Articles

STIS CORONAGRAPHIC IMAGING OF FOMALHAUT: MAIN BELT STRUCTURE AND THE ORBIT OF FOMALHAUT b

, , , and

Published 2013 September 4 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Paul Kalas et al 2013 ApJ 775 56 DOI 10.1088/0004-637X/775/1/56

0004-637X/775/1/56

ABSTRACT

We present new optical coronagraphic data of Fomalhaut obtained with HST/STIS in 2010 and 2012. Fomalhaut b is recovered at both epochs to high significance. The observations include the discoveries of tenuous nebulosity beyond the main dust belt detected to at least 209 AU projected radius, and a ∼50 AU wide azimuthal gap in the belt northward of Fomalhaut b. The two epochs of Space Telescope Imaging Spectrograph (STIS) photometry exclude optical variability greater than 35%. A Markov chain Monte Carlo analysis demonstrates that the orbit of Fomalhaut b is highly eccentric, with e = 0.8 ± 0.1, a = 177 ± 68 AU, and q = 32 ± 24 AU. Fomalhaut b is apsidally aligned with the belt and 90% of allowed orbits have mutual inclination ⩽36°. Fomalhaut b's orbit is belt crossing in the sky plane projection, but only 12% of possible orbits have ascending or descending nodes within a 25 AU wide belt annulus. The high eccentricity invokes a dynamical history where Fomalhaut b may have experienced a significant dynamical interaction with a hypothetical planet Fomalhaut c, and the current orbital configuration may be relatively short-lived. The Tisserand parameter with respect to a hypothetical Fomalhaut planet at 30 AU or 120 AU lies in the range 2–3, similar to highly eccentric dwarf planets in our solar system. We argue that Fomalhaut b's minimum mass is that of a dwarf planet in order for a circumplanetary satellite system to remain bound to a sufficient radius from the planet to be consistent with the dust scattered light hypothesis. In the coplanar case, Fomalhaut b will collide with the main belt around 2032, and the subsequent emergent phenomena may help determine its physical nature.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The number of candidate exoplanets detected by direct imaging techniques has recently surpassed the number of solar system planets. In every case there are a significant number of challenges in observation and interpretation. Observationally, a faint companion to a star or brown dwarf must be shown to be a real astrophysical feature instead of an instrumental artifact. Precision astrometry over multiple epochs of observations is also required to establish common proper motion and orbital motion. Complementary data and analyses are essential for estimating the age of the system, which is required to constrain the mass of the object from the theory of planet luminosity evolution as well as dynamics. And finally, even though an object may pass all of these tests, appear to be a bound companion, and have mass under the brown dwarf mass limit of ∼13 MJ, the question of planet identity may persist, given open questions such as whether or not a planet mass object needs to be bound to a star (as opposed to a brown dwarf), or if the mode of formation should be critical to the definition of a planet (Basri & Brown 2006; Soter 2006; Schneider et al. 2011).

The Fomalhaut system certainly presents a number of open questions and puzzling observations. The discovery of infrared excess due to circumstellar dust was firmly established by a series of infrared and submillimeter observations (Backman & Gillett 1987; Zuckerman & Becklin 1993). Resolved thermal infrared images at 850 μm demonstrated that the debris disk was in fact a torus of material in a region 100–140 AU radius from the star (Holland et al. 1998; Dent et al. 2000). A dust torus could be sustained over the age of the system if planet mass objects served to dynamically halt or delay the inward drift of grains governed by Poynting–Robertson drag (Ozernoy et al. 2000; Moro-Martin & Malhotra 2002). Higher resolution thermal images at 450 μm suggested the presence of warmer dust concentrated within 100 AU radius to the southeast of the star in an arc-like morphology (Wyatt & Dent 2002; Holland et al. 2003).

Motivated by these findings, the first Hubble Space Telescope (HST) coronagraphic observations of Fomalhaut were intended to search for a planet mass object within the dust belt using the Advanced Camera for Surveys High Resolution Channel (ACS/HRC; GO 9862; PI: Kalas). These relatively shallow observations with the F814W filter (0.8 μm) yielded the first detection of the dust belt in optical scattered light. Deeper, follow-up observations (GO 10390; PI: Kalas) were conducted in both the F606W (0.6 μm) and F814W filter in 2004.

The first-epoch, 2004 data revealed that the geometric center of the dust belt is offset to the northwest of the stellar position by ∼15 AU, and the inner edge of the belt is consistent with a knife-edge. Kalas et al. (2005) proposed that these two facts were robust indirect evidence for a planet orbiting interior to the 133 AU inner border of the belt. At roughly the same time, new thermal infrared images using the Spitzer Space Telescope and the Caltech Submillimeter Observatory revealed asymmetry in the disk emission due to the fact that the southeast side of the belt is significantly closer to the star than the northwest side (Stapelfeldt et al. 2004; Marsh et al. 2005).

To search for the putative planet and determine the scattered light colors of the belt, deeper coronagraphic images with HST in three filters (F435W, F606W and F814W) were obtained. The second-epoch, 2006 observations resulted in the discovery of a point source 18 AU interior to the inner border of the belt (Kalas et al. 2008). The point source was verified in multiple data sets, in two filters, and was detected in the F606W data obtained in 2004. Due to the high proper motion of Fomalhaut (0farcs4 yr−1), the point source was easily separated from background stars, but a small offset between epochs suggested orbital motion in the counterclockwise direction as projected on the sky. Thus, the point source was physically associated with the central star and designated Fomalhaut b.

A key surprise with the Fomalhaut b observations was the unexpected blue color (i.e., an unexpectedly high luminosity at optical wavelengths). Kalas et al. (2008) presented ground-based observations at 1.6 μm and 3.8 μm that did not detect Fomalhaut b, establishing that non-thermal sources probably contribute a fraction of its 0.6 μm brightness. They proposed that light reflected from a circumplanetary dust disk could account for the visible-light flux, though the data also indicated a dimming of Fomalhaut b between 2004 and 2006 in the 0.6 μm detections. If the 0.8 μm flux was entirely attributed to thermal emission, then the mass of Fomalhaut would be <3 MJ. This upper limit to the mass was also consistent with dynamical theory that showed the inner edge of the dust belt could not reside as close as 18 AU from a planet unless the planet mass is below a few Jupiter masses (Quillen 2006; Chiang et al. 2009).

An alternative model studied quantitatively in Kalas et al. (2008) is that Fomalhaut b is a rare and short-lived dust cloud that has recently appeared in the system due to the collision of two planetesimals (see also Currie et al. 2012; Galicher et al. 2013). The hypothetical conversion of a 10 km radius planetesimal into 0.1–0.2 μm water ice grains represented the minimum mass (4.1 × 1018 g) that could explain the optical photometry in terms of reflected light. An alternate water ice cloud model that assumes a grain size distribution between 0.08 μm and 1 mm corresponds to a 67 km radius planetesimal (1.2 × 1021 g). However, the scenario of a disrupted planetesimal was deemed less likely than the planetary rings hypothesis because: (1) observing a rare and short-lived event is unlikely; (2) planetesimal collisions are unlikely far from the star where dynamical timescales are relatively long (P ∼ 800 yr); (3) modeling the dimming of the dust cloud requires a fine-tuning of the model such that small-grains are quickly depleted from the cloud just as the observations are conducted; and (4) the simulated dust cloud predicts optical colors that do not agree with the observed color.

A new model was proposed by Kennedy & Wyatt (2011) where they examined the origin and collisional evolution of irregular satellites orbiting solar system planets. The collisional erosion of irregular satellites can produce an hourglass-shaped dust cloud around a planet rather than a flattened circumplanetary dust disk. When applied to Fomalhaut b, a few lunar-mass planetesimal dust cloud orbiting a 2–100 M planet is consistent with the theoretical assumptions and observational constraints.

Additional mass limits for Fomalhaut b have been established by modeling the non-detection of Fomalhaut b at mid-infrared wavelengths with the Spitzer Space Telescope (Marengo et al. 2009; Janson et al. 2012). Adopting the new age determination of ∼400 Myr for Fomalhaut (Mamajek 2012), Janson et al. estimate that the mass of Fomalhaut b is ≲1 MJ, consistent with previous findings. However, their primary conclusion is that the Kalas et al. (2008) dust-cloud model is the best fit to observations because they claim Fomalhaut b resides out of the sky plane, and this geometry rules out reflection from a circumplanetary disk. However, this conclusion has at least three significant problems. First, they presume the circumplanetary dust is optically thick, when in fact there is no such constraint. An optically thin, circumplanetary dust cloud would still permit forward scattering if the current geometry puts the planet between the host star and the observer. Second, the orientation of planetary ring systems relative to planet orbital planes in our solar system varies in the range 0°–177°. Determining the orbital geometry of Fomalhaut b does not directly translate into knowledge of how the planetary rings are oriented relative to the star and observer. Third, the assumption that Fomalhaut b is situated between the star and observer is a tentative result from Le Bouquin et al. (2009). These authors observed Fomalhaut A with Very Large Telescope optical interferometry and the AMBER instrument, finding that the spin axis of the star extends out of the sky plane in the NE. Given a counterclockwise spin, the western portion of the belt and Fomalhaut b would reside out of the sky plane. However, Le Bouquin et al. (2009, L44) clearly state "...no check star is available in the dataset to secure the sign of the AMBER phase...we cannot draw definite conclusions before a real spectroastrometric reference has been observed."

The tentative geometry suggested by Le Bouquin et al. (2009) means that the brightest hemisphere of the belt resides behind the sky plane and the grain scattering phase function is strongly backscattering. Min et al. (2010) suggest that backscattering can dominate in Fomalhaut's dust belt. However, the backscatter model is consistent with the observations only if all grains smaller than 100 μm are absent from the system. The radiation pressure blowout grain size for Fomalhaut is 8–13 μm, depending on composition and porosity (Artymowicz & Clampin 1997; Acke et al. 2012). Therefore, if one accepts that debris disks are replenished by the collisional erosion of larger bodies, there should be a significant population of grains in the 13–100 μm size range that ensures the belt is dominated by forward scattering.

Specifying the belt geometry by finding the orientation of the stellar spin axis is moot in any case because it has been established that the spin–orbit alignment of exoplanets may be highly oblique or retrograde (Hébrard et al. 2008; Anderson et al. 2010; Johnson et al. 2011). Moreover, debris disks are also known to be misaligned with the stellar angular momentum orientation (Kennedy et al. 2012). Therefore, the spin vectors of the host star(s), exoplanet(s) and debris belt(s) within any given exoplanetary system are not necessarily aligned.

In our view the most significant observable is that a dust belt with an asymmetric scattering phase function will exhibit preferential forward scattering. Therefore the bright side of Fomalhaut's belt is out of the sky plane. Fomalhaut b resides near the faint part of the belt, which is likely behind the sky plane due to weaker backscattering. The assumptions underlying the Janson et al. (2012) argument that hypothetical planetary rings surrounding Fomalhaut b would not be seen in reflected light are therefore unsupported.

Janson et al. (2012) note that the HST detections may not be trustworthy, yet two independent groups have analyzed the same HST data and detected Fomalhaut b (Currie et al. 2012; Galicher et al. 2013). Moreover, these groups claim that Fomalhaut b is detected in a third optical filter, F435W (0.4 μm). Fomalhaut b's optical variability is not confirmed, but the photometry presented in Currie et al. (2012) and Galicher et al. (2013) each span the range of photometry given in Kalas et al. (2008). The photometric uncertainties are evidently greater than previously assumed and the case for variability requires further work (cf. Sections 2.2 and 6).

More recent thermal infrared observations of Fomalhaut's dust belt have been made with ALMA (Boley et al. 2012), Herschel (Acke et al. 2012), and the Australian Telescope Compact Array (Ricci et al. 2012). The physical properties of the dust belt derived from these new observations generally support the results of previous studies. One significant new result from a study focused on the parent star is a revised, older age for Fomalhaut (Mamajek 2012). The previous value of 200 Myr (Barrado y Navascues et al. 1997) is superceded by 440 ± 40 Myr. The older age means that any Fomalhaut planets have a lower temperature for a given mass, which has a bearing on predicted infrared detection limits, and other derivations involving dynamical lifetime arguments must be revised. Mamajek (2012) also calls attention to a likely stellar companion 5.7 × 104 AU from Fomalhaut called TW PsA. Furthermore, Mamajek et al. (2013) discovered that LP 876-10, an M dwarf northwest of Fomalhaut, is a likely third stellar component to the Fomalhaut system.

