9.11 14.4 13.2 GJ 862 9.51 0.25 7.72 10.4 15.7 14.8 V834 Tau 9.08 0.18 6.74 9.95 15.3 14.6 HD 166435 8.42 0.17 12.6 9.29 14.7 12.7 LQ Hya 9.82 0.16 9.17 10.7 16.3 15 112

Table 2.5—Continued

Object F1 Separation(AU) H (T8 SpT) mH MH

HD 186704 7.13 0.091 15.1 8 13.6 11.2 HD92945 9.91 0.0099 10.8 10.8 16.6 14.9 UY Pic 9.96 0.11 11.9 10.8 16.7 14.8 HD224228 9 0.15 11 9.87 15.9 14.2 EK Dra 7.85 0.39 17 8.72 14.7 12 HIP 1481 9.22 0.13 20.5 10.1 16.3 13.2 CD-64 1208 9.33 0.087 14.6 10.2 16.5 14.2 HD 135363 7.9 0.27 14.7 8.77 15.1 12.8 HIP23309 8.45 0.092 13.1 9.32 15.7 13.6 GJ 182 8.01 0.16 13.3 8.88 15.3 13.2 V577 PerA 8.9 0.33 16.9 9.77 16.2 13.6 HIP9141 8.92 0.29 21.2 9.79 16.3 13.2 HIP30030 6.91 0.17 26.2 7.78 14.4 10.8 KW Lup 8.76 0.091 20.5 9.63 16.3 13.2 ERX8 9.4 0.2 18.6 10.3 17.2 14.4 ERX6 9.38 0.4 24.6 10.2 17.1 13.6 AO Men 6.91 0.33 19.2 7.78 14.8 11.9 AB Pic 9.65 0.027 22.8 10.5 17.6 14.3 GJ 207.1 7.5 0.094 8.41 8.37 15.5 14.4 HIP112312A 9.09 0.27 11.8 9.96 17.1 15.2 113

Table 2.5—Continued

Object F1 Separation(AU) H (T8 SpT) mH MH

BD +05 378 8.31 0.088 20.3 9.18 16.4 13.4 TWA25 9.5 0.035 22 10.4 17.9 14.7 RX J1224.8-7503 7.16 0.024 12.1 8.03 15.9 14 HD155555C 10.5 0.085 15 11.4 19.3 16.9 TWA 14 8.38 0.03 33.3 9.25 18 13.9 114

Table 2.6. Limiting H mag (5) at 1.0”

Object F1 Separation(AU) H (T8 SpT) mH MH

Eri 11.3 0.2 3.22 12.2 14.1 16.6 Ind A 12 0.16 3.63 12.9 15.2 17.4 HD 201091 9.42 0.05 3.48 10.3 12.8 15.1 HD 114613 7.24 0.13 20.5 8.11 11.5 9.94 HD 17925 11.3 0.19 10.4 12.2 16.4 16.3 HD172555A 11.2 0.098 30 12.1 16.4 14 UMa 9.28 0.14 14.3 10.1 14.4 13.6 1 HD 48189A 9.87 0.24 21.7 10.7 15.4 13.7 GJ803 10.7 0.03 9.94 11.6 16.4 16.4 AB Dor 11 0.17 14.9 11.9 16.7 15.8 HD155555AB 7.3 0.046 30 8.17 13.1 10.7 GJ 417 8.44 0.05 21.7 9.31 14.3 12.6 HD181321 8.63 0.048 20 9.5 14.6 13.1 SRX1 11.2 0.13 23.5 12.1 17.3 15.4 GJ799A 9.55 0.14 10.2 10.4 15.6 15.6 DX Leo 9.98 0.039 17.7 10.8 16 14.8 GJ 862 10.7 0.12 15.4 11.6 16.9 16 V834 Tau 10.2 0.18 13.5 11.1 16.4 15.7 HD 166435 9.98 0.061 25.2 10.8 16.2 14.2 LQ Hya 11 0.035 18.3 11.9 17.5 16.2 115

Table 2.6—Continued

Object F1 Separation(AU) H (T8 SpT) mH MH

HD 186704 7.35 0.052 30.3 8.22 13.8 11.4 HD92945 10.8 0.062 21.6 11.7 17.5 15.8 UY Pic 11.5 0.033 23.9 12.4 18.3 16.4 HD224228 10.8 0.11 22.1 11.7 17.7 16 EK Dra 8.86 0.14 33.9 9.73 15.7 13 HIP 1481 10.8 0.046 41 11.7 17.9 14.8 CD-64 1208 9.88 0.54 29.2 10.8 17.1 14.8 HD 135363 8.65 0.025 29.4 9.52 15.8 13.5 HIP23309 10 0.051 26.3 10.9 17.3 15.2 GJ 182 10.2 0.15 26.7 11.1 17.6 15.5 V577 PerA 10 0.062 33.8 10.9 17.4 14.8 HIP9141 10.5 0.028 42.4 11.4 18 14.9 HIP30030 8.3 0.09 52.4 9.17 15.8 12.2 KW Lup 9.86 0.17 40.9 10.7 17.3 14.2 ERX8 10.7 0.12 37.1 11.6 18.5 15.7 ERX6 10.6 0.12 49.3 11.5 18.4 14.9 AO Men 7.9 0.015 38.5 8.77 15.8 12.9 AB Pic 10.8 0.013 45.5 11.7 18.8 15.5 GJ 207.1 8.74 0.089 16.8 9.61 16.8 15.7 HIP112312A 10.6 0.068 23.6 11.5 18.7 16.8 116

Table 2.6—Continued

Object F1 Separation(AU) H (T8 SpT) mH MH

BD +05 378 9.52 0.074 40.5 10.4 17.6 14.6 TWA25 10.5 0.18 44.1 11.4 18.9 15.7 RX J1224.8-7503 8.04 0.16 24.2 8.91 16.8 14.9 HD155555C 10.8 0.043 30 11.7 19.6 17.2 TWA 14 8.74 0.047 66.7 9.61 18.3 14.2 117 up J M 10 10 and — — — 5.91 5.87 9.24 23.85 5 25.02 15.24 12.28 10.01 10.54 for Separation(AU) up J Separations M 5 — — — — — — — — 20.62 17.87 30.00 24.29 Detectable Planets up Separation(AU) J Minimum M (pc) ojected 3.22 3.63 14.94 33.94 38.48 26.67 17.75 21.72 13.49 40.54 16.82 34.21 Pr Distance (Myr) 70 70 12 12 12 12 800 115 115 160 100 Star/Planet 1300 Age 2.7. 1208 able 378 T A Men -64 Leo Dor Dra 417 207.1 182 174 Object Eri Ind DX BD+05 EK GJ GJ CD AB GJ GJ AO 118 up J M 10 — — — — — — 1.09 1.64 1.74 9.42 3.30 8.30 19.58 Separation(AU) up J M 5 — — — — — — 2.78 3.17 4.70 3.30 13.54 19.52 32.45 Separation(AU) 2.7—Continued (pc) able T 9.94 10.22 10.22 15.45 29.44 10.38 30.03 29.23 20.86 25.24 30.03 16.57 20.48 Distance (Myr) 3 12 12 12 12 12 12 115 160 100 630 6300 4200 Age AB A C 181321 17925 155555 172555 155555 135363 166435 114613 128311 Object 862 799A 799B 803 HD HD HD HD HD HD HD HD HD GJ GJ GJ GJ 119 up J M 10 — — — 9.71 5.51 3.31 9.29 26.44 20.88 12.15 22.18 15.51 129.84 Separation(AU) up J M 5 — — — — — 5.43 59.38 43.14 16.28 47.92 34.92 69.00 51.00 Separation(AU) 2.7—Continued (pc) able T 3.48 40.95 21.57 52.36 26.26 21.67 22.08 23.61 30.26 32.56 37.15 49.29 23.50 Distance (Myr) 30 70 30 12 70 70 12 30 30 70 200 2500 2000 Age 9054 8558 45270 A A HD HD HD / / / Object 8 6 1 1481 30030 112312 23309 92945 224228 186704 81040 48189 201091 HIP HIP HIP HIP ERX ERX HD HD HD HD SRX HD HD 120 up J M contrast 10 — 6.14 2.94 7.40 9.25 4.83 survey 13.20 23.64 12.00 22.45 ent curr Separation(AU) our up J with M 5 — — 8.59 5.32 detected 40.05 28.64 32.00 14.98 78.78 19.34 be to Separation(AU) mass 2.7—Continued in (pc) able low T 40.92 45.52 14.27 23.87 18.34 66.67 44.05 42.35 33.77 24.17 too is Distance object 2 (Myr) 30 70 13 10 10 30 70 16 an 210 Age such star that A that 14 25 Per means 9141 Lup Object for Pic Hya Pic A A UMa — 1 AB HIP TW KW V577 LQ UY RXJ1224.8-7503 TW level 121

Table 2.8. Binary Properties

Object Sep. PA est. F1 Epoch

SDI survey discoveries

a AB Dor AC 0.16” 127 Ks=9.45 0.1 2004-02-02 0.2” 2004-09-28 2004-11-16 HIP 9141 0.15” 355 0.1 2004-09-27 HD 48189AC 0.14” 143 1.5 2004-11-17 HD 135363 0.26” 132 1.7 2005-05-01 CD -64 1208 0.18” 95 2.8 2004-09-16 SDI survey confirmations RXJ 1243.6-7834b 0.068” 171/351 0 2004-02-02 LH 98 062 2.4” 354 0 2004-02-03 TWA 4 0.78” 3 0.05 2004-02-02 EK Dra 0.67” 176 3.0 2005-05-01

aSeparation, position angle, and photometry from Close et al. (2005b). For updated photometry and astrometry see Close et al. (2007b). bAs RXJ 1243.6-7834 is nearly an equal-magnitude binary, we were unable to determine which star was the primary (as selected by Brandner et al. (2000)) and thus present two values for the position angle (assuming each star is the primary in turn). 122

CHAPTER 3

DISCOVERY OF A VERY NEARBY BROWN DWARF TO THE SUN: A METHANE

RICH BROWN DWARF COMPANION TO THE LOW MASS STAR SCR 1845-6357

3.1 Introduction

After decades of little change in the number of known stellar systems within 5 pc, numerous previously unknown low mass stars have recently been discovered in the solar neightborhood (e.g. Scholz et al., 2003; Henry et al., 2004; Hambly et al., 2004). 1 Because they are extremely nearby and intrinsically low luminos- ity, these objects are ideal targets to search for low mass companions, since even a close companion will appear reasonably separated on the sky. For example during commissioning of the Simultaneous Differential Imager (SDI) at the VLT (Close et al., 2005a; Lenzen et al., 2004), one such object, Indi B was resolved into a binary T dwarf (McCaughrean et al., 2004). Indi Ba,Bb are the closest brown dwarfs to the Earth (3.626 0.009 pc). At a distance of 3.85 0.02 pc (Henry et al., 2006), the recently discovered M8.5 star SCR 1845-6357 is just slightly fur- ther away than Indi Ba,Bb (Hambly et al., 2004) and is the 24th closest stellar system from the Earth. We report the discovery of SCR 1845-6357B (hereafter SCR1845B), a methane rich substellar companion to this star. This companion ob- ject is the third closest brown dwarf to the Sun. It is also the only example of a T dwarf companion to a low mass star and one of the tightest known brown dwarf companions to a star.

1This work first appeared as B. Biller, M. Kasper, L. Close, W. Brander, and S. Kellner, 2006, The Astrophysical Journal, 641, L141. Reproduced by permission of the AAS. 123

3.2 Observations and Data Reduction

Data were taken on the night of 2005 May 28 (UT)2 at the 8.2m VLT-UT4 with the unique Simultaneous Differential Imager (SDI) in the facility AO system NACO (Lenzen et al., 2003; Rousset et al., 2003). To guide on this faint red object, the infrared wavefront sensor (IRWFS) was used with the “K” dichroic, sending all of the K band light to the WFS. SDI can be used to calibrate and remove the “speckle noise” in AO images, while also isolating the light from a substellar methane companion from the starlight. This method was pioneered by Racine et al. (1999), Marois et al. (2000), Marois et al. (2002), and Marois et al. (2005). It

exploits the fact that all cool (Teff <1200 K) substellar objects have strong CH4 (methane) absorption redwards of 1.62 m in the H band infrared atmospheric window (Burrows et al., 2001, 2003). The NACO SDI device obtains four images of a star simultaneously through three slightly different narrowband filters (sam-

pling both inside and outside of the CH4 features – Close et al., 2005a; Lenzen et al., 2004). These images are then differenced. This subtracts out the halo and speckles from the bright star to reveal any substellar methane objects orbiting that star. Since a substellar methane object will be brightest in one filter and ab- sorbed in the rest, while the star is bright in all three, a difference can be chosen which subtracts out the star’s light and reveals the companion’s light. Thus, SDI also helps eliminate the large contrast difference between the star and substellar companions (Close et al., 2005a; Lenzen et al., 2004, 2005). The SDI device has already produced a number of important scientific results: the discovery of AB Dor C (Close et al., 2005b) – the faintest companion ever discovered within 0.16” of a star, detailed surface maps of Titan (Hartung et al., 2004), the discovery of the

2Based on observations collected at the European Southern Observatory, Paranal, Chile through proposal 075.C-0357(A) 124

binarity of Indi Ba-Bb, the nearest binary brown dwarf (Scholz et al., 2003; Mc- Caughrean et al., 2004), and evidence of orbital motion for Gl 86B, the first known white dwarf companion to an exoplanet host star (Mugrauer & Neuhauser,¨ 2005). Using the SDI device provides a marked advantage over single band imaging even in situations where the contrast difference between star and companion is not large – images in the 3 different SDI filters immediately provide spectral in- formation about any substellar candidate, particularly regarding the amount of

CH4 present. SCR 1845 was observed for 15 minutes (with 3 30 s subimages taken at 5 different dither positions) at a position angle of 0 and 15 minutes at a position angle of 22. A base integration time (DIT) of 30 s was used and subimages were medianed. Observing the object at different roll angles allows us to immediately confirm if an object is real – an instrumental feature should not rotate with a change of rotator angle; however, a real object on the sky should appear to rotate by the change in rotator angle. Data were sky-subtracted, flat-fielded, and bad- pixel masked. Each data frame was then aligned to a master frame using the IRAF task xreg. After alignment, all frames were median combined. As a comparison, the data were also reduced using a custom IDL SDI pipeline which performs basic data reduction tasks and also precisely aligns images taken in each of the filters using a custom shift and subtract routine (Biller et al., 2006a).