Therefore, Fomalhaut could be newly designated as Fomalhaut A, Fomalhaut b is Fomalhaut Ab, TW PsA is Fomalhaut B, and LP 876-10 is Fomalhaut C. For the sake of continuity with the prior literature, we continue using the term Fomalhaut b.

In the present paper we describe the results of additional imaging observations of Fomalhaut b using HST/STIS in 2010 and 2012 (Table 1). Our main goal is to derive the orbital elements of Fomalhaut b using astrometric measurements from all four epochs of observation. The most significant challenge is that follow-up observations using the original discovery instrument are precluded due to a failure in the ACS/HRC electronics, which could not be restored during Servicing Mission 4. The HRC was ideally suited for high contrast observations, given fine sampling (25 mas pixels), three coronagraphic occulters in the focal plane, and a Lyot stop which suppressed light diffracted around telescope elements. By changing instruments to Space Telescope Imaging Spectrograph (STIS), we had to accept a broader optical bandpass, with different detector and coronagraphic characteristics. When Fomalhaut b was recovered in 2010 with STIS, the third epoch of astrometry indicated that Fomalhaut b's orbit is not nested within the dust belt (Kalas et al. 2010; Kalas 2011). The preliminary orbit was found to have e ∼ 0.7. However, given that these were the first data obtained with a different instrument, it was difficult to determine if the 2010 position measurement was compromised by uncorrected geometric distortion in STIS, or a systematic uncertainty in the roll angle of the telescope due to single guide star guiding. We therefore chose to wait for a second epoch of STIS observations in 2012 to confirm the new findings of the 2010 epoch.

Table 1. STIS Observations

Date Position ORIENTAT a Exposures
2010 Sep 13 WEDGEB2.5 193fdg04 25 × 30 s
    201fdg04 25 × 30 s
    209fdg04 25 × 30 s
    217fdg05 25 × 30 s
2012 May 29 WEDGEB2.5 −169fdg945 28 × 30 s
    −161fdg945 28 × 30 s
    −153fdg945 28 × 30 s
    −145fdg945 28 × 30 s
2012 May 30 WEDGEB2.5 −167fdg945 28 × 30 s
    −159fdg945 28 × 30 s
    −151fdg945 28 × 30 s
    −143fdg945 28 × 30 s
2012 May 31 WEDGEB2.5 −165fdg945 28 × 30 s
    −157fdg945 28 × 30 s
    −149fdg945 28 × 30 s
    −141fdg945 28 × 30 s

Note. aORIENTAT is the position angle of the image y-axis.

Download table as:  ASCIITypeset image

Here we present Fomalhaut b astrometry from a total of four epochs of HST observations spread over 8 yr, as well as new discoveries concerning the morphology of the dust belt in optical scattered light. In Section 2 we briefly revisit and reanalyze the ACS/HRC photometry and astrometry. The STIS data and results are given in Sections 37. Fomalhaut b's astrometry and newly determined orbital parameters are provided in Section 8. We discuss the implications of the new orbit in Section 9, which includes an observational inventory of the objects and structures comprising the Fomalhaut system (Section 9.1). Discussion of Fomalhaut b's physical nature, dynamical history, relationship to the main belt, and comparison to the solar system are also found in Section 9.

2. ACS/HRC OBSERVATIONS REVISITED

Here we present new work to understand the sources of astrometric and photometric uncertainty in the data previously presented by Kalas et al. (2005, 2008). In both the 2004 and 2006 observations, Fomalhaut was occulted by the HRC 1farcs8 diameter occulting spot, which is located near the center of the field. To expand the field of view, in 2006 we also occulted Fomalhaut behind the 3farcs0 diameter occulting spot, which is closer to the top edge of the detector. All of the observations included multiple telescope roll angles so that the point-spread function (PSF) of Fomalhaut A could be self-subtracted while recovering astrophysical objects in the field. The final PSF-subtracted images for any given sequence of observations include a significant number of residuals that appear as point sources or extended features. The key to separating Fomalhaut b from residual noise features is that the noise features vary significantly in position, morphology and intensity among different observing sequences, as the telescope roll angle changes.

2.1. ACS/HRC Astrometry

A significant source of astrometric uncertainty in the HRC data, as in various other coronagraphic data sets (Digby et al. 2006), is determining the position of the central star behind the occulting spots. The two techniques that we use are to: (1) centroid the core of the stellar image as it appears through the occulting spot, and (2) determine the centroid of the stellar PSF halo seen beyond the outer edge of the occulting spot. The former is possible because the ACS/HRC occulting spot transmits 4.5% of incident light (ACS Instrument Handbook 2005). The 2006 observing strategy included three 0.20 s integrations (the minimum allowable) at the conclusion of every orbit targeting Fomalhaut. In these short exposures, the morphology of the PSF core (after the image distortion correction is applied) is highly asymmetric, with a tail of light extending to the lower left of a distinct few pixel PSF core (Figure 1). This asymmetric morphology is less pronounced for the 3farcs0 spot data.

Figure 1.

Figure 1. Fomalhaut A viewed through the 1farcs8 occulting spot (left) and 3farcs0 occulting spot (right) of the ACS/HRC in the shortest exposure data (0.20 s) from 2006. The color scale is linear with white pixels representing 128,000 electrons for the 1farcs8 spot data and 32,000 electrons for the 3farcs0 spot data. Neither the CCD nor the analog-to-digital conversion is saturated (gain = 2.2 electrons DN−1). The morphology of the 3farcs0 spot data includes a dark band extending downward from the top left of the frame, representing the ACS/HRC occulting bar that resides in the corrected beam between the spot and the focal plane array.

Standard image High-resolution image

For the longer exposures (e.g., 340 s), the core is significantly saturated. Our technique for centroiding the PSF involves rotating the image 180° and subtracting the rotated image from the unrotated image. The center of rotation is adjusted iteratively to symmetrically subtract the PSF. This technique provides two additional centroid positions because we choose two different regions to assess the success of the rotated self-subtraction. The first is the region interior to the occulting spot where the PSF core is saturated, but nevertheless the rotation center can be adjusted so that the core is azimuthally uniformly subtracted. The second is the region exterior to the occulting spot which is not saturated.

To summarize, the three methods for estimating the stellar position in ACS/HRC data use: (1) the PSF core in short exposure data; (2) the PSF core in 180°, self-subtracted, long-exposure data; and (3) the PSF halo in 180°, self-subtracted, long exposure data. The standard deviation of these position measurements is given in Table 2 as the 1σ uncertainty in the stellar location. The astrometry cited in Kalas et al. (2008) utilized only the third method, with an estimated 1σ uncertainty of 12.5 mas. The larger uncertainties in Table 2 demonstrate that the differences between techniques account for additional uncertainty in estimating the stellar position. An important note is that changing the assumed location of the star propagates throughout the data reduction because observations made at different telescope orientations must be rotated to a common orientation based on this stellar location. In effect, measurements made in this manuscript are based on different final versions of the 2004 and 2006 observations and the results are not expected to be identical with Kalas et al. (2008). Table 2 also gives the uncertainty in obtaining the location of Fomalhaut b using a variety of centroiding algorithms (IRAF/pradprofandIDL/IDP3) and applied to different versions of the final, reduced images at each of the epochs. Again the standard deviation is larger than quoted in Kalas et al. (2008) because the latter work adopted only one type of centroiding algorithm (IRAF/pradprof).

Table 2. Astrometric Error Terms

  Value Source  
Uncorrected systematic errors
Geometric distortion STIS 66 mas (not used) Unpublished Calibration Program  
Geometric distortion STIS 17 mas Measured in data  
Stability of optical distortion STIS 5 mas STIS astrometric report  
Geometric distortion and stability ACS/HRC 3 mas Instrument Handbook  
Detector position angle 0fdg06 (1 mas in X, 1 mas in Y) Measured in data (see text)  
Statistical errors
ACS/HRC 2004 centroiding star 16 mas in X, 10 mas in Y Measured in data  
ACS/HRC 2004 centroiding on Fom b 18 mas in X, 13 mas in Y Measured in data  
ACS/HRC 2006 centroiding star 19 mas in X, 17 mas in Y Measured in data  
ACS/HRC 2006 centroiding on Fom b 6 mas in X, 21 mas in Y Measured in data  
STIS centroiding star 25 mas in X, 25 mas in Y Measured in data  
STIS 2010 centroiding Fom b 28 mas in X, 31 mas in Y Measured in data  
STIS 2010 recovery of artificial implants 8 mas in X, 5 mas in Y Measured in data  
STIS 2012 centroiding Fom b 16 mas in X, 19 mas in Y Measured in data  
STIS 2012 recovery of artificial implants 5 mas in X, 10 mas in Y Measured in data  

Download table as:  ASCIITypeset image

Another possible source of astrometric error is the position angle uncertainty of the observations. The lack of guide stars means that Fomalhaut observations are made with single guide star guiding (true for all HST observations discussed in this paper), which means that the telescope roll angle is maintained using the gyros. The drift rate has a typical value of 28 mas orbit−1, with a maximum value of 53 mas orbit−1 (ACS Long Data Handbook Version 5, 2006). Given the focal plane geometry of the ACS/HRC relative to the guide fields, the drift means that the entire HRC field may suffer a translation as high as 53 mas during an orbit with a small change in position angle within the field. In discussion below, we determine an 0fdg06 uncertainty in determining the position angle. Telescope jitter is in the range 3–5 mas.

The positions of Fomalhaut b in 2004 and 2006 and the associated uncertainties are given in Table 3. Compared to Kalas et al. (2008) these measurements are within their mutual 1σ error bars. The greatest difference lies in the revised 2004 position, which in this manuscript is 36 mas eastward (i.e., closer to the star) and 26 mas south of the position published in Kalas et al. (2008). As noted above, the 2004 data were re-reduced based on new estimates of the stellar location, thus accounting for differences in the astrometric analysis. The independent data reductions and analyses of these ACS observations performed by Currie et al. (2012) and Galicher et al. (2013) give measured positions within the mutual 2σ error bars.

Table 3. Star–Planet Astrometry

UT Start/End Midpoint West Offset North Offset Detector
(JD) (mas) (mas)
2004 Oct 25/26 2453304.2510995 8587 ± 24  9175 ± 17 ACS/HRC
2006 Jul 17/20 2453935.3606890 8597 ± 22  9365 ± 19 ACS/HRC
2010 Sep 13 2455452.9415740 8828 ± 42  9822 ± 44 STIS
2012 May 29/31 2456078.1699655 8915 ± 35 10016 ± 37 STIS

Download table as:  ASCIITypeset image

2.2. ACS/HRC Photometry

Kalas et al. (2008) reported that Fomalhaut b dimmed by up to 0.8 mag between the 2004 and 2006 epochs in the F606W bandpass. At the time, this result was validated by (1) checking that the 2004 Fomalhaut PSF halo beyond the edge of the occulting spot subtracts the 2006 Fomalhaut PSF halo without any additional image scaling and (2) performing aperture photometry on a background star common to both epochs of data. The former test showed no more than 2% change in calibration (with 2004 as the dimmer of the two epochs), and the latter test showed that the 2004 image of the background star was no more than 0.1 mag dimmer than the 2006 data. The key difference between the control star and Fomalhaut b is that the background star has higher signal-to-noise ratio because it is brighter than Fomalhaut b by ∼1.0 mag, and it is located farther away from the residual speckle halo. The 2006 location of the background control star was 5farcs6 west and 13farcs1 south of Fomalhaut, outside the boundary of the dust belt.

These results indicated that systematic calibration uncertainties would make Fomalhaut b appear dimmer in 2004, when in fact Kalas et al. (2008) reported that it was brighter in 2004. Kalas et al. (2008) quoted the standard error as 0.05–0.10 mag for the Fomalhaut b photometry, which translates to a standard deviation σ ≈ 0.2 mag. Therefore the photometric variability measured for Fomalhaut b was interpreted as significant.

Independent analyses of the same observations by Currie et al. (2012) and Galicher et al. (2013) do not confirm astrophysical variability, but the differences between these two follow-up studies suggest that the photometric uncertainties are larger and of order the variability given by Kalas et al. (2008). For the F606W apparent magnitude, Currie et al. (2012) give 24.97 ± 0.09 mag and 24.92  ±  0.10 mag for 2006 and 2004, respectively. These values are consistent with the 2006 measurement of 25.1 ± 0.2 in Kalas et al. (2008). However, the photometry presented by Galicher et al. (2013) is consistent with the 2004 photometry given by Kalas et al. (2008). Since the observations are the same, these results suggest that there are systematic photometric uncertainties due to the choices of data reduction and analysis methods for high contrast imaging. In an experiment discussed below, we show that the uncertainty in STIS photometry may be >35% due to the residual speckle noise.