3.3 Results and Discussion

Reduced data for the 0 dataset are presented in Fig. 3.1. SCR 1845B appears at a separation of 1.170” 0.003” and at a position angle of 170.20 0.13 from the M8.5 primary in all four of the SDI filters and rotates by 22 (as expected) between datasets. A three color image generated from the SDI filter images is presented 125

in Fig. 3.2. For comparison, an image reduced using the SDI pipeline (Biller et al., 2006a) is also presented in Fig. 3.2. While SCR 1845B is far from the primary and easily detected, we would also be capable of detecting similar or lower mass companions closer to the primary for this system (down to 0.1” separations). Deacon et al. (2005) measured a trigonometric parallax to SCR 1845-6357 of 282 23 mas, corresponding to a distance of 3.5 0.3 pc. With additional epochs of observation, Henry et al. (2006) provide an updated distance of 3.85 0.02 pc. Thus, the candidate lies 4.50 0.02 AU from its primary. The candidate object is 3.57 0.057 mag fainter than the primary in the F1(1.575 m) filter (all photometry performed with the IRAF DAOPHOT PSF fitting package).

3.3.1 Spectral Type

The candidate object appears brightest in the F1(1.575 m) filter, slightly fainter in the F2(1.6 m), and then drops by a factor of 2.7 0.1 between the F1(1.575 m) and F3(1.625 m) filters. The spectral signature of this dropoff is consistent with methane absorption in the atmosphere of a substellar object (Geballe et al., 2002). Previous observations of the T6 spectral type brown dwarf Indi Bb (McCaugh- rean et al., 2004) with the SDI device found that the flux of Indi Bb also dropped by a similar factor between the F1(1.575 m) and F3(1.625 m) filters (see Fig. 2.5). To determine an accurate spectral type for SCR 1845B, we define an SDI methane spectral index calculated from our SDI F1(1.575 m) and F3(1.625 m) filter im- ages (similar in concept to the methane spectral index defined by Geballe et al., 2002). The SDI device measures the location and strength of the 1.6 m methane absorption break, which is a principle spectral feature used to determine spectral types for T dwarfs – this SDI methane index should be sufficient to estimate an accurate spectral type for this object. The SDI methane spectral index is defined 126

2 SF 1()d F 1 R1 as: index( ) = 4 F 3 SF 3()d R3 Each SDI filter was manufactured by Barr Associates to have a precise band- width of 0.025 m, so the wavelength intervals in the numerator and denomina- tor have the same length for the SDI methane index. We only possess SDI data on a limited number of T dwarfs (this object, Gl 229B, Indi Ba (T1), Indi Bb (T6)). In order to compare SCR 1845B to a wider range of L and T dwarf objects we calculated these same SDI spectral indices from spectra of 56 L dwarfs and 35 T dwarfs (Knapp et al., 2004). Spectra for these objects were obtained from Sandy Leggett’s L and T dwarf archive3. In order to make an accurate comparision, SDI filter transmission curves were con- volved into these calculations. Since we have full spectral data for these objects, we also calculated the 1.6 m methane spectral index defined by Geballe et al. (2002), which were found to be similar to our SDI methane spectral indices. In Fig. 2.5, SDI methane spectral indices are plotted for SCR 1845B, the T dwarfs Gl 229B, Indi Ba, Indi Bb, and 94 other L and T dwarfs. SCR 1845B appears to have a noticeable methane break with somewhat lower indices than the T6 dwarfs Gl 229B and Indi Bb. However, Geballe et al. (2002) note that Gl 229B has an anomalously high methane index for its spectral type and assign a large uncertainty to Gl 229B’s spectral type – T6 1. For our SDI methane indices, SCR 1845B has spectral indices similar to that of T4.5-T6.5 dwarfs. Thus, we determine an initial spectral type of T5.5 1 for SCR 1845B. 3.3.2 H magnitude

To determine an accurate H magnitude, the spectra of both the primary and sec- ondary components of SCR 1845 must be taken into account. The M8.5 primary is extremely red – and will appear brighter in the H band than in our blue F1 band.

3http://www.jach.hawaii.edu/ skl/LTdata.html 127

Additionally, the T5.5 1 companion is blue compared to the primary and will ap- pear brighter in the F1 band than in the H band. To convert from our F1 filter mag- nitudes into calibrated H band magnitudes we must calculate the H band magni- tude offsets for the M8.5 primary star and the T4.5-T6.5 companion (OffsetM and

OffsetT respectively): H = HT HM = (OffsetT + F 1T ) (OffsetM + F 1M ) = (OffsetT OffsetM ) + F 1 Using the spectrum of the star VB10 (an M8 template, and thus a reasonable approximation of an M8.5 spectrum), an H transmission curve, and our F1 filter transmission curve, we calculate a magnitude offset of OffsetM =-0.12 0.08 mag. Assuming spectral types of T4.5-T6.5, we can perform a similar calculation for the companion. Offsets were calculated for 15 objects with spectral types of T4.5- T7 (spectra from Knapp et al., 2004), then averaged together by spectral type to derive an average offset for each spectral type. For instance, for a T5 companion,

OffsetT = 0.5 0.05. For a T6 companion, OffsetT = 0.6 0.07. Magnitudes in the 5 6 H filter for both primary and candidate object are presented in Table 3.1. Uncer- tainties are provided for a companion spectral type of T5.5 1. Background T4-T7 dwarfs possess absolute H magnitudes of 14.5-16.0 (Burgasser et al., 2003), so +0.31 our calculated absolute H magnitude of 15.230.26 for SCR 1845B is quite reason- able.

3.3.3 Likelihood of Being a Bound Companion and T Dwarf Number Densities

This object has not been observed at multiple epochs with the SDI device so we must consult other sources to determine if it is indeed truly bound, i.e. shares a common- with its primary. SCR 1845 possesses a large proper motion of 2.5”/ year. A bound companion would possess a similar proper mo- tion whereas a background object would appear to stay in the same spot on the sky. On 2000 January 1, SCR 1845A had an RA of 18h45m05.2” and DEC of - 128

6357’47.355” (J2000, Deacon et al., 2005). Taking proper motion into account, during the 2005 May 28 SDI observations, SCR 1845A therefore had an RA of 18h45m07.21” and DEC of -6357’43.586”. SCR 1845B (1.17” separation at a PA of 170.2) had an RA of 18h45m07.33” and DEC of -6357’42.786” during the 2005 May 28 SDI observations. If the faint companion is actually a background T dwarf, this RA and DEC should be reasonably correct for other epochs of ob- servation. Checking the 2MASS point source catalog (2MASS images taken 2000 May 29), we found no objects within 20” of this position and no objects with T dwarf colors in this part of the sky. Hence, it is impossible that this object is a background T-dwarf and it is highly likely to be a bound companion. In the last few years, 3 new T dwarf companions have been discovered within 6 pc of the Sun –SCR 1845B and Indi Ba-Bb. All three of the nearest T dwarfs are bound companions to stars. Combining these three new T dwarf compan- ions with Gl 229B (5.8 pc) and Gl 570D (5.9 pc), we find a number density of T dwarf companions of 5.5 103 pc3 within 6 pc of the Sun. In contrast, only two isolated, field T dwarfs (2MASS J0415 and 2MASS J0937 from Vrba et al., 2004) are known within 6 pc of the sun, leading to a rough number density of field T dwarfs of 2 103pc3 in this volume. Granted, this number of field dwarfs is somewhat incomplete since there may be nearby T dwarfs without accurate trigonometric parallaxes, but the number is unlikely to change by more than a factor of 2 since this population of bright T-dwarfs is very well studied. The ex- istence of 5 brown dwarfs within binary systems < 6 pc of the Sun suggests that the number density of T dwarfs in binary systems may be higher than that of isolated, single T dwarfs. This may be difficult to expain with “ejection” theories of brown dwarf formation (Reipurth & Clarke, 2002). 129

3.3.4 Mass Estimate for SCR 1845B

While the distance to SCR 1845 is well known, the age of the system is uncon- strained. Ages between 100 Myr and 10 Gyr are all plausible for this system at this time. Future observations of lithium absorption might constrain the age of this system or at least rule out very young ages. However, with a vtan=41 km/s it is unlikely that this system is very young. Using the Baraffe et al. (2003) COND models with this age range, an absolute H mag range of 15.1 to 15.6, and spectral

types of T4.5-T6.5 (Teff 850 K 100 K), we find a mass range of 9 - 65 MJup. While SCR 1845B is clearly substellar at any age, the uncertainty in the age of this system means that we cannot derive an unambiguous mass for this object from the COND models. However, since this object is so close to its primary (currently 4.5 AU), orbital motion should be evident within a few years. Both the primary and secondary mass can be measured accurately within a decade, making SCR 1845B a key T-dwarf mass-luminosity calibrator.

3.4 Conclusions

SCR 1845B is the brightest mid-T dwarf yet discovered. In addition, it is the first T dwarf companion found around a low mass star. At only 4.5 AU from its primary, it is one of the tightest known brown dwarf companions to a star and is a further piece of evidence that the brown dwarf desert does not exist for companions to very low mass stars (Close et al. , 2003; Gizis et al., 2003). Both the primary and secondary mass can be accurately measured within a decade. 130

Table 3.1. SCR 1845 Photometry

Spectral Type H mag (primary) F1 H (companion) absolute H

+0.31 +0.31 T5.5 1 8.967 0.027 3.57 0.057 13.16 15.30 0.26 0.26

from 2MASS point source catalog 131

Figure 3.1 An SDI image of SCR 1845. This 15 minute long image was taken at a position angle of 0 and was reduced using a custom IDL pipeline and the IRAF xreg tool. A substellar companion appears at a separation of 1.170” 0.003” (4.5 AU at 3.85 pc) from the primary and a position angle of 170.20 0.13 in each of the 4 SDI filters. The platescale is (0.01725” 0.00025”)/pix (Nielsen et al., 2005). The companion appears brightest in the F1 filter (out of the CH4 absorption) and drops by a factor of 2.7 in the F3 filter (inside the CH4 absorption), consistent with a T5.5 dwarf spectral type. North is up and east is to the left. Upper Right Inset: Three color image of SCR 1845 A and B generated from the SDI filter im- ages (blue=1.575 m, green=1.600 m, red=1.625 m). The substellar companion appears blue in this image. This image was created with a log10 stretch and each filter is equally weighted. Note how similar in color (white) each of the PSF speckles are for the M8.5, while the faint companion SCR 1845B is considerably bluer due to strong CH4 absorption. The structure in the PSF is typical of the NACO IR WFS for a faint guide star such as SCR 1845A. 132

Figure 3.2 Images of SCR 1845 using the SDI device and reduced using a custom SDI pipeline (Biller et al., 2006a). This 30 minute long image was taken at position angles of 0 (white) and 22 (black). Datasets from each roll angle were subtracted from each other and smoothed with a 1 pixel FWHM gaussian. A substellar com- panion appears at a separation of 1.17” from the primary in each of the 4 SDI filters. Note that the speckles from the M8.5 are almost totally removed. With the high contrasts achievable by SDI, a methane object like SCR 1845B (H=4.2 mag) could have been detected at 10 10 closer in at a separation of only 0.1”. 133

CHAPTER 4

HIGH RESOLUTION MID - INFRARED IMAGING OF THE AGB STAR RV BOO

WITH THE STEWARD OBSERVATORY ADAPTIVE OPTICS SYSTEM

4.1 Introduction / Application to Planet-finding Science

In the second half of this thesis, I discuss the development of technology and techniques that may someday be used to detect earthlike extrasolar planets. Us- ing the unique adaptive secondary mirror AO system at the 6.5m MMT (Wildi et al. 2003, Brusa et al. 2003), it is possible to achieve nearly perfect (Strehl ratio 0.97 0.03), high resolution ( 0.100) images at mid-IR wavelengths. In the current work, we used this capability to probe asymptotic giant branch star and proto-planetary nebulae morphologies on finer scales than ever before possible in the mid-IR. Through deconvolution, these extremely high fidelity images pro- duced with AO at the MMT allow resolutions better than that of the FWHM of the diffraction limit of the telescope. These very high Strehl ratio images also al- low us a glimpse into the future of planet-finding, where high-order AO systems such as the planned AO system for GPI, will make such high Strehl ratios routine at near-IR wavelengths as well.

4.2 Circumstellar Structure around Asymptotic Giant Branch Stars

Extensive mass loss in the asymptotic giant branch (AGB) phase has been well established.1 However, the mode (or modes) by which this mass loss occurs is less well known. While some objects present CO lines indicative of spherically symmetric mass loss, other objects display much more complicated mass loss line

1This work first appeared as B. Biller, L. Close, A. Li, J. Bieging, W. Hoffmann, P. Hinz, D. Miller, G. Brusa, M. Lloyd-Hart, F. Wildi, D. Potter, and B. Oppenheimer, 2005, The Astrophysical Journal, 620, 450. Reproduced by permission of the AAS. 134

profiles (Knapp et al., 1998; Kerschbaum & Olofsson, 1999). A small number of AGB and post-AGB stars display a particular type of anomalous CO features: a very narrow peak (as narrow as 1 km s1, see Kahane et al. 1998, but generally 5 km s1), with or without a broader underlying pedestal feature (with widths of 10-20 km s1). Stars which display just the narrow peak include AC Her, BM Gem, and the Red Rectangle. X Her (Kahane & Jura, 1996), RV Boo (Bergman, Kerschbaum, & Oloffson 2000), EP Aqr, RS Cnc, and IRC +50049 display narrow peaks as well as underlying broader components. Jura & Kahane (1999) inter- pret the narrow features in these objects as reservoirs of dust and molecular gas which are nearly at rest with respect to the central AGB stars. They suggest that a binary companion is necessary in these cases in order to entrain gas and dust into a circumbinary disk (Morris, 1987; Mastrodemos & Morris, 1998, 1999); both AC Her and the Red Rectangle have companions. The broader CO line compo- nents are interpreted as spherical outflows or in some cases (RS Cnc and X Her) as bipolar outflows (Kahane & Jura, 1996). RV Boo is unusual since interferomet- ric images of its CO emission suggest the presence of a large disk in Keplerian rotation (Bergman et al., 2000). Interestingly, Vinkovic et al. (2004) has also re- cently imaged another sort of asymmetry in AGB star envelopes – a small bipolar outflow observed in the near-IR around the AGB star IRC+10011. Determining the spatial structure of winds around AGB stars is important for constraining models of bipolar planetary nebulae. Generalized Interacting Stel- lar Winds (GISW) models of planetary nebulae invoke some initial toroidal struc- ture which can collimate and shape the fast winds produced by these objects. In these models, the existence or non-existence of this structure regulates whether the forming planetary acquires a round, elliptical, or bipolar morphol- ogy (Balick & Frank, 2002). This pre-existing structure is likely to have formed 135

by the end of the central object’s AGB stage, since a sizable percentage of proto- planetary nebulae display bipolar reflection nebulae. Could the disks and molec- ular reservoirs observed in CO around AGB stars be the initial stages in the for- mation of such pre-existing structure? The molecular reservoirs observed by Ka- hane et al. (1998) may be the diffuse precursors to the formation of a denser disk or torus around the star at the end of the AGB phase. Such dense disks might be capable of collimating the fast winds produced when these objects evolve to the PN phase. At mid-IR wavelengths, we expect dusty disks around AGB and post-AGB stars to be detected in thermal emission. Do we observe the disks implied by CO observations at these wavelengths? Meixner et al. (1999) observed a number of protoplanetary nebulae (pPNe) in the mid-IR. In those which were resolved, they found two primary mid-IR morphologies – core/elliptical and toroidal. Ueta, Meixner, & Bobrowsky (2000) found that each of these mid-IR morphologies also corresponds to a specific optical morphology. The optically thick core/elliptical mid-IR morphologies possessed bipolar reflection nebulae and heavily obscured central stars. The optically thin toroidal mid-IR morphologies possessed ellip- tical reflection nebulae and non-obscured central stars. In a number of cases dusty disks have not been observed directly but may collimate reflection nebulae – for instance, Roddier et al. (1995) observe double lobed reflection in the NIR (J and K bands) around the Red Rectangle (IRAS 06176-1036) and Frosty Leo (IRAS 09371+1212). Meixner et al. (1999) resolved the Red Rectangle into an unresolved core and an extended elliptical nebulosity but found Frosty Leo to be unresolved. RV Boo is an O-rich type b semiregular variable. It varies in luminosity and in spectral class between M5III and M7III with a period of 140 days. Previously, Bergman et al. (2000) observed a 400 diameter disk in CO around RV Boo. They 136

interpret this disk as possibly the first known Keplerian disk around an AGB star. The resolution in CO radio lines is limited to 1-200 even using interferometry; by observing at high resolution in the mid-IR, we can probe the structure of this disk on much finer scales. Using the unique adaptive secondary mirror AO system at the 6.5m MMT (Wildi et al. 2003, Brusa et al. 2003), we can observe AGB stars at mid-IR wave- lengths with 0.100 resolution (Close et al. , 2003). Through deconvolution, the SR 100% images produced with AO at the MMT allow resolutions better than that of the FWHM of the diffraction limit of the telescope. With such resolutions, we can probe AGB star and PPN morphologies on finer scales than ever before possible in the mid-IR. Here, we present the first adaptive optics high resolution images of RV Boo.

4.3 Observations and Data Reduction

Data were taken on the night of 2003 May 13 (UT) at the 6.5m MMT using the adaptive secondary mirror AO system with the BLINC-MIRAC3 camera (Hoff- man et al., 1998; Hinz et al., 2000). The adaptive secondary corrected the first 52 system modes at 550 Hz and achieved Strehl ratios as high as 0.97 0.03 from 8.8 - 18 m. These Strehl ratios are the highest ever presented in the literature for ground based telescopes; previous Strehl ratios for large telescopes have rarely exceeded 0.7 at any wavelength. Our unique ability to do AO correction at 10 m leads to very high Strehl ratios regardless of the seeing, airmass, or wind (Close et al. , 2003). With the very stable PSFs that result from such high Strehls it is pos- sible to detect structures with spatial scales smaller than the diffraction-limited FWHM ( 0.98 rad) through the use of image deconvolution. D Images of all the PSF stars observed on 2003 May 13 are displayed in Fig. 4.1. 137

To further illustrate stability, we subtracted one PSF star ( Her) from another ob- served later in the evening (AC Her). The residuals are displayed in Fig. 4.2. The residual flux after PSF subtraction is < 0.5% of AC Her’s original flux. Similar residuals resulted from PSF subtractions at 9.8 m and 18 m. Based on these ex- cellent subtractions, we conclude that the PSF obtained from the MIRAC3 camera with the MMT adaptive secondary AO system is extremely stable. RV Boo was observed in the 9.8 m wavelength band. Point sources UMa and Her were observed in this band before and after RV Boo to use as PSF cal- ibrators. Her is a relatively wide binary with a separation of 4.700 (Jeffers and Vasilevskis, 1978); the brighter component was used as a PSF while the fainter component falls outside our field of view. We note that Her also does not possess a dust shell or other extended structure (Close et al. , 2003). To elimi- nate the high sky background at mid-IR wavelengths, we used a standard chop- ping/nodding scheme and flat-fielded our data. A chopping frequency of 1 Hz (throw 2000) was used with a nodding cycle of 60 sec (throw 6-800). To avoid sat- uration of the high sky background, a base integration time of 29 msec was used. These images were coadded to produce an output frame every 15 seconds. For RV Boo, four 15-second integrations were taken at each of eight nod positions, giving a total exposure time of eight minutes. UMa was observed using eight nod positions, for a total exposure time of eight minutes. Her was observed using four nod positions, for a total exposure time of four minutes. We used the internal BLINC cold chopper and kept the AO in closed loop for both chop and nod beam positions. Flat fields were taken the night of 2003 May 15 (UT). The base integration time was set to 10 msec. These images were coadded every two seconds. Flats were taken of the inside of the dome (hot) and the sky (cold). The sky flats were 138

subtracted from the dome flats. The resulting flat fields were normalized by the mean. To determine an astrometric calibration, we used our 25 November 2002 (UT) observations of the WDS 02589+2137 BU. These data were taken in the M band using MIRAC with the MMT adaptive secondary AO system. At the time of observation, the binary had a position angle of 269 and a separation of 0.50900 (Mason et al., 2001). To align our RV Boo data with North, we must rotate it by (270 – the parallactic angle) at the time of observation. We also determine a plate scale of 88 milliarcsec/pixel from this standard. Data were reduced using a custom IRAF pipeline which first flat fields and re- moves bad pixels. After the pipeline completes these basic data reduction tasks, it then rotates the nod images by (270– the parallactic angle) and coadds them so that North is up and East is left. A coadded image of RV Boo alongside similar im- ages of the PSF stars Uma and Her as well as the AGB star AC Her (Close et al. , 2003) is presented in Fig. 4.3. The vertical axis is telescope altitude while the hor- izontal axis is telescope azimuth. RV Boo appears slightly extended (FWHM 4 pixels) relative to the PSF stars (FWHM 3.8 pixels.) All three PSF stars are very slightly (eccentricity 2%) elliptical along the horizontal (azimuthal) direction – this is a systematic instrumental feature of the PSF. However, RV Boo appears somewhat more elliptical (eccentricity 5%) than the other three stars. We argue that this extension and slight ellipticity are indicative of actual physical structure. The observed extension lies near the limit of resolution of the telescope, however, and is broadened by diffraction. To discern the actual small scale structure of the extension around RV Boo, we must deconvolve it with a PSF star. In order to determine if the extension we see is real (and not just the result of a vibration in the telescope mount, for instance), we deconvolved each of the eight 139

RV Boo nod images with the UMa PSF. After subpixel interpolation by 3x (to a new platescale of 29.3 milliarcsec/pixel), we used the Lucy deconvolution algo- rithm in IRAF with 1000 iterations. We chose to deconvolve for 1000 iterations for two reasons – first, object properties (position angle and deconvolution) vary rapidly up until the 800th iteration. By the 1000th iteration, properties have converged. Secondly, we convolved a set of thermal disk models at inclination angles from edge on of 5, 15, and 30 (see 4.5 for details on our modeling) with § the UMa PSF, then deconvolved for 100, 500, 1000, 1500, and 2000 iterations using AC Her (Close et al. , 2003) as the PSF. The 1000 iteration deconvolutions best recreated the position angles and eccentricities of our models. If the extension is real we expect the position angle of the deconvolved semi- major axis to track the parallactic angle as the sky rotates between exposures. We measured position angle and eccentricity for each deconvolution of the 8 indi- vidual RV Boo nod images using the imexam tool in IRAF. The imexam tool is only accurate for measuring isophote semi-major position angles to within 6 of accuracy. We calculate =5.65 as our error in position angle measurements. This error was determined by stretching an image of UMa to 5% eccentricity (about the eccentricity of RV Boo previous to deconvolution), rotating it through a number of angles, and then taking the standard deviation of (measured angle - actual angle) as the error. We plot the behavior of the major axis position angle of the deconvolved im- age (henceforth PA) and parallactic angle with time in Figure 4.4. We fit the expected parallactic angles (calculated using the SKYCALC software package – Thorstensen (2001)) to the position angle data for each star with a minimized 2. Thus, PA measures how much the trend in position angle deviates from the trend in parallactic angle. Expected parallactic angles are plotted as solid lines; 140

observed values of PA are plotted using a variety of points. We plot the behav- ior of PA for RV Boo deconvolved with UMa as the PSF. As a comparison, we plot PA for UMa deconvolved with Her, a PSF deconvolving a PSF (both apparent point sources). The PA values measured from the RV Boo deconvo- lutions follow the correct rotation of the sky with a reduced 2 of 0.207 for the deconvolution with UMa. Thus, RV Boo’s extension is rotating on the sky to within a probability of 98%. In contrast, the PA values measured from the PSF and PSF deconvolution show much more scatter and have a best fit to the paral- lactic angles with a reduced 2 of 13.2. Hence, the extension of a PSF deconvolved with a PSF is purely an artifact; the probability that PA for UMa deconvolved with Her rotates on the sky is less than 0.1% as one would expect. The variation of PSF FWHM with source eccentricity for RV Boo, UMa, Her, and AC Her is presented in Fig. 4.5. These quantities are also presented in Fig. 4.5 for the best fit thermal emission model of the RV Boo disk convolved with the UMa PSF (see 4.5). An extended source should have a larger FWHM and § possibly a higher eccentricity (depending on position angle, source shape, etc.) than a point source; RV Boo indeed shows this trend. The RV Boo disk models accurately recreate the FWHM and eccentricity of the RV Boo data. Models were fit to the deconvolved data; the match between convolved model and undecon- volved data implies that 1000 iterations of the Lucy algorithm produces an ac- curate deconvolution of the RV Boo data. Based on the rotation of the extension with the sky and its non-PSF FWHM and eccentricity, we tentatively conclude that the extension we observe around RV Boo is indeed real and not an artifact of the telescope. However, to strengthen this conclusion, we reobserved RV Boo at a variety of wavelengths at another epoch. 141