3. HST/STIS OBSERVATIONS AND DATA REDUCTION

We observed Fomalhaut in 2010 and 2012 using the coronagraphic mode of STIS (Woodgate et al. 1998) on board HST. STIS includes a 1024 × 1024 pixel CCD with two orthogonal occulting wedges and a 3farcs0 wide occulting bar located in the focal plane. STIS does not have a Lyot pupil plane stop and therefore the diffraction spikes due to the secondary support spider are evident. Also, STIS imaging is conducted without filters. The effective sensitivity of the CCD covers the full range between 0.20 μm and 1.03 μm. The 0farcs05077 pixel scale results in a 52'' × 52'' field of view. We set the gain to 4.015 electrons DN−1.

Table 1 summarizes the STIS observations of Fomalhaut used in the present paper. Calibration of the data, such as bias subtraction and flatfielding, are executed by the OPUS pipeline. We manually processed the image frames (with extension flt.fits) further by identifying cosmic rays or chip defects and replacing these pixels with values derived from an interpolation over neighboring pixels. For sky subtraction in each exposure, we take the median value of a 400 pixel2 region in the upper left corner of the CCD (furthest away from the star). The images are then divided by their exposure times. The geometric distortion correction is performed manually using the currently recommended calibration file in the archive (o8g1508do.idc.fits).

The next steps involve subtracting the central PSF using roll deconvolution. Each orbit has a fixed telescope position angle orientation that differs by a few degrees from neighboring orbits (Table 1). The first step is to coregister all of the orbits in x, y translation by adopting a single fiducial orbit and subtracting the PSF's of all the other orbits, iteratively adjusting the x and y offsets until the PSF subtraction residuals are minimized. The registered frames from all of the orbits at different roll angles are then median combined. Astrophysical features rotate in the CCD frame whereas the stellar PSF structure is fixed. The median value of each pixel effectively removes astrophysical features and produces a clean image of the stellar PSF. This approximation of the Fomalhaut PSF is then subtracted from each of the frames taken at different position angles. These PSF-subtracted frames are then rotated so that astrophysical features are registered, and these frames are then median-combined, recovering a PSF-subtracted astrophysical scene.

In the 2010 September data set we discovered time variable distortions that contribute to our astrometric errors for Fomalhaut b. In the first two orbits, two field stars are visible at the top and at the bottom of the full field (36farcs5 and 27farcs5 radius from Fomalhaut, respectively). Splitting each orbit into quarters, we find the top field star drifts upward (+Y) by 0.2 pixels (10 mas), whereas the bottom field star remains stationary (1σ = 0.03 pixels). In the horizontal direction (X), both field stars remained stationary (1σ = 0.03 pixels) throughout the orbit. This effect is observed for both orbits. The third and fourth orbits have these field stars outside the field of view due to the changes in telescope roll orientation.

One possible source of position drift within an orbit is the fact that all of our observations utilized a single guide star. However, we rule out this effect because the bottom field star does not appear to drift, nor is there a systematic offset in the PSF of Fomalhaut that corresponds in direction and magnitude to the offset of the top field star.

To estimate astrometric error due to uncorrected geometric distortion, we report recent private communication with STIS instrument scientists (J. Duval and A. Aloisi 2012) concerning ongoing astrometric calibration measurements. The calibration program consists of observing Omega Cen once a year with STIS and WFC3. The STIS observing sequence consists of four 10 s exposures in a four-point dither pattern obtained within a single HST orbit. After pipeline processing, the STIS field is registered to the WFC3 field and the R.A. and decl. offsets between the STIS and WFC3 centroids are computed. Only exposures with >10 stars in common between the STIS and WFC3 fields are used. In 2011 the mean and standard deviations of these offsets in STIS pixel units are (1.84, 0.82, 0.96, 1.46, 0.92) ± (2.13, 1.43, 1.83, 1.74, 1.94), respectively. In 2012 there are five different exposures giving (0.59, 0.52, 0.57, 1.00, 0.97) ± (0.33, 0.31, 0.30, 1.45, 1.49).

These data suggest that the uncorrected geometric distortion is variable not only from year to year, but also within a single orbit. The 0.2 pixel drift that is detected in the Fomalhaut data is therefore a general characteristic of STIS imaging observations and not necessarily a consequence of our specific guide star uncertainty. The time dependence suggests effects such as thermal breathing are important. An astrometric calibration program for Fomalhaut would therefore require observations of fields near Fomalhaut and interspersed in time with the Fomalhaut observations.

Since an astrometric calibration program designed specifically for Fomalhaut is not practical, one option is to adopt the rms value of the astrometric uncertainties given by the entire STIS astrometric calibration from 2001 to the present epoch. This value is 1.303 pixels, or 66 mas (Table 2). However, as discussed below and in Section 8.1, this overestimates the uncorrected geometric distortion uncertainty.

As a second method to estimate the error in geometric distortion correction, we measure the positions of two field stars in the 2012 data before and after the geometric distortion is applied to the data. The largest difference measured between the before and after positions is 0.33 pixels, or 17 mas. Thus we adopt 17 mas as an estimate of geometric distortion "measured in data" (Table 2).

The single guide star guiding also introduces the possibility that the position angle orientation of the telescope contains significant systematic error. To quantify the position angle uncertainty, we found two field stars that are contained in the STIS field, as well as previous observations of Fomalhaut with WFC3/IR and WFPC2. All three observations were made with a single guide star. The two field stars are separated from each other by 47'' and are located east and north of Fomalhaut. The measurement of position angle in these three data sets gives an empirical uncertainty of 0fdg06 (1σ) in the telescope orientation.

Finally, as with the ACS/HRC data, we used several methods for determining the uncertainty in the position of the star behind the occulting wedge. Note that since all of the orbits were registered to a single fiducial orbit (see above), the stellar center is determined for the single, fiducial orbit. Unlike the HRC, the STIS images have prominent diffraction spikes that can be used to estimate the stellar position. However, since the STIS occulting wedges are not partially transparent, the core of the PSF is inaccessible using short exposures. Perhaps the greatest advantage with STIS is the larger field of view that will contain more field stars (however, the occulting wedges often block the field stars at various telescope roll angles). Since the data are obtained at different roll angles about Fomalhaut, it is possible to verify if the rotation center is accurate by studying how well the field stars are coregistered after derotation to a common reference frame.

Our first estimate for the stellar position is derived from the diffraction spikes. We fit two straight lines to the spikes, yielding an rms residual of 0.15 pixels (8 mas), with no systematic curvature. This result can be validated by rotating the image 180° to demonstrate that the diffraction spikes self-subtract. Using this center position, the frames are rotated so that north is up and east is left. The centroids of the field stars can be determined for each north-rotated orbit, and the procedure is repeated again using 0.5 pixel and 1.0 pixel deviations from the initial center position determined from the diffraction spikes. If the field-star positions in the north-rotated frames differ from each other in separate roll angle observations by more than 0.2 pixels, we consider the assumed rotation center position as invalid. In this way the center position is tested and refined so that the field-star centroids are statistically identical from orbit to orbit in the registered, de-rotated frames. A weakness of this technique is that there are only two field stars for reference, and some frames contain one star and not the other because the STIS occulting wedges block different portions of the field at different roll angles. The center position determined using the field-star-rotation technique differs from the diffraction spike center position by 0.5 pixels in x and 0.5 pixels in y. However, the diffraction spike center does not violate the 0.2 pixel cut-off defined above. Therefore, adopting a conservative, worst-case scenario, we establish 0.5 pixels (25 mas) as the uncertainty in determining the stellar center in the STIS coronagraphic data ("STIS centroiding star" in Table 2).

4. 2010 RECOVERY OF FOMALHAUT b

Figure 2 presents the final, unsmoothed image resulting from the combination of the four orbits executed in 2010 September. The southeast side of the belt lies outside the STIS field of view. The northwest portion of the belt is detected for the first time in scattered light since this region was outside the field of view in the ACS/HRC data. As reported in Kalas et al. (2010) and Kalas (2011), the northwest side of the belt reveals an extended halo of nebulosity, indicating that the belt is broader than previously reported from the ACS/HRC observations.

Figure 2.

Figure 2. False-color, linear-scale image of a portion of the Fomalhaut field containing the 2010 detection of the belt and Fomalhaut b to the northwest of the star. North is up, east is left. Also shown is the discovery of a halo of nebulosity northwest of the main belt. The box inset is 1'' on a side and magnifies the image of Fomalhaut b. A smoothed version appears in Figure 12.

Standard image High-resolution image

The source we identify as Fomalhaut b is the brightest object in a 1'' × 1'' search box at the expected location of Fomalhaut b at the epoch of observation (Figure 2, inset). The centroid of Fomalhaut b is offset from the central star by ΔR.A. = −8.828 ± 0farcs042, Δdecl. = +9.822 ± 0farcs044 (J2000.0), with projected stellocentric separation ρ = 13farcs206 (Table 3). This 2010 position is consistent with the independent detection of Fomalhaut b in these data by Galicher et al. (2013), who measure ΔR.A. = −8.81 ± 0farcs07, Δdecl. = +9.79 ± 0farcs07.

The apparently extended morphology of Fomalhaut b along the vertical axis will be discussed further when we introduce the 2012 detection. The signal-to-noise of the STIS detection is decreased relative to the ACS/HRC observations by significant residual speckle noise and subtraction artifacts. Compared to the ACS/HRC data, STIS has a coarser pixel scale (51 mas versus 25 mas for the HRC) that results in poorer sampling of the PSF halo and speckles. The significantly broader bandpass of the STIS data also increases the radial extent of speckles relative to the ACS/HRC observations.

We quantify the signal-to-noise of the Fomalhaut b detection using aperture photometry. We placed a circular aperture of 2.4 pixel diameter at the location of the source. We subtracted an average background level determined from an annulus 3–5 times the target aperture size. After correcting for the encircled energy fraction,6 we measure an overall source flux of 0.491 counts s−1. We adopt the relatively small aperture radius because the residual speckles that may produce false positive detections of point sources have this characteristic size. Therefore our estimate of signal-to-noise using this method is an attempt to quantify the speckle noise in the reduced images. To assess the noise, we sampled photometry using the same aperture in an arc-like region at the same radial distance as Fomalhaut b. We use this shape because speckle noise decreases with increasing radial distance form the star, and it is the dominant source of noise at the location of Fomalhaut b. To avoid the regions dominated by diffraction spikes and Fomalhaut's dust belt, the arc resides entirely within the inner boundary of the dust belt. We choose pixels with radii within 2.5 pixels of Fomalhaut b's projected radius and are at least 10 pixels away from Fomalhaut b. The distribution of photometric sample values in this residual speckle dominated region are shown in the top panel of Figure 3. The distribution is Gaussian-like with a sample standard deviation of 0.138 counts s−1. This metric suggests that Fomalhaut b is a 3.5σ detection in the 2010 data.

Figure 3.

Figure 3. Histograms for each epoch show the distribution of photometry collected in 2.4 pixel diameter apertures spread across an arc at the same radius as Fomalhaut b. The azimuths of the apertures were selected to be interior to the dust belt. Values were aperture corrected and background subtracted. Overplotted in dotted lines are the Gaussian distribution with mean and standard deviation taken from the samples. The dashed vertical line shows the resulting photometry at the location of Fomalhaut b.

Standard image High-resolution image

The surprising result from the 2010 detection is that Fomalhaut b is detected three pixels (150 mas) westward from the position that would be expected for a low-eccentricity orbit (e ∼ 0.1) nested within the inner boundary of the debris belt. Kalas et al. (2010) and Kalas (2011) reported that the uncertainties in the roll angle of the telescope and the uncorrected geometric distortion may plausibly account for the three pixel deviation. Subsequent work reported in Section 3 quantifies both the position angle and uncorrected geometric distortion uncertainties (Table 2). Adopting a geometric distortion uncertainty of 66 mas based on the unpublished STIS calibration program leads to an astrometric uncertainty (1σ) of 76 mas in right ascension. Therefore the 150 mas westward deviation observed in 2010 could be considered a 2σ result. We concluded that a fourth epoch of observation was required in order to test the significance of the westward deviation. As discussed in subsequent sections, after the 2012 epoch confirmed a highly eccentric orbit, the error analysis given all four epochs of astrometry justifies the adoption of a smaller value for the uncorrected geometric distortion (17 mas instead of 66 mas; Table 2).