4.4 Followup Observations

RV Boo was reobserved on the night of 2004 February 2 (UT) at the 6.5m MMT using the adaptive secondary mirror AO system with the BLINC-MIRAC3 cam- era (Hoffman et al., 1998; Hinz et al., 2000). Similar Strehl Ratios were observed as during the first observations, however, the MIRAC camera was used with a platescale of 0.07900 (as opposed to 0.08800 during the previous run.) These data was taken during an engineering test and are of considerably lower quality (en- gineering grade rather than science grade) than the May 2003 dataset. As before, we used a standard chopping/nodding scheme. A chopping fre- quency of 1 Hz (throw 2000) was used. Each object was observed at three differ- ent wavelengths (8.8 m, 9.8 m and 11.7 m) and two different nod positions (throw 4-500). To avoid saturation of the high sky background, a base integration time of 30-70 msec was used. These images were coadded to produce an output frame every 10 seconds. For RV Boo, 10x10s integrations were taken at two nod positions for each wavelength, giving a total exposure time of 200 seconds per wavelength. Boo was observed at two nod positions per wavelength, for a total exposure time of 200 seconds per wavelength. We used the internal BLINC cold chopper and kept the AO in closed loop for all chop and nod beam positions. The data were reduced with the same custom IRAF pipeline as before. With shorter total integration times (200 s per wavelength for RV Boo during reobservations vs. 480 s at 9.8 m during the initial observations), our signal to noise during the followup observations is only 65% that achieved during the initial observations. PSF and data FWHMs for both the May 2003 and February 2004 datasets are presented in Table 4.1. We estimate an uncertainty in our measurements of 6 milliarcsec from the scatter of the PSF FWHMs in the May data. The FWHM of RV Boo was slightly larger than that of the PSF in all wavelengths in February; 142

May 2003 February 2004 UMa Her AC Her RV Boo Boo RV Boo (PSF) (PSF) (PSF) (PSF) 8.8 m 0.29200 0.29700 9.8 m 0.32400 0.33200 0.31900 0.34800 0.32200 0.33800 11.7 m 0.38200 0.39300 The uncertainty for all of the FWHMs is 6 milliarcsec. Table 4.1 FWHMs for RV Boo and PSF stars however, the extension is within the measurement error for the 8.8 and 11.7 m data. The February RV Boo 9.8 m data possesses a FWHM comparable to that of the May RV Boo data – thus, we reobserve the slight extension found in May in our February dataset. Perhaps it is not surprising that the object appears most extended at 9.8 m – right at the center of a prominent silicon emission feature seen in the ISO spectrum (see 4.5 and Fig. 4.7). For RV Boo, we measured ec- § centricities of 2% in February vs. 5% in May. This may have simply been a function of position angle during the observations. Unfortunately, the engi- neering grade observations of the only PSF star ( Boo) were found to have low spatial frequency noise and were unsuitable to be used as a PSF for deconvolu- tion. As a test, we deconvolved these observations using the UMa PSF from the previous observations. The February data possessed similar deconvolved source properties (FWHM, position angle, etc.) compared to the May data.

4.5 Analysis

The deconvolved image of RV Boo is presented in Fig. 4.6. This image was pro- duced by coadding all of the RV Boo nod images which had been deconvolved 143 for 1000 iterations using the Lucy algorithm with UMa as the PSF. (Due to the slight saturation of the Her image, UMa was chosen as the PSF over Her for this analysis.) After deconvolution, the extended structure appears to be a disk seen at an inclination angle between 20 to 30 from edge-on (inclination es- timated from modeling; see below). From their CO J=2-1 interferometric data, Bergman et al. (2000) found evidence of a rotating disk around RV Boo in the form of a 400 diameter disk with a position angle of 150. Using the imexam tool in IRAF, we measured a FWHM at =9.8 m of 0.1600 along the major axis and a PA of 120. The 9.8 m emission traces dust thermal re-emission; RV Boo heats only a fraction of the dust in the entire CO disk to temperatures where dust is luminous in the mid-IR. If circumstellar gas and dust are well-mixed around RV Boo, we would expect to find a much smaller disk extent at 9.8 m than in CO, since all of the CO will emit, but only the dust near enough to the star to be heated to temperatures of 300 K will emit at 9.8 m. To map out more extension in the dust disk, observations at longer wavelengths (which trace a larger region of dust re-emission) are necessary. We measure a total flux for RV Boo at 9.8 m of 145 24 Jy. +250 At a distance of 390100 pc (Perryman et al., 1997), the mid-IR disk of RV Boo +40 subtends a FWHM of 6015 AU. RV Boo is a relatively nearby AGB star; it would be more difficult to observe a similar disk around a more distant star like AC Her, which is 2 times more distant. Since RV Boo is so nearby relative to the distance of the average AGB star, it is not surprising that such small scale structure usually is unresolved in other systems. Small mid-IR disks around AGB stars could be common but difficult to resolve. We model the IR emission of RV Boo as an optically passively heated 144

by the star.2 We approximate the stellar by the Kurucz (1979) model

spectrum for M6IIIe stars with an effective temperature of Te = 3000 K. The dust is taken to be amorphous silicate since RV Boo is an M star. We assume a power- law dust size distribution dn(a)/da a which is characterized by a lower- ∝ 3 cutoff amin, upper-cutoff amax and power-law index . We take amin = 0.01 m since grains smaller than this will undergo single-photon heating (Draine and Li, 2001) while the observed IR spectral energy distribution of RV Boo does not appear to show evidence for stochastically heated dust, and amax = 1000 m since larger grains are not well constrained by the currently available IR photometry. The dust spatial distribution is taken to be a modified power-law dn/dr ∝ (1 r /r) (r /r) where r is the inner boundary of the disk which we take min min min to be the location where silicate dust sublimates. For RV Boo with a luminos- ity L? 8100 L (Bergman et al., 2000) submicron-sized silicate dust achieves an equilibrium temperature of T 1500 K and starts to sublimate at r 3 AU. Therefore we take rmin = 3 AU. This functional form has the advantage that on one hand, it behaves like a power-law dn/dr r at larger distances (r r ), ∝ min and on the other hand it peaks at rp = rmin ( + ) /, unlike the simple power- law which peaks at rmin. The latter is unphysical since one should not expect dust to pile up at rmin where dust sublimates! We take = 2 as expected from a stationary outflow. We take rmax = 120 AU which is large enough for our dust IR emission modeling purpose since there is very little dust beyond this outer boundary. Therefore, we have only two free parameters: – the power-law expo- nent for the dust size distribution and – the dust spatial distribution parameter

2RV Boo has a B–V color of 1.47 mag; this low reddening implies that the mid-IR emitting region is not obscuring the entire star (as it would be if the dust was spherically distributed around the star) and is largely concentrated along one plane. Thus, we have explicitly modeled the IR emission around RV Boo as a disk. Since we have apriori considered a disk morphology, our model cannot constrain the vertical geometry of the emission. 3We assume all grains are spherical in shape with a being their spherical radius. 145

which determines where the dust peaks. Using the dielectric functions of Draine and Lee (1984) for “astronomical sil- icates” and Mie theory (Bohren & Huffman, 1983), we calculate the absorption cross sections of amorphous silicate grains as a function of size and their steady- state temperatures (as a function of radial distance from the central star) in ther- mal equilibrium with the illuminating starlight intensity. We then obtain the dust model IR emission spectrum by integrating the dust emission over the entire size range and the entire disk and compare with the 12, 25, 60, and 100 m IRAS (Infrared Astronomical Satellite) broadband photometry, the 7.7–22.7 m low res- olution spectrum obtained with the IRAS Low Resolution Spectrometer (LRS; with a resolution of / 20), the 2.5–45 m high resolution spectrum obtained with the Short-Wavelength Spectrometer (SWS) instrument (with a resolution of / 2000) on board the Infrared Space Observatory,4 and the 9.8 m photome- try presented in this work. We also compare the dust model image at 9.8 m con- volved with the AO instrument PSF with the AO image of RV Boo and inclined at a variety of inclinations from edge on. A number of models with 25 30 and inclinations from edge on of 30-45 fit the data to within 20% accuracy. In Figures 4.7 and 4.8, we respectively show the best-fit IR emission spectrum and disk image. The disk image is also shown convolved with the instrument PSF spectrum (see Fig. 4.7) and the AO image (see Fig. 4.8). The source properties (eccentricity, FWHM) for this model are plotted alongside source properties for the telescope data and PSF stars in Fig. 4.5. This model, with 3.3 and 25, 27 6 has a total dust mass m 3.17 10 g 1.6 10 M and an inclination dust 4The IRAS LRS spectrum integrated with the IRAS 12 m band filter function results in a factor of 9.8 higher than the IRAS 12 m photometric flux. This is not surprising in view of the fact that RV Boo is a . We have therefore reduced the flux level of the IRAS LRS spectrum by a factor of 9.8 to bring it into agreement with the IRAS 12 m broadband photometric data. Similarly, the flux level of the ISO SWS spectrum has also been reduced by a factor of 1.15, in order to agree with the IRAS 12 m photometry. 146

from edge on of 40. The dust spatial distribution peaks at r 40.5 AU. The p maximum vertical visual optical depth is V 0.024, confirming the validity of the optical-thin assumption made at the beginning of this modeling effort.

4.6 Discussion

Where do the wide variety of PNe shapes come from? Mass loss has practically ended by the PPN stage; therefore, any underlying structure which produces the shape of PN (via theories such as the Generalized Interacting mod- els) must form by the end of the AGB stage. While PNe possess an extremely wide range of morphologies, Meixner et al. (1999) found two primary mid-IR morphologies for PPNe corresponding to different values of optical depth – opti- cally thick core/elliptical morphologies and optically thin toroidal morphologies. Ueta et al. (2000) found that each of these mid-IR morphologies also corresponds to a specific optical morphology. Optically thick core/elliptical mid-IR morpholo- gies possess bipolar reflection nebulae and heavily obscured central stars (called DUPLEX by Ueta et al. (2000)) while optically thin toroidal mid-IR morphologies possess elliptical reflection nebulae and non-obscured central stars (called SOLE by Ueta et al. (2000)). The PPNe observed by Meixner et al. (1999) are more evolved objects than AGB stars. Since mass loss has ceased by the PPN stage, the dense core/elliptical dust structures which produce DUPLEX sources (and are the precursors to struc- tures formed during the PN stage) must have formed during the end of the AGB phases. AGB mass loss can be divided into two separate phases: an initial, spher- ically symmetric AGB wind ( 10 km s1), supplanted by a faster superwind ( 20 km s1) for a brief period at the end of the AGB phase (Ueta et al. (2000), Renzini et al. (1981)). With the onset of the superwind, mass loss rates are expected to 147

rise by factors of 10 (Steffen et al., 1998). Ueta et al. (2000) propose that this su- perwind at the end of the AGB phases is intrinsically asymmetric, producing an equatorially flattened toroid which collimates later bipolar structure. However, it is not clear what mechanism could produce the asymmetry. An AGB star like RV Boo which displays asymmetric structure is an excellent laboratory for studying the very beginnings of PN and PPN structure formation. RV Boo is one member of a set of AGB stars which display hallmarks of asym- metric structure. This small group of AGB stars ( 20%) display very narrow CO linewidths (as narrow as 1 km s1, see Kahane et al. 1998, but generally 5 km s1), with or without a broader underlying pedestal feature (with widths of 10-20 km/s). BM Gem displays just a narrow peak, while X Her (Kahane & Jura, 1996), RV Boo (Bergman et al., 2000), EP Aqr, RS Cnc, and IRC +50049 display narrow peaks as well as an underlying broader component. Jura & Kahane (1999) inter- pret the narrow features in these objects as reservoirs of dust and molecular gas which are nearly at rest with respect to the central AGB stars. The broader CO line components are interpreted as spherical outflows or in some cases (RS Cnc and X Her) as bipolar outflows (Kahane & Jura, 1996). RV Boo is unusual among these stars since images of its CO emission suggest the presence of a large disk in Keplerian rotation (Bergman et al., 2000). One pos- sible explanation for the presence of the disk is that mass loss from the AGB star has become equatorially enhanced through entrainment by a binary companion (Jura & Kahane, 1999; Morris, 1987; Mastrodemos & Morris, 1998, 1999). Could this be the mechanism (in progress) which produces the dense collimating toroid of material invoked by Ueta et al. (2000) – essentially, Ueta’s equatorial super- wind? While RV Boo does not have a known companion, a close companion (>100 AU) may be currently undetectable. 148

Interestingly, similar narrow CO line structure has been observed around a number of evolved post-AGB systems – two RV Tauri stars, AC Her (which has only a narrow line without a broad pedestal, suggesting perhaps that mass loss has ceased and only a molecular reservoir remains) and IRAS 08544-4431 (Maas et al., 2003) and the PPN Red Rectangle (Jura et al., 1995). Additionally, a dusty disk has been detected around IRAS 08544-4431 and a Keplerian disk has been detected in CO around the PPN Red Rectangle (Bujarrabal et al., 2003). Are the disks around IRAS 08544-4431 and the Red Rectangle similar to that around RV Boo, but at the corresponding later stages of evolution? All three of these objects have known companions, a fact which supports theories of binary entrainment for disk formation. If RV Boo is nearing the end of its AGB phase, the disk structure observed in CO and in the mid-IR may be the beginning of a denser disk/torus capable of col- limating the fast winds produced during the PN phase. However, to determine whether the structure around RV Boo is the precursor of PPN and PN structure requires knowledge of the evolutionary status of RV Boo and also whether ob- jects of a similar evolutionary status possess similar CO and mid-IR structure. RV Boo is a “red” SRb, according to the classification system of Kerschbaum & Hron (1992). They interpret the red SRb’s as a transitional stage between the blue SRb’s, which sit near the tip of the branch, and Miras, which mark the

7 1 tail end of the AGB phase. With a moderate mass loss rate of 10 M yr , RV Boo is probably not at the very tail end of AGB evolution. However, Kerschbaum & Hron (1992) use single-star stellar models to constrain evolutionary state. If disk formation does require a binary companion, estimates of evolutionary state from such models may be inaccurate for RV Boo and similar objects. While narrow CO lines have been found in 20% of AGB stars, it remains to 149

be seen how common extended mid-IR structure such as that observed around RV Boo occurs in AGB stars. Indeed, even if mid-IR structure around narrow CO line AGB stars is common, it remains to be seen how similar RV Boo’s CO and mid-IR structures are to those of other AGB stars with narrow CO lines. This is a comparison we can not yet make, since no other similar object has both been mapped interferometrically in CO and imaged at high resolution in the mid-IR. Several of these objects share very similar observational properties with RV Boo – in particular, X Her, EP Aqr, RS CNc, BM Gem, and EU And. It would be worthwhile observe these objects both at high resolution in the mid-IR and also interferometrically in CO.