5. 2012 CONFIRMATION OF FOMALHAUT b

The main difference between the 2012 STIS data and the 2010 STIS data is a factor of three increase in integration time and telescope roll angles. Figure 4 shows the 2012 confirmation of Fomalhaut b. We employ the aperture photometry measurements of the previous section and measure Fomalhaut b to have a flux density of 0.498 counts s−1. The standard deviation of the comparison measurements in the residual speckle arc-like region is 0.109 counts s−1, resulting in a 4.6σ detection (lower panel of Figure 3). Noise due to quasi-static speckles should be reduced by increasing the number of realizations of speckles. Figure 5 shows a horizontal line cut through the image that intersects Fomalhaut b. Fomalhaut b is a prominent feature compared to the local noise and the western portion of the dust belt.

Figure 4.

Figure 4. False-color, unsmoothed image of the 2012 STIS data. A smoothed version appears in Figure 12. The box inset is 1'' on a side and magnifies the image of Fomalhaut b. North is up, east is left.

Standard image High-resolution image
Figure 5.

Figure 5. Horizontal cut across the image in Figure 4. The cut represents the average of three lines centered on Fomalhaut b, where the center line is 10farcs00 above the stellar position. The x-axis plots the position relative to the star. This horizontal cut passes through a range of radii from the star. Therefore the residual speckle noise is prominent in the range −5farcs0 to 5farcs0. However, Fomalhaut b is a prominent feature relative to the local noise in the region ∼13'' radius from the star.

Standard image High-resolution image

Fomalhaut b appears to have an elliptical morphology with major axis at P.A. ≈ 137°. The major axis has FWHM of 5.2 pixels (264 mas) and full-width quarter-maximum (FWQM) of 7.4 pixels (376 mas). The minor axis has FWHM and FWQM 2.5 and 3.5 pixels, respectively. The background field star south of Fomalhaut b in the lower quarter of the frame has FWHM of 2.4 and 2.1 pixels in the x (R.A.) and y (decl.) directions (FWQM 3.6 pixels and 2.9 pixels, respectively). Therefore Fomalhaut b is consistent with a point source along its minor axis direction, but appears significantly extended along its major axis. If the extended morphology is astrophysical, then the 376 mas FWQM of the major-axis corresponds to 2.9 AU.

However, the radial direction relative to the star at the position of Fomalhaut b is P.A. = 138°, invoking the possibility that the extended morphology is due to the residual speckle noise. To assess this possibility, we conduct an experiment using an STIS artificial stellar PSF generated by TinyTim (Krist et al. 2011). First, we determine if the data reduction steps distort the morphology of the artificial PSF. The PSF is inserted into blank fields at positions corresponding to the various roll angles of the data. When these experimental data are processed in a manner identical to the real data, and measurements conducted in an identical manner, the resulting TinyTim PSF has FWHM of 2.0 and 1.8 pixels (FWQM of 3.4 and 3.2 pixels) along the x and y directions, respectively. The field star is ∼15% broader than this, most likely because of noise (demonstrated below), the uncertainty in determining the rotation center in the real data, and the fact that a real PSF core may be land over several pixels instead of a single pixel.

The second step in the experiment is to insert the artificial PSF described above into the 2012 data to quantify the magnitude of PSF distortion due to residual speckle noise. We insert nine copies of the artificial PSF at locations in the image where positive noise features of at least four contiguous pixels are apparent, and nine more locations where the noise is negative. Figure 6 demonstrates that point sources may appear extended due to positive noise. For example, source number nine has FWHM and FWQM of 3.9 and 6.8 pixels, respectively. We plot the respective FWHM and FWQM measurements for the nine positive noise sources and Fomalhaut b in Figure 7. Fomalhaut b is the most extended source in both cases, but it does not appear to be an outlier when measuring the FWQM. The minor axis measurements shown in Figure 7 for the artificial implants demonstrate that in most cases the minor axis of a point source is also broadened due to noise.

Figure 6.

Figure 6. Experiment with implanted point sources in the 2012 data. Circles mark locations where an artificial point source is inserted at the position of positive noise features in the image. Squares denote locations where the artificial source is inserted at negative noise features. The oval marks Fomalhaut b.

Standard image High-resolution image
Figure 7.

Figure 7. Measurements of the implant sources and Fomalhaut b FWHM (left) and FWQM (right) along the minor and major axes. In both plots Fomalhaut b is the topmost data point. Fomalhaut b appears to be an outlier in the measurement of the FWHM of the major axis. However, Fomalhaut b is indistinguishable in the graph of FWQM.

Standard image High-resolution image

We conclude that since residual noise is a plausible explanation for Fomalhaut b's extended morphology, other data sets are required to establish whether or not Fomalhaut b is extended. In the 2010 STIS data, Fomalhaut b appears somewhat extended. The FWHMs of the minor and major axis are 2.1 × 7.2 pixels (107 × 366 mas). The corresponding FWQMs are 3.5 × 8.9 pixels (178 × 452 mas). However, the orientation of the major axis is north–south in 2010. Thus, if both the 2010 and 2012 STIS observations detect extended structure, rotation by ∼45° in 18 months must be explained. The deepest, best-sampled images of Fomalhaut b are the 2006 observations obtained with the ACS/HRC and the F606W filter. In these data, Fomalhaut b appears to be a point source with FWHM = 69 ± 6 mas (Kalas et al. 2008). Therefore if the 2012 extended morphology is real, it would require that Fomalhaut b is spreading over time, or has a triaxial shape that occasionally appears point like as the major axis rotates into our line of sight, minimizing the projected size. Future observations with HST/STIS or other instruments can refute or confirm the extended, time-dependent morphology.

The implant experiment also quantifies the astrometric error for centroiding on Fomalhaut b. The median difference between the known implant locations and the recovered centroid positions of the implants is 0.1 pixels in X and 0.2 pixels in Y. These values are used in quantifying the astrometric error in Table 2.

6. STIS PHOTOMETRY OF FOMALHAUT b

As noted above, background and aperture corrected photometric measurements give 0.491 counts s−1 and 0.498 counts s−1 for the respective epochs of STIS observation. We use the IRAF/STSDAS package countrate to obtain a STIS zeropoint (i.e., 1 counts s−1) of mv = 24.48 mag in the VEGAMAG system for an input spectrum comprising a Teff = 8500 K, log (g) = 4.0 Kurucz model spectrum. In other words, this particular experiment with the photometry assumes that Fomalhaut b has the same spectrum as the host star. This gives a factor of 4.628 × 10−4 for converting counts to mJy. The sky and aperture corrected photometry on Fomalhaut b therefore corresponds to 0.227 μJy and 0.230 μJy for 2010 and 2012, respectively.

One measure of the photometric uncertainty is to adopt the speckle noise of 0.138 counts s−1 and 0.109 counts s−1 (or 0.064 μJy and 0.050 μJy, respectively) as the 1σ photometric uncertainties. A second measure derives from the experiment with artificial point sources inserted into the data. The implants on negative noise regions result in very weak detections or non-detections. This is because the artificial point source flux was scaled so that the peak pixels, when implanted on positive noise features, result in pixel values close to that of Fomalhaut b in the image (and thereby giving a proper comparison of FWHM and FWQM). For the nine implants on positive noise features, the standard deviation in photometry gives 25% uncertainty in the photometry. If we add in seven more implant locations on the negative region (excluding two implants that give none-detections), the standard deviation in photometry is 35%. Given these significant uncertainties, the 2010 and 2012 photometry on Fomalhaut b are consistent with each other. Kalas et al. (2008) report Fomalhaut b photometry of 0.75 and 0.36 μJy in the ACS/HRC F606W filter in the 2004 and 2006 observations, respectively. Using countrate with the same input spectrum as above, we convert the ACS F606W fluxes to STIS values, giving 0.63 and 0.30 μJy for the 2004 and 2006 photometric points, respectively. The two epochs of STIS data are therefore in better agreement with the 2006 ACS/HRC data. These results would suggest that the 2004 ACS/HRC photometry may be anomalous. We have found no source of error in our 2004 photometric analysis other than residual speckle noise. On the other hand, for the 2010 STIS data, Galicher et al. (2013) independently report 0.61 ± 0.21 μJy in the ABMag system using PSF fitting based on TinyTim models. These discrepancies further reinforce our conclusion at the end of Section 2.2 that different data reduction and photometry techniques lead to systematic differences in the absolute measured flux. What is important in the context of the present manuscript is that we do not measure a change in photometry between 2010 and 2012 given the same instrument and uniform methods.

We note that in the scenario where the optical flux of Fomalhaut b originates from light scattering by dust grains in a cloud orbiting a planet (Kalas et al. 2008), Fomalhaut b must dim over time, everything else being equal, because the stellocentric distance is currently increasing. In 2006, at a stellocentric distance of 119 AU (assuming a co-planar geometry with the belt), the incident flux on Fomalhaut b was 1.58 W m2. At the 2012 epoch the distance has increased to 125.4 AU, and the incident flux decreased to 1.42 W m2. In this scenario we expect an 0.12 mag dimming from 2006 to 2012. This effect is not apparent in the nominal photometry derived above, nor would the effect be detectable to sufficiently high confidence because the photometric noise described above is of the order of a few tenths of a magnitude. Moreover, the conversion of instrumentation from ACS to STIS involves observations in different bandpasses, requiring additional assumptions that need to be made regarding the optical spectrum of Fomalhaut b. If we instead consider only the STIS observations, the 2010 position of Fomalhaut b is 124.1 AU from the star and the incident flux is 1.46 W m2. Fomalhaut b would be dimmer in 2012 by only ∼0.02 mag, which is undetectable given the current photometric uncertainties. The nominal STIS photometry above shows an 0.02 mag brightening between 2010 and 2012, which if it were astrophysical cannot be attributed to the motion of Fomalhaut b away from the star in the coplanar case. Continued photometric measurements of Fomalhaut b with STIS in future epochs will certainly build a time series of photometry that would explore whether or not Fomalhaut b's apparent magnitude decreases over time.

7. MAIN BELT PROPERTIES FROM THE STIS OBSERVATIONS

Figure 8 is a mosaic of the 2010 and 2012 STIS images in the reference frame of the star. The main difference between this combined STIS image and the 2004/2006 ACS/HRC observations is that the northwestern side of the belt is now fully contained within the field of view, revealing an extended halo of nebulosity to the northwest. Neither the STIS nor the ACS/HRC observations include the field significantly beyond the southwest side of the belt so as to ascertain if the extended halo is a symmetric feature to either side of the belt, which we refer to as the "main belt" to distinguish it from other belts in the system (Section 9.1).

Figure 8.

Figure 8. Mosaic of the 2010 and 2012 STIS data registered to the location of Fomalhaut A. In the central regions of overlap (see Figures 2 and 4) this combined figure represents the average value from the two epochs of observation, which means that background objects and Fomalhaut b are blurred due to their motion between epochs. North is up, east is left. The circle and diamond mark the stellar center and the geometric center of the belt, respectively. The data are not smoothed.

Standard image High-resolution image

7.1. Main Belt Geometry

Figure 9 shows a cut along the apparent major axis of the belt. In the majority of cuts, there is a double peak of maximum brightness (see also Figure 3 in Kalas et al. 2005). To avoid confusion with the azimuthal "gap" discussed below, we characterize this specific radial morphology as a "bisected" plateau. The apparent major axis measured between the inner peaks of the belt has length 277.34 AU. The center of the belt bisector is 1.85 AU and 1.62 AU (SE and NW sides, respectively) further outward from the inner peak. The outer peak is another 1.54 and 1.85 AU (SE and NW sides, respectively) outward from the gap. Therefore the distance between the two peaks is approximately 3.4–3.5 AU. The major axis is 280.81 AU if we measure the distance between the belt bisectors, which is consistent with the semi-major axis value (140.7 ± 1.8 AU) derived by Kalas et al. (2005).

Figure 9.

Figure 9. Cut along the major axis of the belt shown in Figure 8, which is the average pixel value in a 1 arcsec wide segment along the minor axis direction. The inset to the left magnifies the characteristic double peak structure of the belt, where the peaks are bisected and separated by ∼3.4 AU.

Standard image High-resolution image

To measure the projected (sky plane) shape of the belt, we determine the stellocentric positions of two distinct features in radial cuts through the belt: (1) the bisector and (2) the inner edge of the belt, defined as the half maximum of the line that rises to the inner peak. These position measurements from radial cuts are not possible in the regions closest to the star that are dominated by speckle noise, or crossing the azimuthal belt gap that is discussed below.