4.7 Conclusions

We present the first high resolution (0.100) very high Strehl ratio (0.97 0.03) mid- IR images of RV Boo utilizing the MMT adaptive secondary AO system. These are the first ground-based IR images presented in the literature with such high Strehl ratios; previous Strehl ratios for large telescopes have hardly exceeded 70% at any wavelength. RV Boo was observed at a number of wavelengths over two epochs (9.8 m in May 2003, 8.8, 9.8 and 11.7 m in February 2004) and appeared slightly extended at all wavelengths. While the extension is within the measure- ment error for the 8.8 and 11.7 m data, the extension is more pronounced in the 9.8 m data. The slight extension seen in the 9.8 mum data from both May 2003 and February 2004 suggests that the mid-infrared structure around RV Boo is marginally resolved at 9.8 m. Because of our high Strehl ratios which leads to extremely stable PSFs, we can deconvolve our images with those of PSF stars for a super-resolution of 0.100. Based on the rotation of the extension with the sky and its non-PSF FWHM (at 150

8.8-11.7 m) and eccentricity, we conclude that the extension around RV Boo is indeed real and not an artifact of the telescope. We tentatively detect a 60 AU FWHM (0.1600 at 390 pc) disk at 9.8 m around RV Boo. Previously, Bergman et al. (2000) found a 400 disk in CO with a position angle of 150; we find a position angle of 120 6 for the 9.8 m disk. We measure a total disk flux of 145 24 Jy at 9.8 m. We closely reproduce the observed IR spectral energy distribution and the AO image in terms of an optically thin dust disk consisting of amorphous silicates with a power-law size distribution. We estimate a disk inclination angle of 30 to 45 from edge on and a disk dust mass of 1.6 106 M . 151

Figure 4.1 The 9.8, 11.7, and 18 m images of the PSF stars AC Her, UMa, and Her as observed at the MMT. The box size of the MMT images is 1.5 1.000. The faint point source in the lower left of each MMT image is a MIRAC3 ghost. 152

Figure 4.2 The 11.7 m PSF of AC Her before (left) and after (right) PSF subtrac- tion (using Her as the PSF) with DAOPHOT’s ALLSTAR task. The residual flux after PSF subtraction is < 0.5% of AC Her’s original flux. Similar residuals resulted from PSF subtractions at 9.8 m and 18 m. Based on these excellent subtractions, we conclude that the PSF obtained from the MIRAC3 camera with the MMT adaptive secondary AO system is extremely stable. Note that the small ghost image to the lower left in each frame is not subtracted to show that the vertical scales are the same for both images. 153

Figure 4.3 AO Images of RV Boo, UMa, Her, and AC Her at 9.8 m. The vertical axis is telescope altitude while the horizontal axis is telescope azimuth. The images are Fourier filtered: An image smoothed with a Gaussian with a 3 pixel FWHM was subtracted from each of the original images in order to remove variations on large spatial wavelengths. All images are shown on a logarithmic scale; the bright ring around the images is the first Airy ring. Note that RV Boo appears nominally extended relative to the other stars. The PSF star Her is slightly saturated. A 0.700 diameter circle is overlaid on each image to help aid the eye. 154

Figure 4.4 Position angle of the semi-major axis vs. time (after first observation) for deconvolved RV Boo and UMa nod images. Measured position angles are represented as crosses for RV Boo and circles for UMa; solid lines depict the pre- dicted parallactic angle as a function of time during the observations. Note that the PAs measured for RV Boo track the parallactic angle much more closely than those measured for UMa, with a reduced 2 value of 0.207 for RV Boo de- convolved with UMa versus a reduced 2 value of 13.2 for UMa deconvolved with Her. This implies that the elongation observed was really associated with RV Boo (since it was rotating along with the sky) and is not a PSF artifact. 155

Figure 4.5 Eccentricity vs. PSF FWHM for RV Boo, UMa, Her, and AC Her images. The best fit thermal emission model of RV Boo convolved with the UMa PSF is also plotted (see 4.5 for details on modeling). The three PSF stars § are plotted as green crosses. The RV Boo data is plotted as a red cross and the RV Boo model is plotted as a blue star. FWHM is measured by a Gaussian fit to the enclosed flux at each radius. RV Boo appears slightly extended and has a significantly higher eccentricity and FWHM than the other stars. 156

Figure 4.6 Deconvolved image of RV Boo. The platescale (after magnification) is 29.3 milliarcsec/pixel. North is up, east is left. The disk is at a position angle of 120 and has a major axis FWHM of 0.1600 ( 60 AU at 390 pc). A deconvolved image of UMa (deconvolved using Her as a PSF) is shown on the right for comparison. For each object, there are 10 contours spaced linearly. For RV Boo, the lowest contour level is placed at 3% of the peak flux (275 Jy arcsec2) and the highest contour level is placed at 98% of the peak flux (9640 Jy arcsec2). For UMa, the lowest contour level is placed at 1% of the peak flux (279 Jy arcsec2) and the highest contour level is placed at 95% of the peak flux (2.24 104 Jy arcsec 2). After deconvolution, RV Boo appears extended, while the PSF still appears pointlike. 157

Figure 4.7 IR emission and model fit to the RV Boo disk. The open black squares are the U, B, V, I, K, the 9.8 m AO photometry, and the 12, 25, 60 and 100 m IRAS photometry. Thin solid black line is the IRAS LRS spectrum (reduced by a factor of 9.8 in order to agree with the IRAS 12 m data; see 4). Thin solid cyan line is § the ISO SWS spectrum (reduced by a factor of 1.15; see 4; the spectrum longward § of 30 m is too noisy and therefore not shown in this figure). Green dotted line is the dust model emission. Blue dashed line is the stellar photospheric spectrum approximated by the Kurucz model of Te = 3000 K. Red solid line is the sum of the dust and stellar emission. At a distance of d 390 pc, this best fit model has an inner edge at 3 AU and a peak dust density at r 40.5 AU. p 158

Figure 4.8 Comparison of RV Boo to the best fit dust disk model (rinner = 3 AU, rp 27 6 = 40 AU, 3.3, 25, total dust mass mdust 3.17 10 g 1.6 10 M and an inclination from edge on of 40 .) For comparison with the raw RV Boo image, the model image in the lower right has been convolved with the UMa PSF. North is up and east is to the left in all these images. Note that the RV Boo data is consistent with the best fit SED model from Fig. 4.7. 159

CHAPTER 5

PRELIMINARY RESULTS OF A MULTI-WAVELENGTH DIFFERENTIAL IMAGING

EXPERIMENT FOR THE HIGH CONTRAST IMAGING TESTBED

5.1 Introduction / Motivation

The Simultaneous Differential Imager (SDI) implemented at the VLT and MMT (Biller et al. 2004, Lenzen et al. 2004) uses a quad filter to take images simulta- neously at 3 wavelengths surrounding the 1.62 m methane bandhead found in the spectrum of cool brown dwarfs and gas giants. By performing a difference of images in these filters, speckle noise from the primary can be significantly attenu- ated – for instance, in our survey data, we achieved H band contrasts >25000 (5 F1(1.575 m) > 10 mag, H > 10.8 mag for a T6 spectral type) at a separation of 0.5” from the primary star. With this degree of attenuation, we should be able to image (5 detection) a 2-4 Jupiter mass planet at 5 AU around a 30 Myr star at 10 pc. We believe that our SDI images are the highest contrast astronomical images ever made from ground or space for methane rich companions within 1” of their primary star (Masciadri et al. 2005, Biller et al. 2007, Mugrauer & Neuhauser¨ 2005). SDI on ground based telescopes provides significant speckle attenuations down to star-planet contrasts of 1-3 104. To test the classical SDI technique at con- trasts of 1069, to develop a similar “2nd order” SDI technique incorporating Fresnel propagation in the speckle solution, and to determine whether such a technique would be applicable for the Finder mission, we im- plemented a similar multiwavelength differential imaging scheme for the High Contrast Imaging Testbed. The High Contrast Imaging Testbed tests precursor technologies for the Ter- 160

restrial Planetfinder Mission and consists of a coronagraphic testbed kept in vac- uum and vibration-isolated. High contrasts are provided by a band-limited coro- nagraph – in addition, speckle noise is suppressed at the HCIT by the use of an additional deformable mirror. In the simulated space-like environment of the HCIT, speckles evolve on hours to days timescales (as opposed to minutes on the ground.) Thus, there is plenty of time to measure the error in the wavefront and use the deformable mirror to dial in the exact counter to that wavefront. In this way, speckles are nulled out at the HCIT one by one. Speckles are produced by both phase (symmetric) and amplitude (non-symmetric) errors in the wavefront; however, the deformable mirror can only produce symmetric speckles to null out existing speckles. In the contrast regime (up to 106, as described by Fraun- hofer diffraction) where phase errors predominate, the deformable mirror can then be used to create a symmetric “dark hole” within its control radius. How- ever, at higher contrasts (beyond 107, where the Fresnel formalism is necessary to describe diffraction), amplitude errors begin to predominate. Since they are not symmetric, “nulling” a non-symmetric amplitude speckle removes a bright speckle from the right side of the dark hole but then effectively adds the bright speckle back in to the left side of the hole. Thus, at higher contrasts, the “dark hole” loses its symmetry and appears only on the right side of the image. Con- trast is highest within the “dark hole” region and is considerably lower outside the control radius of the deformable mirror. For our differential imaging experiment, we selected five filters near the promi- nent O2(A) absorption feature at 0.762 m (seen in Earth’s atmosphere and ex- pected for any terrestrial extrasolar planet with an oxygen atmosphere, Woolf et al. (2002), with approximate central wavelengths of F1(768 nm), F2(784 nm), F3(800 nm), F4(816 nm), and F5(832 nm). Exact filter wavelengths and band- 161

widths are presented in Table 5.1 – however, approximate wavelengths are used to refer to the filters throughout the text). Two sets of images were taken in each filter – a long set, with 90 s exposure time per filter, and a short set, with 5 s expo- sure time per filter. The filter set was spaced across a considerable bandwidth in wavelength in order to measure speckle chromatism as a function of wavelength. For ground based observing, simultaneous imaging in several bandwidths is necessary to overcome the stochastic speckle noise floor remaining even af- ter adaptive optics correction. For space-based observing, however, speckles are stable on timescales of hours to days, making simultaneity of imaging unneces- sary. This multi-wavelength differential imaging experiment measures speckle evolution as a function of wavelength and contrast level. We test whether the ground-based simultaneous differential imaging technique can be generalized to a non-simultaneous differential imaging technique for a space mission. By using a 5 filter set, we can attempt to correct for speckle chromatism using both single differences of images and the double difference technique pioneered by Marois et al. 2000. Since the radial position of the speckle pattern in each

filter image is proportional to D , the platescale of each image must be scaled so that the speckles in each filter fall at the same radii despite chromatic differences. After this scaling, a number of single differences as well as a double difference of images in the F1 through F5 filters are calculated:

(F 3 F 1) (5.1)

(F 3 F 2) (5.2)

(F 3 F 4) (5.3) 162

(F 3 F 5) (5.4)

(F 4 F 3) (F 3 F 2) (5.5)

(F 5 F 3) (F 3 F 1) (5.6) In this document, we discuss preliminary results using the classical SDI data reduction method (i.e. no Fresnel propagation is considered.)

5.2 Data Acquisition and Reduction

Data were acquired in January 2007. A series of 5 filter images were acquired at each nominal contrast value (106, 107, 108, and 109). After optical speckle nulling in the F3 800 nm filter, two images (with exposure times of 90 s and 5 s) were ac- quired at each of the four filter wavelengths. In this document, we focus on the 90 s dataset, since its signal to noise is considerably higher than that of the 5 s dataset (although a degree of saturation and bleeding is observed in the 106 contrast im- ages). Since the images in the nulled wavelength are of the highest quality and are the most reliable, the F3 800 nm filter image was used as the “master” image (for adjustment of platescale and alignment purposes) in this analysis. Fluxes in each wavelength image were converted to the equivalent contrast (i.e. the ratio of each pixel’s brightness to the peak of the unocculted star intensity divided by the occulter transmission). Since the radial position of the speckle pattern in each filter image is propor-

tional to D , the platescale of each image was scaled to the 800 nm image so that the speckles in each filter fall at the same radii despite chromatic differences (in 163 other words, we assume only phase errors in the platescale scaling). The exact filter central wavelengths presented in Table 1 were used for this scaling. Finally, each wavelength image was aligned to the 800 nm image using a cus- tom shift-and-subtract alignment algorithm (alignments are to 0.1 pixel precision, see Biller et al. 2004). The resulting aligned images were subtracted from the mas- ter 800 nm image. Galleries of each aligned image and the resulting subtraction are presented for each nominal contrast level in Fig. 5.1 to Fig. 5.4. These images are shown with a logarithmic stretch from contrast levels of 0 to 105. These sin- gle subtractions suppress the speckles outside the dark hole by a factor of 5 - 50. Two double differences were also calculated and are shown (same logarithmic stretch, also by nominal contrast level) in Fig. 5.5 through Fig. 5.8.

5.3 Analysis

To quantify the level of speckle attenuation available through the differential imaging technique both inside and outside the dark hole, we calculated the speckle RMS (in contrast levels) before and after subtraction in two 50 140 pixel regions of the chip for all our reduced images. Regions used for this analysis are shown in Fig. 5.9. One major caveat exists regarding our choice of left-side “compar- ison region” – we chose a comparison region outside the control radius of the deformable mirror, since the speckle nulling algorithm adds speckle noise to the left side of the image when removing a non-symmetric amplitude speckle from the right side of the image. Before nulling we note that a radial trend in speckle intensity exists – there are more and brighter speckles closer to the center of each image. Thus, the speckle noise within the the left-side region is somewhat lower than the pre-null speckle noise in the dark-hole region. For this reason, it is not instructive to compare speckle noise between the right and left side regions, but 164

instead to compare the degree of attenuation provided by the differential imag- ing technique in both of these regions. Speckle RMS information is presented in Table 2 and plotted as a function of from the nulling wavelength (800 nm) in Fig. 5.10. Outside the dark hole, the single differences reduce speckle RMS by a factor of 6 and speckle RMS does not vary by more than a factor of 2 as a function of nominal dark hole contrast level (in fact, in this particular region of the image, the pre-subtraction measured contrast level remains consistently at 3-8 108). Inside the dark hole, speckle RMS appears to depend strongly on from the nulling wavelength, especially at higher contrasts. This is likely a result of amplitude errors in the wavefront which are well corrected at the nulled wavelength but remain uncorrected at other wavelengths. At the highest nomi- nal contrast level (109), a single difference of images will “import” speckles in from the image with the largest from the nulling wavelength (for instance, the 800 nm - 832 nm subtraction introduces speckles from the 832 nm image into the subtraction). The double differences do not provide an advantage over the single differences – speckle RMS inside and outside of the dark hole is not decreased compared to the single differences. Contrast plots for each of the 4 single differences we calculated at each nomi- nal contrast level are presented in Figs. 5.11 to 5.14 . These plots were generated by tracking a 6x6 pixel box along a specific trajectory in the subtracted image (trajectories used are shown in Fig. 5.9) and calculating the absolute value of the median in that box for each position along the trajectory. Two trajectories are considered – one going through the dark hole and one going through the un- corrected half of the image. Outside of the dark hole, the subtraction suppresses speckles by a factor of 5 to 50 in the contrast plots. For lower contrast levels, a sim- ilar speckle suppression with subtraction is also observed within the dark hole. 165

At high contrasts (109), however, the chromaticity of the HCIT occulter (differ- ent phase errors at different wavelengths) strongly affects speckle evolution as a function of wavelength – the contrast degrades strongly for wavelengths away from the nulling wavelength and we end up ”importing” in speckles from other wavelengths (i.e. the 800 nm - 832 nm subtraction introduces chromatic speckles from the 832 nm image into the subtraction). Interestingly, the differential imag- ing technique appears to work equally well both in the pure phase error regime (contrasts less than 107) and in the phase + amplitude error regime (contrasts above 107, where the dark hole becomes asymmetric) before being ultimately limited by the chromaticity of the occulter. We are currently working on model- ing methods to remove occulter chromaticity effects from our images – once we have corrected for this chromaticity, we expect the differential imaging technique to continue to yield contrast improvements at dark hole contrasts of 109 and above. Increased speckle RMS in the dark hole as image wavelength differs from the nulling wavelength translates to a lower contrast in that image relative to the image at the nulled wavelength. Our multifilter experiment lets us simulate the variation in speckle RMS we would expect from a wideband image. Our filters cover a wavelength range of 80 nm – equivalent to a 10% bandwidth filter. At a contrast of 108, speckle RMS is 1.7 higher at the red end (830.9 nm) than at the nulled wavelength (798.9 nm) and at the blue end (768.1 nm), speckle RMS is 2 higher than at the nulled wavelength. Over the entire 80 nm bandwidth, speckle RMS is on average 1.5 higher than at the nulled wavelength (and, hence, the contrast would 1.5 lower over a 10% bandwidth filter than over a 2% band- width filter.) At the highest contrasts (109), contrast over the entire bandpass is considerably diminished – at the red end (830.9 nm), speckle RMS is 11 higher 166

than at the nulled wavelength (798.9 nm) and at the blue end (768.1 nm), speckle RMS is 15 higher than at the nulled wavelength. Over the entire 80 nm band- width, speckle RMS is on average 7.2 higher than at the nulled wavelength (and, hence, the contrast would 7.2 lower over a 10% bandwidth filter than over a 2% bandwidth filter.) Thus, apparent speckle chromatism in these images due to the chromaticity of the occulter itself is sufficient to predict considerable contrast degradation in a wide band filter.

5.4 Conclusions

At nominal contrasts of 106 to 109, single differences of filter images can reduce speckle noise outside of the dark hole by factors of 5 - 50. For contrasts of 106 to 108, a similar result is also found within the dark hole, with speckle attenu- ation achieved through a single difference decreasing as a function of increasing contrast. However, at high contrasts (109), considerable increase in speckle RMS between filters is observed to “pollute” the dark hole in all of our single differ- ences (where, in each difference, only one of the two wavelengths had undergone optical speckle nulling). At all contrast levels, a double difference of images does not seem to decrease speckle noise relative to the single differences. Significant differences in speckle RMS ( 2-10 RMS) is found between filters separated in wavelength by >20 nm. Speckle RMS between filters increases strongly as a function of increasing contrast. In this preliminary analysis, we have used the classical SDI data reduction method (i.e. no Fresnel propagation or correction for occulter chromaticity is considered.) We find that the differential imaging technique appears to work equally well both in the pure phase error regime (contrasts less than 107) and in the phase + amplitude error regime (contrasts above 107, where the dark 167

Table 5.1 Filter Wavelengths and Bandwidths Filter Center Wavelength Wavelengths at half max transmittance F1 768.1 nm 760.5 nm – 775.6 nm F2 782.5 nm 774.7 nm – 790.3 nm F3 798.9 nm 790.9 nm – 806.8 nm F4 814.8 nm 806.6 nm – 822.9 nm F5 830.9 nm 823.1 nm – 838.6 nm

hole becomes asymmetric) before being ultimately limited by the chromaticity of the occulter. We are currently working on modeling methods to remove occulter chromaticity effects from our images – once we have corrected for this chromatic- ity, we expect the differential imaging technique to continue to yield contrast im- provements at dark hole contrasts of 109 and above. 168

Table 5.2 Speckle RMS in Right-Side Dark Hole and Left-Side Comparison Region Image Region 106 107 108 109 F1 768.1 nm dh 1.1 106 7.1 108 1.4 108 1.2 108 oh 7.5 108 4.0 108 4.7 108 5.4 108 F2 782.5 nm dh 1.4 106 7.0 108 1.1 108 3.7 109 oh 7.2 108 3.7 108 4.3 108 5.0 108 F3 798.9 nm (nulled ) dh 8.9 107 6.9 108 7.3 109 7.9 1010 oh 6.9 108 3.7 108 4.4 108 4.9 108 F4 814.8 nm dh 9.8 107 7.7 108 1.0 108 3.0 109 oh 6.5 108 3.6 108 4.2 108 4.7 108 F5 830.9 nm dh 1.3 106 7.8 108 1.3 108 8.9 109 oh 6.6 108 3.5 108 4.1 108 4.7 108 798.9 nm - 768.1 nm dh 5.1 107 1.7 108 1.0 108 1.2 108 oh 8.8 109 6.4 109 6.8 109 9.3 109 798.9 nm - 782.5 nm dh 7.6 107 7.5 109 3.9 109 3.0 109 oh 6.4 109 2.6 109 2.9 109 3.5 109 798.9 nm - 814.8 nm dh 3.0 107 7.6 109 3.8 109 2.5 109 oh 3.9 109 2.7 109 3.4 109 4.0 109 798.9 nm - 830.9 nm dh 6.6 107 1.3 108 8.6 109 8.5 109 oh 5.7 109 4.7 109 5.3 109 5.8 109 (814.8 nm - 798.9 nm) dh 1.0 106 8.0 109 3.7 109 4.8 109 - (798.9 nm - 782.5 nm) oh 8.2 109 5.2 109 5.2 109 5.4 109 (830.9 nm - 798.9 nm) dh 1.2 106 1.8 108 1.3 108 1.9 108 - (798.9 nm - 768.1 nm) oh 1.1 108 9.0 109 1.1 108 1.4 108 169

Figure 5.1 Gallery of single wavelength images and single differences with a nominal contrast level of 106. All images are shown with the same logarithmic stretch from contrasts of 0 to 105. The single subtractions suppress the speckles outside the dark hole by a factor of 5 - 50. 170

Figure 5.2 Gallery of single wavelength images and single differences with a nominal contrast level of 107. All images are shown with the same logarithmic stretch from contrasts of 0 to 105. The single subtractions suppress the speckles outside the dark hole by a factor of 5 - 50. 171

Figure 5.3 Gallery of single wavelength images and single differences with a nominal contrast level of 108. All images are shown with the same logarithmic stretch from contrasts of 0 to 105. The single subtractions suppress the speckles outside the dark hole by a factor of 5 - 50. 172

Figure 5.4 Gallery of single wavelength images and single differences with a nominal contrast level of 109. All images are shown with the same logarithmic stretch from contrasts of 0 to 105. The single subtractions suppress the speckles outside the dark hole by a factor of 5 - 50. 173

Figure 5.5 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 106. These images are shown with the same logarithmic stretch from contrasts of 0 to 105 as Fig. 5.1. Significant speckle residuals are present compared to the single differenced im- ages. In contrast to theory (Marois et al. 2000), the double difference method is less effective at suppressing speckle noise than a single difference. 174

Figure 5.6 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 107. These images are shown with the same logarithmic stretch from contrasts of 0 to 105 as Fig. 5.1. Significant speckle residuals are present compared to the single differenced im- ages. In contrast to theory (Marois et al. 2000), the double difference method is less effective at suppressing speckle noise than a single difference. 175

Figure 5.7 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 108. These images are shown with the same logarithmic stretch from contrasts of 0 to 105 as Fig. 5.1. Significant speckle residuals are present compared to the single differenced im- ages. In contrast to theory (Marois et al., 2000), the double difference method is less effective at suppressing speckle noise than a single difference. 176

Figure 5.8 Double differenced images – (F4 - F3) - (F3 - F2) ((F3 - F2) - (F2 - F1)) and (F5 - F3) - (F3 - F1) for a nominal contrast level of 109. These images are shown with the same logarithmic stretch from contrasts of 0 to 105 as Fig. 5.1. Significant speckle residuals are present compared to the single differenced im- ages. In contrast to theory (Marois et al., 2000), the double difference method is less effective at suppressing speckle noise than a single difference. 177

Figure 5.9 Boxes used for speckle RMS calculation (Table 2) and trajectories for contrast plots (Fig. 5.11 through Fig. 5.14). 178

Figure 5.10 Contrast as a function of wavelength ( in nm from 798.9 nm) in- side and outside of the dark hole. To compare RMS in different regions of the chip, RMS values have been scaled to the value at 798.9 nm. RMS increases by a factor of 2 at wavelength differences of 30 nm from 798.9 nm. Thus, contrast will decrease by a similar factor at these wavelengths. The increase in RMS as a function of wavelength is most pronounced at the highest contrast level (109), where amplitude error in the wavelength plays an important role in the speckle noise left after nulling. 179

Figure 5.11 Contrast plots inside and outside of the dark hole for single filter images and our single differenced images at a nominal contrast level of 106. The contrast plots on the left follow trajectories through the nulled right side of the dark hole while the contrast plots on the right show trajectories through the uncorrected left side of the image. Trajectories used are shown in Fig. 5.9. At this contrast level, the single difference subtraction suppresses speckles by a factor of 5 to 50 in the contrast plots both inside and outside of the dark hole. In the dark hole, speckle noise increases slightly with increasing from the nulled wavelength (thus the filter subtractions which minimize are slightly better than those with larger values of .) 180

Figure 5.12 Contrast plots inside and outside of the dark hole for single filter images and our single differenced images at a nominal contrast level of 107. The contrast plots on the left follow trajectories through the nulled right side of the dark hole while the contrast plots on the right show trajectories through the uncorrected left side of the image. Trajectories used are shown in Fig. 5.9. At this contrast level, the single difference subtraction suppresses speckles by a factor of 5 to 50 in the contrast plots both inside and outside of the dark hole. In the dark hole, speckle noise increases slightly with increasing from the nulled wavelength (thus the filter subtractions which minimize are slightly better than those with larger values of .) 181

Figure 5.13 Contrast plots inside and outside of the dark hole for single filter images and our single differenced images at a nominal contrast level of 108. The contrast plots on the left follow trajectories through the nulled right side of the dark hole while the contrast plots on the right show trajectories through the uncorrected left side of the image. Trajectories used are shown in Fig. 5.9. At these contrasts, the single difference subtraction suppresses speckles by a factor of 5 to 40 in the contrast plots outside of the dark hole. In the dark hole, speckle noise increases strongly with increasing from the nulled wavelength. Inside the dark hole, for 16 nm, a single difference still attenuates speckle noise by a factor of 10, however, for 20 nm, speckle patterns between filters are different enough that little speckle attenuation is achieved in a single difference. 182

Figure 5.14 Contrast plots inside and outside of the dark hole for single filter images and our single differenced images at a nominal contrast level of 109. The contrast plots on the left follow trajectories through the nulled right side of the dark hole while the contrast plots on the right show trajectories through the uncorrected left side of the image. Trajectories used are shown in Fig. 5.9. At this contrast level, the single difference subtraction suppresses speckles by a factor of 5 to 50 in the contrast plots outside of the dark hole. However, inside the dark hole, chromatic variation of the speckles between filters is considerable – by doing a single difference, we end up ”importing” in speckles from other wavelengths (i.e. the 800 nm - 832 nm subtraction introduces speckles from the 832 nm image into the subtraction). 183

CHAPTER 6

CONCLUSIONS AND FUTURE DIRECTIONS

We summarize the results of each chapter of this thesis and then discuss applica- tions of these results to future work.