Figure 10 plots the bisector positions and a corresponding least-squares Keplerian fit. The fit assumes that the apparent belt structure traces a simple Keplerian orbit. We consider the orbital phase at which a hypothetical belt particle would pass each measured point and solve jointly for the orbital elements and these orbital phases. These phases are "nuisance parameters" in the problem, and the posterior distributions are marginalized over these parameters. Table 4 summarizes the belt properties from Kalas et al. (2005) and these two new measurements and Keplerian fits for the STIS data. Even though we use a Keplerian orbit for the apparent structure, the belt is presumed to be a circle inclined to the line of sight. If the belt represents a non-circular Keplerian orbit, the eccentricity is the ratio of the stellocentric offset of the belt center to the semi-major axis. Therefore the projected ellipticity of the belt is due to both the inclination and the inherent non-circular morphology. Table 5 gives revised values for the belt's properties assuming a non-circular structure inclined to the line of sight. We note that Kalas et al. (2005) derived these values from the Kowalsky construction using apparent ellipses to find the true orbital elements (from Smart, pp. 352–353). Here we have revised the Kalas et al. (2005) values using the Keplerian orbit approach. The new values derived from the STIS data are in good agreement with those published by Kalas et al. (2005).

Figure 10.

Figure 10. Left panel: measurements of the peak scattered light from the main belt (red crosses) and a fit (blue crosses) based on a Keplerian orbit. Right panel: residuals between the Keplerian fit and the measured points. θ is the best-fit value of ν + ω (true anomaly + argument of perapse).

Standard image High-resolution image

Table 4. Apparent (Sky–Plane) Belt Geometry

Data a P.A. Semi-minor $\sqrt{1-b^2/a^2}$ Inclination R.A. Offset Decl. Offset
(AU)
Kalas et al. (2005) 140.7 ± 1.8 AU 156fdg0 ± 0fdg3 57.5 ± 0.7 AU 0.91 ± 0.01 65fdg9 ± 0fdg4 −2.2 ± 0.3 AU 13.2 ± 0.9 AU
STIS (bisector) 141.3 ± 0.4 AU 156fdg2 ± 0fdg1 57.5 ± 0.2 AU 0.91 ± 0.01 66fdg0 ± 0fdg2 −2.66 ± 0.15 AU 13.06 ± 0.25 AU
STIS (inner edge) 135.0 ± 0.4 AU 156fdg4 ± 0fdg1 52.6 ± 0.2 AU 0.92 ± 0.01 67fdg0 ± 0fdg2 −1.13 ± 0.15 AU 12.75 ± 0.25 AU

Download table as:  ASCIITypeset image

Table 5. Derived Main Belt and Fomalhaut b Keplerian Orbital Elementsa

Data a e i Ω ω
(AU)
STIS (bisector) 141.77 ± 0.28 0.10 ± 0.01 −66fdg1 ± 0fdg1 156fdg1 ± 0fdg1 29fdg6 ± 1fdg3
STIS (inner edge) 136.28 ± 0.28 0.12 ± 0.01 −67fdg5 ± 0fdg1 156fdg2 ± 0fdg1 41fdg9 ± 1fdg1
Fomalhaut b  177 ± 68 0.8 ± 0.1  −55° ± 14°  152° ± 13°  26° ± 25°

Note. aAssumes stellar mass = 1.92 M

Download table as:  ASCIITypeset image

7.2. Main Belt Outer Halo

The slope of the belt halo brightness along the apparent semi-major axis between 20'' and 27'' projected radius (Figure 9) is fit by a power law with index −3.3. The right portion of Figure 9 shows that halo is detected more than 6'' beyond the inner edge of the belt. To improve the signal-to-noise of low surface brightness nebulosity along the apparent major axis, we bin the data along the apparent minor axis direction. Figure 11 indicates that the extended halo extends at least as far as 209 AU from the star, or 57 AU beyond the inner belt edge (Figure 9). At 209 AU the isophotes bend westward by ∼35° (relative to the major axis, when the image is restored from its collapsed state). Other debris disk midplanes show evidence for bending, such as HD 32297, which has a ∼31° bend (Kalas 2005), and HD 61005, which has a ∼10° bend (Maness et al. 2009). Another possible explanation for the apparent bending of the Fomalhaut belt halo in the STIS data is that the lowest surface brightness contours are influenced by a small mismatch (<0.005 counts s−1) in the background sky levels between frames in the 2012 epoch, and in between epochs. These mismatches are emphasized when the images are binned and smoothed. Future observations are required to confirm whether or not the Fomalhaut dust belt is detected beyond 209 AU and with a bend in the position angle. In any case, the main finding is that the belt is significantly more extended in scattered light than previously known, with a detection out to at least 209 AU radius from the star.

Figure 11.

Figure 11. Binned image showing the northwest side of Fomalhaut's belt. The combined image is rotated such that the major axis of the belt lies along a horizontal, with the northwest side pointing right. The data are then binned 2 pixels along the x-axis and 20 pixels along the y-axis, and then convolved with a Gaussian with σ = 2 pixels. The left edge of the frame represents the stellar position, and the distances mark the stellocentric positions.

Standard image High-resolution image

7.3. Main Belt 331° Azimuthal Gap

An additional newly discovered belt feature is an azimuthal belt gap approximately 6'' north of Fomalhaut b. Figure 12 emphasizes this gap by smoothing the data and displaying the images with a hard stretch. We classify it as a real astrophysical feature (as opposed to an instrumental artifact) because it is apparent in both epochs of STIS data. In the projected (sky plane) view, the gap appears to be ∼3° wide, beginning at P.A. = 329fdg8. Since this is the faintest portion of the belt, we examined the possibility that the gap is an artifact introduced by instrumentation or data reduction. For the 2012 observations we exclude from data reduction 5 of the 12 orbits where this gap region lands near a diffraction spike, occulting mask, or field edge. We also studied the subtraction pairs sufficiently separated in roll angle such that self-subtraction of the belt would not be possible. The final images produced with the remaining seven orbits continues to show the belt gap. We therefore conclude that since artifacts are excluded and the gap appears in both STIS data sets (this region is not in the field of view for the ACS/HRC observations), it is likely a real astrophysical gap.

Figure 12.

Figure 12. Northwest portions of the 2012 data (left) and 2010 data (right) with 5 × 5 median binning and a hard grayscale stretch to emphasize the northwest gap. The dotted lines are at P.A. = 329fdg8 and 332fdg8 in both data sets. Ovals indicate background galaxies and Fomalhaut b is marked between white line segments.

Standard image High-resolution image

Figure 13 shows a deprojection of the belt and a radius squared multiplicative scaling of the image centered on the star to normalize for the fall-off in stellar illumination. The deprojections assumes a circular structure inclined to the line of sight by 66fdg5 (Table 4). In 2012, Fomalhaut b is 98 AU and 78 AU to the right and below the star, respectively, in this reference frame. The four-epoch motion is essentially to the right in the +X direction at roughly 0.00260 AU day−1. Assuming Fomalhaut b is coplanar with the belt, it has to travel ∼19 AU to reach the inner edge of the belt. Therefore we might expect to witness the real or projected belt crossing around 2032.

Figure 13.

Figure 13. Deprojection of the belt (Figure 8) by 66fdg5 after the apparent semi-major axis is rotated to horizontal (clockwise by 66fdg0). The right panel has been normalized by multiplying the image by radius squared centered on the stellar location. The angle markings on the right panel are belt centric, so that 0° and 180° mark the belt's apparent minor axis.

Standard image High-resolution image

The azimuthal brightness asymmetries are due to an asymmetric scattering phase function and the fact that the star is closest to the southern portion of the belt, as discussed in Kalas et al. (2005). Since the star is 13 AU to the left of the belt center, one effect is that the left hemisphere of the belt receives greater illumination than the right hemisphere, which may account for apparent belt gaps in the right hemisphere. However, the belt gap is still evident in the illumination-corrected image. Figure 14 gives photometric measurements along the circumference of the belt in the illumination-corrected image. The brightness in the gap minimum is approximately 50% of the mirror region in the left hemisphere. The gap width measured as a FWHM is ∼50 AU. We note that the belt minor axis serves as the reference frame for the azimuth (degree) measurements shown in the right panel of Figure 13, which is slightly offset from the reference frame describing the scattering angles relative to the star. This offset between reference frames is ∼5° within 30° of the minor axis (e.g., between azimuth 150°–210°). However, at the azimuth of the belt gap (∼250°) the scattering angle offset is <1° and therefore the gap cannot be explained by a scattering phase function effect.

Figure 14.

Figure 14. Photometry along the illumination-corrected, deprojected belt (Figure 13) in circular apertures with diameter 0farcs5. The blue line is a ninth order polynomial fit to the gap region. The green dashed line is a least-square linear fit to the points excluding the gap, but includes measurements in the 100°–140° region. The green dashed line fit is near horizontal, as expected for the left-right symmetry argument. The gap is a significant depression, but not entirely empty. The full width at half minimum is ∼50 AU (note that at 141 AU radius, 1° on the plot corresponds to 2.461 AU in circumference).

Standard image High-resolution image

8. ORBIT OF FOMALHAUT b

8.1. Astrometry and Uncertainties

Table 3 summarizes the four epochs of astrometry with 1σ error bars derived from combining the error terms in Table 2 in quadrature. Determining the position of the star behind occulting spots or wedges and the residual geometric distortion in STIS are the two most significant sources of astrometric uncertainty. To test for possible systematic errors in any of the epochs, we conducted astrometry on a faint background star south of Fomalhaut, residing outside of the dust belt boundary. This is the only background object detected at all four epochs. Figure 15 compares our astrometry to the predicted locations using proper motion and parallax information from the Hipparcos catalog. The residuals between the expected and measured locations are ∼20 mas, which we take as evidence that the 66 mas value for the residual geometric distortion adopted from the STIS calibration program is an overestimate. In Figure 15 the error bars plotted are derived from the residual geometric distortion for STIS inferred from the Fomalhaut data (17 mas; Table 2). The residuals are now comparable to the 1σ error bars, justifying adoption of the 17 mas value for the assumed uncorrected geometric distortion in STIS. We note this is a likely upper limit given that the background star is ∼1'' (20 pixels) farther from the star than Fomalhaut b in 2012.

Figure 15.

Figure 15. Test of astrometry on a background star in the reference frame of Fomalhaut A. For the 2010 and 2012 epochs of STIS data, the error bars are derived assuming 17 mas in residual geometric distortion instead of the 66 mas from the STIS calibration program (Table 2).

Standard image High-resolution image

8.2. Kinematics

Figure 16 shows the sky–plane motion of Fomalhaut b compared with a uniform motion (unaccelerated) model. The corresponding best-fit speed is v = 4.36 ± 0.17 km s−1, with corresponding χ2 = 4.73 (4 degrees of freedom; the cut-off probability P2 > 4.73) = 0.32). The quoted velocity uncertainty is a propagation of the astrometric error (Table 3). We do not know the escape speed if we assume that projection effects are unknown. However, for M = 1.92 M (Mamajek 2012), the escape speed $v^p_{{\rm esc}} = \sqrt{2GM_\star /r_p}$ = 5.837 km s−1 at the mean observed projected separation (rp ≈ 100 AU), represents an upper limit. For circular orbits $v/v^p_{{\rm esc}} \le 2^{-1/2} \approx 0.707$; hence, the measured value of the ratio $v/v^p_{{\rm esc}}=0.747\pm 0.03$ suggests that we should investigate non-circular orbits.

Figure 16.

Figure 16. Uniform motion model—the dotted line shows motion at constant velocity 4.36 km s−1; the blue diamonds show the predicted positions for the best-fit model.

Standard image High-resolution image

The large observed value of $v/v^p_{{\rm esc}}$ does not mean that the object is unbound. For an ensemble of randomly oriented orbits with e distributed between zero and one, $v/v^p_{{\rm esc}} < 1$. The distribution of $v/v^p_{{\rm esc}}$ depends on the details of the eccentricity distribution: for a uniform distribution (0 ⩽ e ⩽ 1) then $\langle v/v^p_{{\rm esc}}\rangle = 0.425\pm 0.178$; for a more physically based thermal distribution (dp/de = 2e; Heggie 1975) the ratio is 0.402 ± 0.190. Figure 17 shows that cumulative distribution of $v/v^p_{{\rm esc}}$ for randomly oriented, elliptical orbits. Fewer than 9.2% (6.2%) of bound orbits have an observed value of greater than that allowed by the observations at 99% (68%) confidence. Therefore the allowed phase space is not large for bound orbits in random orientations, but not improbable.

Figure 17.

Figure 17. Cumulative distribution of $v/v^p_{{\rm esc}}$ for randomly oriented orbits with an eccentricity distribution dp/de = 2e. The vertical dashed line shows the observed value and its 1σ error bounds.

Standard image High-resolution image

Even though acceleration is not yet detected, the magnitude and direction of motion for a bound object constrain the Keplerian orbital elements. We assume that the stellar mass and heliocentric distance are known and that the errors are sufficiently small that marginalization over the associated uncertainties is unnecessary (a calculation that performs this marginalization over the uncertainty in stellar mass is described at the end of this section). We use standard methods to compute the Cartesian coordinates of an orbiting body by solving Keplers equation for the eccentric anomaly and hence the radius and true anomaly (Green 1985). For hyperbolic orbits (e > 1) we solved Kepler's equation using the approach of Gooding & Odell (1988).

There are six unknowns in this problem: two describe the shape and size of the orbit (eccentricity, e and semimajor axis, a), three angles (argument of perihelion, ω, longitude of the ascending node, Ω, and inclination, i) account for the orientation in space relative to a reference direction (north and position angle) and reference plane (sky plane), and one describes the orbital phase (epoch of perihelion). Figure 18 illustrates the astronomical convention where the ascending node is the point where the orbit penetrates into the sky plane away from us; Ω is an angle in the sky plane measured eastward from north to the ascending node; and ω is the angle in the orbital plane between the ascending node and periastron, q (Green 1985). We assume that Fomalhaut b is currently observed behind (into) the sky plane such that the inclination of the orbital plane is a negative value. We note that the main belt is also described by an orbital plane, where the mutual inclination between this and Fomalhaut b's orbital plane is represented by I.

Figure 18.

Figure 18. Diagram marking several orbital elements for an orbit inclined relative to the reference plane which is the sky plane. Ω and ω are not coplanar. We follow the binary star convention where positive Z (not drawn) is into the sky plane (below the sky plane drawn here), such that the ascending node is the point where the orbiting body crosses the reference plane (red circle) toward positive Z (Green 1985). In this particular sketch, the planet lies out of the sky plane (nearest Earth), which means that the descending node (blue circle) follows periastron passage. The periastron vector lies in the plane of the orbit and represents the direction of the true semi-major axis. This does not necessarily correspond to the apparent semi-major axis of an inclined orbit projected onto the sky plane. At the current epoch, Fomalhaut b has passed through periastron, but it has not yet reached the descending node (i.e., it still resides behind the sky plane).

Standard image High-resolution image

Our data comprise eight measurements: two measurements of position at four distinct epochs (2004, 2006, 2010, and 2012). The problem of finding the orbital elements is therefore over-determined and a statistical approach using, for example, the method of least squares or maximum likelihood is necessary to estimate the orbital elements and their uncertainties.

For initial exploration of the problem we used the Levenberg–Marquardt algorithm to find acceptable sets of parameters (Bevington 1969). It is evident from these investigations that the six-dimensional χ2 surface has many local minima. The Levenberg–Marquardt algorithm finds local minima, not the global minimum; moreover, estimates of the parameter uncertainties, which are derived from a Taylor-series expansion of χ2 about a local minimum, are untrustworthy.

We have therefore used a Markov Chain Monte Carlo (MCMC) method to sample the posterior probability distributions for the orbital elements. For examples of application of these methods for the determination of the orbital elements, see Ford (2006) and Chauvin et al. (2012). The method employed here computes the likelihood function—assuming that the measurement errors are normally distributed—and the Metropolis–Hastings algorithm to select new members of the chain from a proposal distribution (Sivia & Skilling 2006). The Metropolis–Hastings algorithm guarantees convergence of the Markov chains to the posterior distribution, but convergence is slow when a high rate of rejection (≪50%) of the proposed values occurs; a common circumstance for problems with a large number of parameters. To speed convergence we included an initial phase during which an adaptive proposal distribution is used at each step. The adaptation is known as simulated tempering (Gregory 2001). We adopt uniform priors for the proposal distributions of the orbital elements and each chain is started with a random value within the prior range. A burn-in period proceeds and convergence and independence of the Markov chains are established using the statistical methods of Raftery & Lewis (1995).

Figure 19 shows the results of this analysis. The adopted priors for the free parameters are a ∈ [80, 800], e ∈ [0.4, 1.0], Ω ∈ [110°, 200°], and i ∈ [0°, 90°]; no priors were imposed on ω or the epoch of perihelion. The limits on semimajor axis and longitude of the ascending node were imposed after extensive exploration of the entire range of these parameters. No viable solutions were found outside of these ranges (apart from the π-periodicity in Ω and ω) and therefore these priors were adopted for convenience to speed the convergence of subsequent Markov chain calculations; the lower limit of e = 0.4 was adopted for the same reason. The prior probabilities for semimajor axis and inclination are uniform in log a and cos i, respectively. In each case the posterior distribution is sharply peaked in contrast to the initial uniform prior and characterized by a standard deviation that is significantly smaller than the prior range.

Figure 19.

Figure 19. Plots showing the distributions and correlations for orbital elements a, e, Ω, ω, and i for Fomalhaut b. The histograms along the diagonal show the marginalized probability distribution. The off-diagonal plots show the correlation between the corresponding parameters—each dot represents a Markov chain element. The mean and standard deviation of each marginal distribution are listed in the accompanying legend.

Standard image High-resolution image

Figure 19 shows that the current observations favor an elliptical orbit (e = 0.8 ± 0.1) with large semimajor axis (a = 177 ± 68 AU); a low-eccentricity orbit (e ≈ 0.1) that is nested within the belt is ruled out. Figure 20 shows a sample of 100 orbits drawn from the Markov chain, representing orbital elements that are consistent with the astrometric data. This figure demonstrates graphically that the projected motion of Fomalhaut b crosses the main belt. However, because of the mutual inclination of the belt and the orbit Fomalhaut b does not necessarily penetrate the belt.

Figure 20.

Figure 20. Sample of 100 orbits drawn from the Markov chains, representing orbits that are consistent with the astrometric data (+ symbol). The background shows the HST/STIS image; the white line shows the loci of the peak of the main belt. Orbits are drawn in two segments between the ascending node (red dot) and the descending node (blue dot) with respect to the plane of the main belt. From the ascending node to the descending node the orbit is drawn as a dashed line (i.e., behind the sky plane); between the descending node and the ascending node the orbit is drawn as a solid line (i.e., in front of the sky plane). The inset shows a zoomed view (30 × 60 AU) where the astrometric data are plotted with a + symbol.

Standard image High-resolution image

Figure 21 shows the face-on and edge-on views of 30 orbits. The majority of ascending nodes relative to the belt are concentrated interior to the belt, near Fomalhaut b's periastron. The edge-on view emphasizes that the mutual inclination is most likely I ≲ 36° (90% confidence; Figure 22). Fomalhaut b's orbit is unlikely to intersect the main belt at a ∼ 140 AU because the main belt is relatively flat (the model-dependent opening angle for the belt is 1fdg5; Kalas et al. 2005) and the nodes are distributed at many locations interior and exterior to the belt.

Figure 21.

Figure 21. Sample of 30 orbits from Figure 20, viewed face-on (top) and edge-on (bottom) with the same orientation as the deprojected images shown in Figure 13. The dashed black line represents the main belt, with pericenter to the lower left. The ascending (red dots) and descending (blue dots) nodes with respect to the belt plane are mostly concentrated within the perimeter of the belt.

Standard image High-resolution image
Figure 22.

Figure 22. Posterior distribution (left) and cumulative distribution (right) of mutual inclination, I, between the orbit of Fomalhaut b and the main belt. The mean difference in inclination is 17fdg0 ± 12fdg0. 50% of allowed orbits lie within 13fdg3; 90% of allowed orbits lie within 36fdg3.

Standard image High-resolution image

The probability of Fomalhaut b directly interacting with the main belt depends on how the problem is defined, such as considering the size of Fomalhaut b's Hill sphere at the intersection region (which depends on the planet mass estimate) and the assumed physical boundaries of the belt. If Fomalhaut b is massive, then it can still gravitationally perturb a portion of the belt without crossing through it. To quantify the belt-crossing probability, we simply calculate the fraction of ascending and descending nodes that occur within various annuli representing the belt, without consideration of Fomalhaut b's mass and Hill radius. We find that 12% of nodes occur in the region 133 AU ⩽a ⩽ 158 AU. This 25 AU wide annulus was defined in the scattered light observations of Kalas et al. (2005) and it is roughly equal to the FWQM of the ALMA radial profile measurements at 870 μm (Boley et al. 2012). For a wider annulus starting from the belt inner edge at 133 AU, to the newly detected outer edge at 209 AU (Figure 11), the probability is 43%. These values suggest that the geometric deformations of the tentative belt detection beyond 209 AU may be dynamically linked to Fomalhaut b, whereas there is a smaller, ∼10% chance that Fomalhaut b interacts with the main concentration of belt mass near ∼140 AU.

Inspection of Figure 20 and Table 5 suggests that the orientation of the orbit of Fomalhaut b, within the uncertainties, is apsidally aligned with the main belt. The inclination of the orbit, ib = −55° ± 14° is similar to that of the belt (ibelt = −66°), and the longitude of the ascending node, Ωb = 152° ± 13°, is also consistent with that of the belt (Ωbelt = 156°). The posterior cumulative distribution of mutual inclination between the orbit of Fomalhaut b and the main belt is shown in Figure 22. The mean difference in inclination is 17° ± 12°. Fifty percent of allowed orbits lie within 13° and 90% of allowed orbits lie within 36°; the corresponding solid angles cover 0.8% and 5% of the sky respectively, indicating a small chance of this alignment occurring at random. Moreover the longitudes of periapse, Ω + ω, for Fomalhaut b and the belt are aligned within the errors (Figure 23).

Figure 23.

Figure 23. Posterior distribution of the longitude of periapse, Ω + ω, for Fomalhaut b. The vertical dotted line denotes the longitude of periapse of the main belt.

Standard image High-resolution image

Figure 24 demonstrates that periapse occurs interior to the main belt, behind the sky plane, and south of the star as projected on the sky plane. The posterior distribution of the periapse distance (Figure 25) has a mean value 32 ± 24 AU. Many of the Fomalhaut b orbits intersect the belt plane near periapse, suggesting that the region near periapse is where Fomalhaut's system may be most dynamically disturbed.

Figure 24.

Figure 24. Locations of periastra. The blue line traces the belt, the red dot is the geometric center of the belt, the yellow dot is the stellar location, and the green points represent the projected pericenters derived from the distribution of orbital elements.

Standard image High-resolution image
Figure 25.

Figure 25. Posterior distributions of the periapse (left) and apoapse (right) distributions.

Standard image High-resolution image

We repeated the MCMC fitting procedure using the affine invariant ensemble sampler of Goodman & Weare (2010), which is computationally efficient. This time we included the uncertainty in the mass of Fomalhaut (1.92 ± 0.02 M; Mamajek 2012) as a nuisance parameter. The resultant properties of the marginalized posterior probability density functions are listed in Table 6. These results, which were computed using a Python implementation of Goodman & Weare's algorithm by Foreman-Mackey et al. (2013), confirm our previous conclusions regarding the orbital elements.

Table 6. Properties of the Marginalized Posterior Probability Density Functionsa

Value a e i Ω ω
(AU)
Mode 162.2 0.851 −60fdg6 155fdg2 11fdg9
Median 170.0 0.846 −57fdg7 155fdg3 16fdg8
Mean 174.1 0.842 −56fdg1 155fdg5 17fdg2
rms 30.1 0.070 10fdg1 8fdg2 19fdg5
Percentiles          
0.05 126.6 0.694 −71fdg5 141fdg1 −19fdg2
0.95 242.9 0.952 −31fdg9 172fdg8 52fdg9

Note. aThe mass of Fomalhaut A (1.92 ± 0.02 M) is included as a nuisance parameter and marginalized over in these results.

Download table as:  ASCIITypeset image

9. DISCUSSION

Fomalhaut is emerging as an increasingly complex planetary system. It is therefore helpful to review and define its various elements before assessing the possible nature of Fomalhaut b.

9.1. Inventory of the Fomalhaut System

Figure 26 is a notional sketch of the Fomalhaut system, where we adopt a nomenclature based on the approximate positions of features in radius and position angle in the sky plane.