6.1 Conclusions from the SDI Imaging Survey for Extrasolar Planets and Rami- fications for Future Planet Imaging Surveys

As part of the SDI Imaging Survey, we obtained SDI datasets for 54 stars (45 stars were observed in the southern sky at the VLT, 11 stars were observed in the northern sky at the MMT, and 2 stars were observed at both telescopes). In our VLT data, we achieved H band contrasts > 10 mag (5) at a separation of 1.0” from the primary star on 45% of our targets and H band contrasts of > 9 mag at a separation of 0.5” on 80% of our targets. With this degree of attenuation, we should be able to image (5 detection) a 7 MJup planet 15 AU from the star around a 70 Myr K1 star at 15 pc or a 7.8 MJup planet at 2 AU from a 12 Myr M star at 10 pc. We believe that our SDI images are the highest contrast astronomical images ever made from ground or space for methane rich companions within 1” of their primary star. Eight tentative candidates were identified (none with S/N > 2 ). Had these candidates been real, they would have possessed separations of 3 - 15.5 AU and masses of 2-10 MJup. However, none of the candidates were detected in second epoch observations. Thus, we find a null result from our survey. Nonetheless, our result still has serious implications for the distribution of extrasolar planets. In the course of our survey, we also discovered 5 new close stellar binary systems with measured separations of 0.14” to 0.26”. 184

For 20 of our survey stars, we attained 50% completeness for 4-8 MJup plan- ets at semi-major axes of 20-40 AU. Thus, our completeness levels are sufficient to significantly test theoretical planet distributions. From our survey null result, we can rule out (at the 90% level) a model planet population using a constant distri- bution ( dN a0) of planet semi-major axis out to a distance of 45 AU. In Nielsen da ∝ et al. (submitted), we consider a number of other possible model planet popula- tions and examine the implications for the distribution of extrasolar planets based on the null results from the SDI survey combined with the VLT NACO adaptive optics deep imaging survey of Masciadri et al. (2005). Combining the measured contrast curves from 23 of the stars observed with the VLT NACO adaptive op- tics system by Masciadri et al. (2005), and 47 of the stars observed with the VLT NACO SDI and MMT SDI devices by Biller et al. (2007) (for a total of 59 unique stars), we consider what distributions of planet masses and semi-major axes can be ruled out by these data, based on Monte Carlo simulations of planet popula- tions. We can set this upper limit with 95% confidence: the fraction of stars with

planets with semi-major axis from 20 to 100 AU, and mass >4 MJup, is 20% or less. We also discovered a very interesting nearby T5.5 type brown dwarf with the SDI device. SCR 1845B is the brightest mid-T dwarf yet discovered. In addition, it is the first T dwarf companion found around a low mass star. At only 4.5 AU from its primary, it is one of the tightest known brown dwarf companions to a star and is a further piece of evidence that the brown dwarf desert does not exist for companions to very low mass stars (Close et al. , 2003; Gizis et al., 2003). Both the primary and secondary mass can be accurately measured within a decade. SCR 1845B is also proof of concept that the SDI device works. Indeed, we could have detected SCR 1845B at just 0.5 AU from its parent star with SDI. 185

The results from the SDI survey will be immediately important for the Near Infrared Coronagraphic Imager (henceforth NICI) science campaign at Gemini South, a deep 50-night survey to directly image extrasolar planets slated to begin in fall 2007. In many ways, the NICI science campaign will serve as a larger scale continuation of the SDI survey. NICI utilizes the same spectral differential speckle-attenuation technique as SDI but can use both the 1.6 m and 2.2 m methane absorption features. Additionally, NICI possesses a coronagraph and is planned to operate at higher Strehl ratios than SDI. I will be adapting the SDI pipeline and database tools I have developed for SDI to use with NICI and the NICI campaign. The NICI science campaign survey directly picks up where the SDI young star survey leaves off. With 150 target stars already selected, Monte Carlo modeling (Nielsen et al. 2007, similar to our SDI Monte Carlo modeling) predicts 12 de- tected planets as opposed to 3 for the SDI survey (assuming every star has one giant planet.) While the SDI survey null result is just barely statistically relevant, a null result for the NICI science campaign would have a strong statistical relevance and would drastically constrain the distribution of extrasolar giant planets.

6.2 Development of Technology for Future Planet Searches

6.2.1 High Strehls

We present the first high resolution (0.100) very high Strehl ratio (0.97 0.03) mid- IR images of RV Boo utilizing the MMT adaptive secondary AO system. These are the first ground-based, large telescope IR images presented in the literature with such high Strehl ratios; previous Strehl ratios for large telescopes have hardly exceeded 70% at any wavelength. RV Boo was observed at a number of wave- lengths over two epochs (9.8 m in May 2003, 8.8, 9.8 and 11.7 m in February 186

2004) and appeared slightly extended at all wavelengths. While the extension is within the measurement error for the 8.8 and 11.7 m data, the extension is more pronounced in the 9.8 m data. The slight extension seen in the 9.8 m data from both May 2003 and February 2004 suggests that the mid-infrared structure around RV Boo is marginally resolved at 9.8 m. Because of our high Strehl ratios which leads to extremely stable PSFs, we can deconvolve our images with those of PSF stars for a super-resolution of 0.100(3 better than D ). Based on the rotation of the extension with the sky and its non-PSF FWHM (at 8.8-11.7 m) and eccentricity, we conclude that the extension around RV Boo is indeed real and not an artifact of the telescope. These very high Strehl ratio images also allow us a glimpse into the future of planet-finding, where high- order AO systems such as the planned AO system for GPI, will make such high Strehl ratios routine at near-IR wavelengths as well.

6.2.2 High Strehls and High Contrasts

SDI on ground based telescopes provides significant speckle attenuations down to star-planet contrasts of 1-3 104. To test the classical SDI technique at con- trasts of 1069 and to develop a similar “2nd order” SDI technique incorporating Fresnel propagation in the speckle solution, as well as determine whether such a technique would be applicable for the Terrestrial Planet Finder mission, we im- plemented a similar multiwavelength differential imaging scheme for the High Contrast Imaging Testbed at the Jet Propulsion Laboratory. Five filters were se- lected near the prominent O2(A) absorption feature at 0.762 m (seen in Earth’s atmosphere and expected for any terrestrial extrasolar planet with an oxygen at- mosphere, Woolf et al. (2002), with approximate central wavelengths of F1(768 nm), F2(784 nm), F3(800 nm), F4(816 nm), and F5(832 nm). Exact filter wave- lengths and bandwidths are presented in Table 5.1 – however, approximate wave- 187

lengths are used to refer to the filters throughout the text). Two sets of images were taken in each filter – a long set, with 90 s exposure time per filter, and a short set, with 5 s exposure time per filter. The filter set was spaced across a con- siderable bandwidth in wavelength in order to measure speckle chromatism as a function of wavelength. For ground based observing, simultaneous imaging in several bandwidths is necessary to overcome the stochastic speckle noise floor remaining even af- ter adaptive optics correction. For space-based observing, however, speckles are stable on timescales of hours to days, making simultaneity of imaging unneces- sary. This multi-wavelength differential imaging experiment measures speckle evolution as a function of wavelength and contrast level. We tested whether the ground-based simultaneous differential imaging technique can be generalized to a non-simultaneous differential imaging technique for a space mission. At nominal contrasts of 106 to 109, single differences of filter images can reduce speckle noise outside of the dark hole by factors of 5 - 50. For contrasts of 106 to 108, a similar result is also found within the dark hole, with speckle atten- uation achieved through a single difference decreasing as a function of increasing contrast. However, at high contrasts (109), considerable increase in speckle RMS between filters is observed to “pollute” the dark hole in all of our single differ- ences (where, in each difference, only one of the two wavelengths had undergone optical speckle nulling). At all contrast levels, a double difference of images does not seem to decrease speckle noise relative to the single differences. Significant differences in speckle RMS ( 2-10 RMS) is found between filters separated in wavelength by >20 nm. Speckle RMS between filters increases strongly as a function of increasing contrast. We find that the differential imaging technique appears to work equally well 188

both in the pure phase error regime (contrasts less than 107) and in the phase + amplitude error regime (contrasts above 107, where the dark hole becomes asymmetric) before being ultimately limited by the chromaticity of the occulter (in this case due to HEBES glass). We are currently working on modeling meth- ods to remove occulter chromaticity effects from our images – once we have cor- rected for this chromaticity, we expect the differential imaging technique to con- tinue to yield contrast improvements at dark hole contrasts of 109 and above. 189

REFERENCES

Alcala, J. M., Krautter, J., Schmitt, J. H. M. M., Covino, E., Wichmann, R., & Mundt, R. 1995, A&AS, 114, 109

Balick, B., & Frank, A. 2002, ARA&A, 40, 439

Baraffe, I., Chabrier, G., Barman, T.S., Allard, F., & Hauschildt, P.H. 2003, A&A, 402, 701

Bely, P. 1993, STScI report SESD-93-16 (Baltimore)

Benedict, G. F., et al. 2006, AJ, 132, 2206

Bergman, P., Kerschbaum, F., & Olofsson, H. 2000, A&A, 353, 257

Berton et al. 2006, PASP, 118, 1144

Bidelman, W. P. 1951, ApJ, 113, 304

Biller, B. A., Close, L. M., Lenzen, R., Brandner, W., McCarthy, D., Nielsen, E., Kellner, S., & Hartung, M. 2006, IAU Colloq. 200: Direct Imaging of Exoplanets: Science & Techniques, 571

Biller, B.A., Kasper, M., Close, L.M., Brandner, W., & Kellner, S. 2006, ApJ, 641, L141

Biller, B. A., et al. 2006, Proc. SPIE, 6272, 74

Bohren, C.F., & Huffman, D.R. 1983, Absorption and Scattering of Light by Small Particles (New York: Wiley)

Boss, A. P. 2003, ApJ, 599, 577 190

Brandeker, A., Jayawardhana, R., Ivanov, V. D., & Kurtev, R. 2006, ApJL, 653, L61

Brandner et al. 2000, AJ, 120, 950

Brusa, G. et al. 2003, Proc. SPIE 4839, 691

Bujarrabal, V., Neri, R., Alcolea, J., & Kahane, C. 2003, A&A, 409, 573

Burgasser, A. 2001, Caltech PhD Thesis

Burgasser, A. et al. 2003, AJ, 125, 850

Burrows, A., Hubbard, W. B., Lunine, J. I., & Liebert, J. 2001, Reviews of Modern Physics, 73, 719

Burrows, A., Sudarsky, D. & Lunine, J. 2003, ApJ, 596, 587

Butler, R. P., et al. 2006, ApJ, 646, 505

Chauvin, G., Lagrange, A.-M., Dumas, C., Zuckerman, B., Mouillet, D., Song, I., Beuzit, J.-L., & Lowrance, P. 2005, A&A, 438, L25

Chauvin, G., et al. 2005, A&A, 438, L29

Christian, D. J., & Mathioudakis, M. 2002, AJ, 123, 2796

Close, L.M., Siegler, N., Freed, M., & Biller, B.A. 2003, ApJ, 587, 407

Close, L.M., Lenzen, R., Biller, B.A., Brandner, W. & Hartung, M. 2005, Science with AO ESO Workshop, 136

Close, L.M. et al. 2005, Nature, 433, 286

Close, L.M. 2007, ApJ in press, astro-ph/0608574

Close, L.M. 2007, ApJ in press, astro-ph/0703564 191

Codona, J. L., & Angel, R. 2004, ApJL, 604, L117

Cowley, A. P., Hiltner, W. A., & Witt, A. N. 1967, AJ, 72, 1334

Cruz, K.L., Reid, N.I., Liebert, J., Kirkpatrick, J.D., & Lowrance, P.J. 2003, AJ, 126, 2421