  • 1.  
    Fomalhaut A is the central A3V star (α PsA, HD 216956, GJ 881). Mamajek (2012) finds Teff = 8590 ± 73 K, L = 16.65 ± 0.48 L, M = 1.92 ± 0.02 M and age 440 ± 40 Myr. The heliocentric distance is 7.704 ± 0.028 pc and the angular radius for the stellar photosphere of 1.01 mas corresponds to 1.84 R (Di Folco et al. 2004). The position angle of the stellar spin axis is aligned with the minor axis of the main belt (Le Bouquin et al. 2009). Given the ∼16 × greater luminosity, the radiation environment at a given radius in the Fomalhaut system is roughly four times greater than for the solar system. The tidal radius, at, set by the Galactic tidal field is (Tremaine 1993):
    Equation (1)
    Assuming ρ = 0.11 M pc−3 (Holmberg & Flynn 2000), then at = 234 kAU (1.1 pc). Therefore the stellar companions Fomalhaut B and C, discussed below, are well within the sphere of Fomalhaut A's gravitational influence.
  • 2.  
    Fomalhaut B is a common proper motion stellar companion, also known as TW PsA. This is a K4Ve star with Teff = 4594 ± 80 K, L = 0.189 ± 0.013 L, and M = 0.73$^{+0.02}_{-0.01}$M. In the sky plane TW PsA is located 1fdg96 (55 kAU) southwest of Fomalhaut A, and therefore lacks a projected alignment with the major axis of the belt (Figure 27). Mamajek (2012) gives a three-dimensional project separation of 57.4$^{+3.9}_{-2.5}$ kAU. The heliocentric distances of Fomalhaut A and B are within 2σ of each other. No further information is currently available with respect to the possible orbit of Fomalhaut B. If 57.4 kAU is adopted as the semi-major axis value, then the orbital period is ∼8 Myr. We note that given a single astrometric observation of a binary separation, the most likely value for the orbital semi-major axis is the observed projected separation (Savransky 2011).
  • 3.  
    Fomalhaut C (LP 876-10) is a newly discovered M4V (M = 0.18 ± 0.02 M) proper motion companion to Fomalhaut A with 158 ± 2 kAU (0.77 ± 0.01 pc) projected separation from Fomalhaut A (Mamajek et al. 2013). Fomalhaut C is currently situated at P.A. = 336° relative to Fomalhaut A, which is equivalent to the position angle of Fomalhaut A's projected main belt semi-major axis.
  • 4.  
    Fomalhaut b could alternately be named Fomalhaut Ab. The present work revises the previous notion that Fomalhaut b's orbit is nested within the belt. Instead, Fomalhaut b's orbit is highly eccentric. In the 2012 epoch of observation, Fomalhaut b is 125 AU from the star assuming an inclination identical to that of the main belt. Apastron will likely be beyond 300 AU, where its velocity will be ∼1 km s−1. Fomalhaut b's current blackbody temperature is 50 K, whereas a 32 AU periastron and 322 AU apoastron give temperatures of 99 K and 31 K, respectively. The mass of Fomalhaut b is ≲ 1 MJ due to the non-detection at infrared wavelengths. Initially reported variability at optical wavelengths is not confirmed. The possibility that it is a resolved object also requires future observations for confirmation. If the optical brightness of Fomalhaut b is due to circumplanetary dust grain scattering, then compared to the present epoch it was approximately 16 times brighter at periastron (∼30 AU), and may become undetectable at apastron when it becomes at least eight times fainter.
  • 5.  
    The main belt is the primary source of far-infrared emission and is also prominent in optical scattered light (Kalas et al. 2005), with a sharp inner edge at a semi-major axis of 133 AU. The geometric center of the main belt is offset from the stellar location by 15 AU. The eastern hemisphere of the belt is brighter than the western hemisphere because the former lies out of the sky plane in the forward scattering direction, and the grain surface area is dominated by ∼10 μm sized grains that are preferentially forward scattering. The main belt has mass 1022–1024 kg in directly observed grains. However, given the age of the system, Wyatt & Dent (2002) argue that objects as large as a few km participate in the collisional cascade, yielding a total main belt mass of 20–30 M. Including primordial bodies as large as 1000 km that are not yet collisionally evolved, the belt mass could be near 1 MJ.
  • 6.  
    The main belt 331° gap refers to the approximate position angle of the azimuthal dust depletion detected in optical scattered light. In the deprojected reference frame it has FWHM ≈ 50 AU. As noted below, the region of the interbelt dust disk contains an arc of 450 μm emission located 190° from the 331° gap.
  • 7.  
    The main belt outer halo is a tenuous dust component extending to radii exceeding 200 AU. The scattered light color of the belt halo is currently unknown and the morphology may bend westward at large distances from the main belt.
  • 8.  
    The 10 AU belt is inferred from an unresolved component of 24 μm excess emission detected with the Spitzer Space Telescope (Stapelfeldt et al. 2004). This mid-infrared bandpass corresponds to blackbody emission with T ≈ 125 K, which at Fomalhaut is located at ∼20 AU radius (and roughly equal to the resolution limit of the observations). Re-analysis of these data suggest that the emitting grains can be constrained to lie between 8 and 12 AU, depending on their size and composition (Su et al. 2013). The significance of this region is that the blackbody grain temperature is 170 K, which is the canonical ice-line temperature in a circumstellar disk (e.g., Ida & Lin 2005; Kennedy & Kenyon 2008). Therefore the 10 AU belt could be called the "ice-line belt."
  • 9.  
    The interbelt dust disk. There is evidence for dust located inward from the main belt. Kalas et al. (2005) referred to an "inward intrusion of nebulosity" from the main belt to as close as ∼100 AU radius (a sensitivity limited value). This inner dust component is also detected as 24 μm (Stapelfeldt et al. 2004) and 70 μm thermal emission (Acke et al. 2012). Acke et al. (2012) assumed that the inner edge is ∼35 AU, stating that this is the 100 K water ice line of the system. Their best-fit model has grain surface density increasing linearly with radius out to the inner edge of the main belt (133 AU). The total grain mass is 4 × 1025 g, or half the total grain mass in their model fit to the main belt.
  • 10.  
    The interbelt 141° arc. The interbelt dust disk also contains an arc of 450 μm emission consistent with 0.075 lunar mass of dust located at ∼100 AU radius from the star (Holland et al. 2003). In the sky plane, the peak of arc emission is ∼4'' east and ∼5'' south of Fomalhaut. A background galaxy identified in Kalas et al. (2005) is 8farcs3 east and 8farcs0 south of Fomalhaut in HST observations made in 2004, approximately 3 yr after the SCUBA data were obtained. The proper motion of Fomalhaut would place the galaxy even farther away from the star in 2001 and therefore the optically detected galaxy is an unlikely explanation for the 450 μm arc. The position angle of peak emission in the arc is ≈141°, close to our estimate of Fomalhaut b's ascending node (152° ± 13°), ∼40° smaller than the longitude of periapse (178° ± 18°), and 190° away from the main belt 331° Gap. The relative geometry of these features may help in revealing the active dynamical mechanisms governing the Fomalhaut system.
  • 11.  
    The 25 AU radial gap refers to the radial location where the interbelt dust disk has a minimum mass, which lies just outside the boundary of the 10 AU belt. Periastron for Fomalhaut b it potentially located in this gap region. It is also notable that the 15 AU stellocentric offset of the main belt is located near the inner edge of the gap.
  • 12.  
    The hot disk is the region within a few AU radius from the star responsible for excess near-infrared emission (Absil et al. 2009). The 1000–2000 K grain temperature makes it a distinct component of dust from the 10 AU belt. Recently reported observations using the Keck Interferometer Nuller suggest that the hot disk could be further subdivided into <0.3 μm sized carbon-rich grains at ∼0.1 AU, and micron sized grains near ∼1 AU (Mennesson et al. 2013).
  • 13.  
    The 5 AU radial gap is the region between the hot disk and the 10 AU belt. The habitable zone for Fomalhaut A lies in the 2–5 AU region (Kasting et al. 1993).Given this inventory information, we evaluate several paradigms based on the possible orbits of Fomalhaut b.
Figure 26.

Figure 26. Notional sketch of the Fomalhaut system viewed face-on. The radial locations of features are approximate and not to scale. For example, the 10 AU belt represents dust near 10 AU radius, but the width of the belt is not precisely known.

Standard image High-resolution image
Figure 27.

Figure 27. Fomalhaut system. North is up, east is left, and the sky–plane separation between Fomalhaut A and B is 2fdg0. The background image is a false-color, log-scale, gnomonic (tan) projection from the optical Digitized Sky Survey 1 (red plates), centered on α = 344fdg00, δ = −28fdg00 (Credit: National Geographic Society, Caltech, STScI). The STIS Fomalhaut image is overlaid.

Standard image High-resolution image

9.2. Implications of Fomalhaut b's High e, Large a Orbit

The revised, larger values for e and a suggest that Fomalhaut b is not a planet that is solely responsible for the main belt stellocentric offset and sharp inner edge. Fomalhaut b's present dynamical state could be a consequence of an interaction with at least one other massive object that formed in the system. However, before exploring the scenarios that predict the existence of other Fomalhaut planets, is there a paradigm where Fomalhaut b achieves its dynamically hot state using only the inventory of observationally confirmed objects and structures presented above?

9.2.1. No Other Undetected Massive Bodies?

The stellar companions Fomalhaut B or C could disturb the Fomalhaut A system either by a close flyby interaction (Larwood & Kalas 2001; Kenyon & Bromley 2002; Ardila et al. 2005; Reche et al. 2009; Malmberg et al. 2011) or a secular perturbation (Augereau & Papaloizou 2004; Wyatt 2005). For example, a flyby interaction studied numerically for the β Pic debris disk gives a geometry that qualitatively resembles that of Fomalhaut A, Fomalhaut B, and the main belt (Larwood & Kalas 2001). An initially symmetric circumprimary disk of material perturbed by a close stellar encounter results in eccentric belts of material (technically, tightly wound spiral arms) where the apastra of the belts point toward the direction of the perturber's periastron. This means that the apastron of the perturber (or its post-flyby trajectory in a hyperbolic orbit) and the apastra of the eccentric belts are pointed in opposite directions. This simple geometrical picture is consistent with the present epoch location of Fomalhaut B south of Fomalhaut A, and the main belt and Fomalhaut b apastra to the north of Fomalhaut A (Figure 27). However, the critical problem is that if Fomalhaut B is bound to Fomalhaut A, and given a system age ≈400 Myr, there are repeated periastron passages that would wipe out the belt structure created by the first periastron passage.

Instead of a flyby, Fomalhaut B and C may influence the Fomalhaut A system via a secular perturbation. The Kozai resonance has been invoked as one mechanism to explain highly eccentric exoplanets (Wu & Murray 2003; Takeda & Rasio 2005). For Fomalhaut B or C to be responsible for a Kozai resonance, it must have a mutual inclination of >39fdg2 relative to the orbital planes of either Fomalhaut b or the main belt. If Fomalhaut b and the main belt are not coplanar, then it is possible that the Kozai mechanism operates only on one component of the Fomalhaut system. The approximate period for a Kozai oscillation (Ford et al. 2000) is

Equation (2)

Here the subscripts A, b, and B refer to the respective components of the Fomalhaut system. Using the approximate values of Pb = 103 yr, mA = 2 M, mB = 1 M, mb = 0, aB = 6 × 104 AU, ab = 102 AU and e = 0.5, the Kozai period is of order 1011 yr (and longer for the more distant Fomalhaut C). The Kozai resonance is therefore relatively ineffective for separations as large as observed between Fomalhaut A and B.

Instead of Fomalhaut B or C, the main belt mass may be responsible for a secular perturbation on Fomalhaut b. This scenario has been studied by Terquem & Ajmia (2010), but the initial conditions presume that a planet begins with a mutual inclination ≳ 30° relative to the main belt. To reach this starting point, the scenario needs to invoke an additional dynamical interaction with some other body, e.g., a Fomalhaut c, and therefore the planet-belt Kozai effect does not give a dynamical history consistent with no other massive bodies.

In addition to the problems of explaining Fomalhaut b's dynamically hot orbit, the properties of the main belt are left without adequate explanation. The eccentric orbit of Fomalhaut b tends to exclude the possibility that it dynamically sculpts the inner edge of the belt since only a small fraction of the belt could be disturbed during each orbital period of Fomalhaut b. Moreover, the secular perturbation theory invoked to explain the main belt stellocentric offset is second order with respect to eccentricity (Wyatt et al. 1999) and breaks down at high eccentricity. If secular theory is applicable, then Fomalhaut b's high eccentricity would predict that the main belt's eccentricity should be larger than observed. On the other hand, apsidal alignment between Fomalhaut b and the main belt continues to be indicated by the new orbit determination (Table 5). Future work needs to determine if the orbital parameter space presented here could be consistent with secular theory and the observed stellocentric offset. In any case, the main belt's sharp inner edge and the azimuthal gap are consistent with the existence of another planet orbiting near the main belt.

To summarize, the observed Fomalhaut inventory (Section 9.1) does not appear to be sufficient for explaining Fomalhaut b's high eccentricity and the main belt morphology. Other perturbing objects must be present in the dynamical history of the Fomalhaut system.