Cutispoto, G., Pallavicini, R., Kuerster, M., & Rodono, M. 1995, A&A, 297, 764

Cutri, R.M. et al. 2003, 2MASS All-Sky Catalog of Point Sources

Deacon, N.R. et al. 2005, AJ, 129, 409

Dohlen, K., Beuzit, J.-L., Feldt, M., Mouillet, D., Puget, P., Antichi, J., Baruffolo, A., Baudoz, P., Berton, A., Boccaletti, A., Carbillet, M., Charton, J., Claudi, R., Downing, M., Fabron, C., Feautrier, P., Fedrigo, E., Fusco, T., Gach, J.-L., Grat- ton, R., Hubin, N., Kasper, M., Langlois, M., Longmore, A., Moutou, C., Pe- tit, C., Pragt, J., Rabou, P., Rousset, G., Saisse, M., Schmid, H.-M., Stadler, E., Stamm, D., Turatto, M., Waters, R., & Wildi, F. 2006, Proc. SPIE, 6269, 24

Draine, B.T., & Lee, H.M. 1984, ApJ, 285, 89

Draine, B.T., & Li, A. 2001, ApJ, 551, 807

Favata, F., Barbera, M., Micela, G., & Sciortino, S. 1995, A&A, 295, 147

Favata, F., Micela, G., & Sciortino, S. 1997, A&A, 326, 647

Fischer, D. A., & Valenti, J. 2005, ApJ, 622, 1102

Geballe, T.R. et al. 2002, ApJ, 564, 466

Geißler, K., et al. 2007, A&A, 461, 665

Gizis, J.E. et al. 2003, AJ, 125, 3302 192

Gliese, W., & Jahreiß, H. 1979, A&AS, 38, 423

Gliese, W., & Jahreiß, H. 1991, On: The Astronomical Data Center CD-ROM: Selected Astronomical Catalogs, Vol. I; L.E. Brotzmann, S.E. Gesser (eds.), NASA/Astronomical Data Center, Goddard Space Flight Center, Greenbelt, MD

Gray, R. O., Corbally, C. J., Garrison, R. F., McFadden, M. T., & Robinson, P. E. 2003, AJ, 126, 2048

Gray, R. O., Corbally, C. J., Garrison, R. F., McFadden, M. T., O’Donoghue, A. A., Knox, E. R., McGahee, C. E., & Bubar, E. J. 2004, Bulletin of the American As- tronomical Society, 36, 1488

Gray, R. O., Corbally, C. J., Garrison, R. F., McFadden, M. T., Bubar, E. J., McGahee, C. E., O’Donoghue, A. A., & Knox, E. R. 2006, AJ, 132, 161

Hambly, N.C., Henry, T.J., Subasavage, J.P., Brown, M.A., & Jao, W-C. 2004, AJ, 128, 437

Hartung, M. et al. 2004, Astronomy and Astrophysics, 421, L17

Henry, T.J. et al. 2004, AJ, 128, 246

Henry, T.J. et al. 2006, AJ, 132, 2360

Hinkley, S., et al. 2007, ApJ, 654, 633

Hinz, P.M. et al. 2000, SPIE 4006, 349

Hoffman, W. et al. 1998, SPIE 3354, 647 193

Houk, N., & Cowley, A. P. 1975, “Michigan Catalog of Two-Dimensional Spectral Types for the HD Stars”, Ann Arbor: University of Michigan, Department of Astronomy, 1975

Houk, N. 1978, Ann Arbor : Dept. of Astronomy, University of Michigan : dis- tributed by University Microfilms International

Houk, N. 1982, Michigan Spectral Survey, Ann Arbor, Dep. Astron., Univ. Michi- gan, 3 (1982), 0

Houk, N., & Cowley, A. P. 1988, “Michigan Catalog of Two-Dimensional Spectral Types for the HD Stars, Vol. 4”, Ann Arbor: University of Michigan, Depart- ment of Astronomy

Houk, N., & Smith-Moore, M. 1988, Michigan Spectral Survey, Ann Arbor, Dept. of Astronomy, Univ. Michigan (Vol. 4) (1988), 0

Ida, S., & Lin, D. N. C. 2005, ApJ, 626, 1045

Janson et al. 2007, AJ, subm.

Jeffers, H.M. & Vasilevskis, S. 1978, AJ, 83, 411

Jura, M., Balm, S.P., & Kahane, C. 1995, ApJ, 453, 721

Jura, M., & Kahane, C. 1999, ApJ, 521,302

Kaisler, D., Zuckerman, B., & Becklin, E. 2003, ASP Conference Series, 294, 91

Kahane, C., & Jura, M. 1996, A&A, 310, 952

Kahane, C., Barnbaum, C., Uchida, K., Balm, S.P., & Jura, M. 1998, ApJ, 500, 466 194

Kasper, M., Biller, B., Burrows, A., Brandner, W., Budaj, J. & Close, L. 2007, A&A, accepted

Kellner et al. 2007, submitted

Kenworthy, M.A., Hinz, P.M., Angel, J.R.P, Heinze, A.N., & Sivanandam, S. 2006, Proc. SPIE, 6272, 104

Kerschbaum, F., & Hron, J. 1992, A&A, 263, 97

Kerschbaum, F., & Oloffson, H. 1999, A&AS, 138, 299

Knapp, G.R., Young, K., Lee, E., & Jorissen, A. 1998, ApJS, 210, 231

Knapp, G. R. et al. 2004, ApJ, 127, 3553

Kurucz, R.L. 1979, ApJS, 40, 1

Lachaume, R., Dominik, C., Lanz, T., & Habing, H. J. 1999, A&A, 348, 897

Leaton, B. R., & Pagel, B. E. J. 1960, MNRAS, 120, 317

Lenzen, R. et al. 2003, Proc. SPIE, 4841, 944

Lenzen, R., Close, L., Brandner, W., Hartung, M., & Biller, B. 2004, Proc. SPIE, 5492, 970

Lenzen, R., Close, L., Brandner, W., Biller, B., & Hartung, M. 2005, Science with AO ESO Workshop, 46

Li, J. Z., & Hu, J. Y. 1998, A&AS, 132, 173

Liu, M. et al. 2005, NICI campaign science proposal

Liu, W. et al. 2006, accepted to ApJ 195

Lowrance, P. J., et al. 2005, AJ, 130, 1845

Luhman, K.L., et al. 2006, ApJ, 649, 894

Maas, T., Van Winckel, H., Lloyd Evans, T., Nyman, L.-A., Kilkenny, D., Martinez, P., Marang, F., & van Wyk, F. 2003, A&A, 405, 271

Macintosh, B., Graham, J., Palmer, D., Doyon, R., Gavel, D., Larkin, J., Oppen- heimer, B., Saddlemyer, L., Wallace, J., Bauman, B., Evans, J., Erikson, D., Morzinksi, K., Phillion, D., Poyneer, L., Sivaramakrishnan, A., Soummer, R., Thibault, S., & Veran, J.-P. 2006, Proc. SPIE, 6272, 18

Marley, M.S., Fortney, J.J., Hubickyj, O., Bodenheimer, P., and Lissauer, J.J. 2006, astro-ph/0609739

Marois, C., Doyon, R., Racine, R., & Nadeau, D. 2000, PASP, 112, 91

Marois, C., Doyon, R., Racine, R., & Nadeau, D. 2002, Proc. SPIE, 4008, 788

Marois, C. et al. 2005, PASP, 117, 745

Marois, C., Lafrenier` e, D., Doyon, R., Macintosh, B., & Nadeau, D. 2006, ApJ, 641, 556

Masciadri, E., Mundt, R., Henning, T., Alvarez, C., & Barrado y Navascues, D. 2005, ApJ, 625, 1004

Mason, B.D., Wycoff, G.L., Hartkopf, W.I., Douglass, G.G., & Worley, C.E. 2001, AJ, 122, 3466

Mastrodemos, N. & Morris, M. 1998, ApJ, 497, 303

Mastrodemos, N. & Morris, M. 1998, ApJ, 523, 357 196

McCaughrean, M.J., Close, L.M., Scholz, R.-D., Lenzen, R., Biller, B., Brandner, W., Hartung, M., & Lodieu, N. 2004, Astronomy and Astrophysics, 413, 1029

Meixner, M., Ueta, T., Dayal, A., Hora, J.L., Fazio, G., Hrivnak, B.J., Skinner, C.J., Hoffman, W., & Deutsch, L.K.)] 1999, ApJS, 122, 221

Metchev, S. A., & Hillenbrand, L. A. 2004, ApJ, 617, 1330

Meyer, M. R., Edwards, S., Hinkle, K. H., & Strom, S. E. 1998, ApJ, 508, 397

Mochnacki, S.W. et al. 2002, AJ, 124, 2868

Montes, D., Lopez-Santiago,´ J., Galvez,´ M. C., Fernandez-Figuer´ oa, M. J., De Cas- tro, E., & Cornide, M. 2001, MNRAS, 328, 45

Morris, M. 1987,PASP,99,1115

Mugrauer, M. & Neuhauser¨ , R. 2005, MNRAS, 361, L15

Nakajima, T., Oppenheimer, B.R., Kulkarni, S., Golimowski, D.A., Matthews, K., & Durrance, S.T. 1995, Nature, 378, 463

Neuhauser¨ , R., & Brandner, W. 1998, A&A, 330, L29

Nielsen, E. L., Close, L. M., Guirado, J. C., Biller, B. A., Lenzen, R., Brandner, W., Hartung, M., & Lidman, C. 2005, Astronomische Nachrichten, 326, 1033

Nielsen, E. L., Close, L. M., & Biller, B. A. 2006, IAU Colloq. 200: Direct Imaging of Exoplanets: Science & Techniques, 111

Noyes, R. W., Weiss, N. O., & Vaughan, A. H. 1984, ApJ, 287, 769

Olofsson, H., Gonzalez Delgado, D., Kerschbaum, K., & Shoier¨ , F.L. 2002, A&A, 391, 1053 197

Perryman, M.A.C. et al. 1997, ApJ, 323, 49

Racine, R., Walker, G.A.H., Nadeau, D., Doyon, R., & Marois, C. 1999, PASP, 111, 587

Randich, S., Gratton, R., & Pallavicini, R. 1993, A&A, 273, 194

Reipurth, B. & Clarke, C. 2002, AJ, 122, 432

Renzini, A. 1981, in Physical Processes in Red Giants, ed. I. Iben & A. Renzini (Dordrecht: Reidel), 431

Richardson, L. J., Deming, D., Horning, K., Seager, S., & Harrington, J. 2007, Na- ture, 445, 892

Roddier, F., Roddier, C., Graves, J.E., & Northcott, M.J. 1995, ApJ, 443, 249

Rousset, G. et al. 2003, Proc. SPIE, 4389, 140

Schneider, G., Becklin, E., Close, L., Figer, D., Lloyd, J., Macintosh, B., Hines, D., Max, C., Potter, D., Rieke, M., Scoville, N., Thompson, R., Weinberg, A., & Windhorst, R. 2003, solicited by STScI in preparation for HST Cycle 12

Scholz, R.-D., McCaughrean, M.J., Lodieu, N., & Kuhlbrodt, B. 2003, A&A, 398, L29

Sparks, W. B., & Ford, H. C. 2002, ApJ, 578, 543

Sozzetti, A., et al. 2006, A&A, 449, 417

Steffen, M., Szczerba, R., & Schonberner¨ , D. 1998, A&A, 337, 149

Tamura, M. & Lyu, A. 2006, IAUC 200 proceedings

Thatte, N. et al. 2007, MNRAS, submitted 198

Thorstensen, J.R. 2001, Skycalc User’s Manual, (Tucson:NOAO, contributed soft- ware)

Torres, G., Stefanik, R., Latham, D., & Mazeh, T. 1995, Astrophysical Journal, 452, 870

Trauger, J.T., Burrows, C., Gordon, B., Green, J.J., Lowman, A.E., Dwight, M., Niessner, A.F., Shi, F., & Wilson, D. 2004, Proc. SPIE, 5487, 1330

Trauger, J.T. & Traub, W.A. 2007, Nauture, 7137, 771

Ueta, T., Meixner, M., & Bobrowsky, M. 2000, ApJ, 528, 861

Van Winckel, H., Waelkens, C., Waters, L.B.F.M., Molster, F.J., Udry, S. & Bakker, E.J. 1998, A&A, 336, L17

Vinkovic, D., Blocker, T., Hofmann, K.-H., Elitzur, M., & Weigelt, G. 2004, MN- RAS, 352, 852

Vrba et al. 2004, AJ, 129, 2948

Wallace, L., Hinkle, K., & Livingston, W. 1993, An of the Photospheric Spec- trum from 8900 to 13600 cm1 (7350 to 11230 angstroms), NSO Technical Report #93-001

Wichmann, R., Schmitt, J. H. M. M., & Hubrig, S. 2003, A&A, 400, 293

Wildi, F., Brusa, G., Riccardi, A., Lloyd-Hart, M., Martin, H.M., & Close, L.M. 2003, proc. SPIE 4839, 155

Woolf, N.J., Smith, P.S., Traub, W.A., & Jucks, K.W. 2002, ApJ, 574, 430

Wright, J. T., Marcy, G. W., Butler, R. P., & Vogt, S. S. 2004, ApJS, 152, 261 199

Wright, J. T., Butler, R. P., Marcy, G. W., Vogt, S. S., Fischer, D. A., Rinney, C. G., & Jones, H. R. A. 2005, and Planets V, 8605

Zuckerman, B., Song, I., Bessell, M. S., & Webb, R. A. 2001, ApJL, 562, L87

Zuckerman, B., Webb, R. A., Schwartz, M., & Becklin, E. E. 2001, ApJL, 549, L233

Zuckerman, B., & Song, I. 2004, ARA&A, 42, 685