9.2.2. Additional Fomalhaut Perturbers

Permitting the existence of additional perturbers in the past and/or present epochs allows a variety of plausible dynamical histories that are consistent with the current observables. Three classes of dynamical paradigms could focus on endogenic perturbations, exogenic perturbations, or a blend of both. For example, the dynamical paradigm of Oort cloud comets is a blend that involves the increasing of minor body semi-major axes and aphelia by close-encounters with gas giant planets, followed by the raising of perihelia by passing stars and molecular clouds (Oort 1950; Duncan et al. 1987).

In the endogenic class of paradigms, Fomalhaut b's eccentric orbit was produced by an interaction with at least one other planet in the system. The general idea for planet–planet dynamical interactions is that two or more planets initially form in relative isolation from each other, but subsequent migration mechanisms lead to unstable orbital configurations. Two planets may enter within a few times their mutual Hill sphere (Gladman 1993; Chambers et al. 1996; Rasio & Ford 1996; Levison et al. 1998; Marzari & Weidenschilling 2002; Adams & Laughlin 2003; Veras & Armitage 2004; Chatterjee et al. 2008; Jurić & Tremaine 2008; Veras et al. 2009), or their orbits may cross into an unstable resonance due to planetesimal-driven migration (Tsiganis et al. 2005; Thommes et al. 2008). An instability that modifies the orbital elements of two planets in the system may then lead to unstable orbits for other planets in the system, producing a "global" instability.

Because planet–planet scattering evolution is chaotic, the initially closest planet may end up the farthest and vice versa. Overall, the surviving (i.e., not ejected) planet that ends up with the largest semi-major axis will also have higher eccentricity if its mass is less than or equal to the planet it interacted with. Unfortunately, the upper mass limit of Fomalhaut b (⩽1 MJ) is not particularly helpful for constraining the expected mass of another surviving planet. Moreover, it is also possible that a hypothetical Fomalhaut c was ejected, leaving behind Fomalhaut b as the interior planet. For example, numerical tests by Ford & Rasio (2008) involving two planets indicate that the largest eccentricities are obtained for near equal mass planets, where the surviving (bound) planet has e = 0.624 ± 0.135, but the second planet is lost. Jurić & Tremaine (2008) find that 20% of simulations that begin with multiple planets end with only one bound planet. However, the majority of systems have at least two surviving planets after chaotic evolution, agreeing with simulations of three-planet systems conducted by Chatterjee et al. (2008). Therefore, the detection of Fomalhaut b as a large-a, high-e exoplanet makes the existence of another comparably massive exoplanet in the system more likely than not. This would also mean that the orbit of Fomalhaut b may undergo further dynamical interactions that will evolve the orbits of both Fomalhaut b and Fomalhaut c.

The observational avenue for constraining the problem clearly rests on detecting Fomalhaut c and determining its orbital properties. Direct imaging surveys to date have not detected a second companion at infrared wavelengths (Kalas et al. 2008; Marengo et al. 2009; Kenworthy et al. 2009, 2013; Janson et al. 2012). Since planet–planet scattering or other instabilities may involve a Fomalhaut c with a Jupiter mass or below, the mass limits explored by these surveys, >1 MJ, are not adequate to rule out the existence of a Fomalhaut c.

In the exogenic class of perturbations, an interloping star could have been responsible for perturbing the Fomalhaut system. Deltorn & Kalas (2001) searched for Fomalhaut "nemesis" encounters among 21,497 stars where space motions could be derived from radial velocity and Hipparcos information. The strongest perturbation was from HD 16895 (HIP 12777; SpT = F7V) $474^{+20}_{-19}$ Myr ago. The age of HD 16895 is ∼9 Gyr (Ng & Bertelli 1998), and therefore it did not form as part of the Fomalhaut system. The closest approach distance was $1.15^{+0.41}_{-0.34}$ pc, at which time $105^{+21}_{19}$ kAU represents the approximate radius of a sphere centered on Fomalhaut A where a particle experiences equivalent gravitational forces from Fomalhaut A and HD 16895. Given the present-day separation of 158 kAU (0.77 pc) between Fomalhaut A and C, there is a possibility that Fomalhaut C was gravitationally perturbed by HD 16895. Therefore there is some empirical evidence for a possible exogenic disturbance to the system that could propagate inward, resulting in a global dynamical instability on a secular timescale (e.g., Zakamska & Tremaine 2004). The availability of expanded position, proper motion, and radial velocity catalogs may be used to identify other potential perturbers in future work, and the effect of the galactic tides should also be incorporated in new calculations (Kaib et al. 2013; Veras & Evans 2013).

A blend of endogenic and exogenic perturbations requires a comprehensive analytical and numerical analysis. To gain a rough picture concerning the dynamical lifetime and outcomes of the current orbital configuration, we used the numerical simulator AstroGrav to evolve the orbits of several test cases for 440 Myr. The simulations include Fomalhaut A and B, two planets orbiting Fomalhaut A, but have no test particles representing the belt to minimize the simulation times. One of the test planets represents Fomalhaut b with a = 177 AU and e = 0.8, and the mutual inclination with the second planet is either i = 0° or i = 20°. The second planet is either at a = 30 AU or a = 120 AU. We tested a combination of various masses for the two planets representing Jupiter, Saturn, and Neptune. Fomalhaut B has mass 1.45 × 1030 g, a = 57400 AU, e = 0.0 and i = 0°.

The general outcome is that if Fomalhaut b is coplanar with Fomalhaut c, it is ejected from the system on <107 yr timescales, though there are exceptions where Fomalhaut b survives for the age of the system. In a small fraction of cases, Fomalhaut b approaches Fomalhaut B as its semi-major axis evolves to large values, but capture is unlikely. The exogenic influence of Fomalhaut B appears minor given the fixed assumption of a = 57,400 AU, e = 0.0. Fomalhaut b remains bound to Fomalhaut A for >107 yr timescales in the test cases where the mutual inclination is 20°. We also found cases where Fomalhaut c has high eccentricity (e ∼ 0.8) after Fomalhaut b is ejected. This confirms the previously stated notion that the observed Fomalhaut b could have been a planet on a low-eccentricity orbit, as originally envisioned to account for the main belt properties, but recently acquired high eccentricity via a planet–planet scattering event.

The overall picture is that given the uncertainties concerning the orbital parameters of Fomalhaut b, Fomalhaut B, and the existence of other Fomalhaut A planets, there are configurations where Fomalhaut b obtained its high eccentricity >107 yr ago at early epochs, particularly in the non-coplanar cases. However, there are circumstances where the configuration is younger than 107 yr. A test of which scenario should be favored could look into how likely the belt is to survive in either case, though in such a scenario the assumed mass of Fomalhaut b is increasingly relevant. In the next section we consider the belt survival timescales and other physics that would be implied by a co-planar case, given a variety of masses for Fomalhaut b up to one Jupiter mass.

9.3. Belt Collision Scenarios

In the coplanar scenario, Fomalhaut b is on a collision course with the main belt. Fomalhaut b will begin entering the inner edge of the dust belt around 2032 C.E., at which point the emergent phenomena would elucidate the physical nature of Fomalhaut b. For example, if Fomalhaut b's optical light is due to a dust cloud, it may appear to episodically brighten and change color in scattered light as fresh dust rich in smaller grains is produced by collisions with main belt material. The direction of the main belt orbital motion may also be ascertained depending on which direction new features propagate within the belt. We note that even though the probability of a belt-crossing orbit is of order 10%, the nodes may precess or librate, producing intervals where belt crossing occurs. The belt collision scenario is therefore worth studying even if at the present epoch Fomalhaut b is in a configuration that does not physically intersect the belt.

In Section 9.3.1 we study the case where Fomalhaut b is a relatively low-mass planetesimal surrounded by a dust cloud. We tackle two questions. (1) Can such a dust cloud survive belt passages? (2) What is the lowest mass for Fomalhaut b such that the dust cloud remains bound to it after gravitational shearing occurs at periapse?

In Section 9.3.2 and following we consider the consequences of assuming Fomalhaut b consists of a planet mass. We study the issues of whether or not the main belt will be disrupted (Section 9.3.3), the possibility of observing energetic impacts on the planet (Section 9.3.4), whether or not a giant impact could be the origin of the main belt (Section 9.3.5), and if the main belt gap could be the result of planet crossing (Section 9.3.6).

9.3.1. Planetesimal with a Dust Cloud

Here we assume that Fomalhaut b is a low-mass planetesimal that is optically bright because of reflected light from a fresh dust cloud surrounding it. For example, it could be a planetesimal that was recently disrupted by forces associated with its recent periaston passage. Fomalhaut b is unlikely to be only a dust cloud (i.e., only ⩽1 mm sized grains) because the size of the object required to account for the grain scattering surface area is at least 10 km in size (Kalas et al. 2008). Therefore it resides within the gravity regime of planetesimal collision physics, where "catastrophic" collisions are defined as retaining 50% of the precursor mass in a largest remnant.

An alternative to a collision is tidal, thermal, and/or spin breakup of a weak planetesimal (e.g., Jewitt 2012). Fomalhaut b's precursor could be an analog to Shoemaker–Levy 9 (SL9), tidally disrupted by passing within the Roche radius of the hypothetical Fomalhaut c. Alternately, the analogy may be to a Sun-grazing comet that breaks up near periastron due to thermal and tidal stresses, or elsewhere due to spin (Marsden 2005). One empirical test of this idea is to search for debris along Fomalhaut b's orbital path (Figure 20). Unfortunately, the current data are dominated by speckle noise in most of the region closer to the star than Fomalhaut b's current location. One might classify the Fomalhaut b phenomenon as cometary, but the inferred dust mass and stellocentric distance places it in the "giant" comet category with activity involving supervolatiles, as inferred for the activity of comet Halley and other icy objects at large heliocentric distances (Sekanina et al. 1992; Jewitt 2009).

Does a Fomalhaut b dust cloud survive the belt crossing as it collides with main belt material? A key consequence of Fomalhaut b's e ∼ 0.8 orbit is that the relative velocity of Fomalhaut b with respect to material in the belt is greater than previously assumed. Therefore the collision lifetime, tcc, of cloud particles will be shorter by some scaling factor compared to the lifetime of particles that collide with each other in the main belt. The collision lifetime of dust grains within the main belt has been analytically determined by Wyatt & Dent (2002): 10 μm grains have 105 yr ⩽ tcc ⩽ 106 yr. Therefore our task is to estimate the appropriate scaling factor pertinent to Fomalhaut b's relative velocity.

The relative velocity of particles orbiting within the belt is (Wyatt & Dent 2002)

where vk is the Keplerian orbital velocity. The collisional belt model adopted by Wyatt & Dent (2002) assumes that the belt lies between 125 and 175 AU radius, and the average inclinations and eccentricities of belt particles are I = 5° and e = 0.065. At 150 AU, vk = 3.4 km s−1, yielding vrel = 0.4 km s−1.

To calculate vrel between Fomalhaut b and the belt as it enters the belt two decades from now, we assume that the incidence angle is 45° in the prograde sense. The model belt is 50 AU wide and the oblique path through the belt has length 71 AU. The entry point is 150 AU from Fomalhaut A (recall the stellocentric offset), where the orbital velocity of Fomalhaut b is ∼3.7 km s−1, and hence the belt crossing requires ∼100 yr. The velocity components of Fomalhaut b are 2.6 km s−1 both parallel and orthogonal to the disk velocity vector. Thus, while cutting diagonally through the main belt, Fomalhaut b is rear-ended by belt material moving faster in the parallel direction at vk = 3.4 km s−1, or vrel = 0.8 km s−1. Fomalhaut b will also undergo head-on collisions with an orthogonal component vrel = 2.6 km s−1. In the reference frame of Fomalhaut b, $v_{{\rm rel}} = \sqrt{(}0.8^2 + 2.6^2)$ = 2.7 km s−1 from the lower right (cf. inset of Figure 20).

Given that vrel for Fomalhaut b is six to seven times greater than for belt particles, the catastrophic collision timescale could then be taken as proportionally shorter for dust grains surrounding the planetesimal compared to dust grains colliding with each other in the belt. The dependence is in fact stronger because smaller and smaller particles become catastrophic impactors as the relative velocity increases. Under the assumption that smaller impactors are present, the collision timescale (s):

where α is the exponent for the particle size distribution dependence, taken here as α = 7/2 (Wyatt et al. 2007, 2010, 2011). Therefore the collision timescale for Fomalhaut b dust-cloud particles passing through the belt is (2.7/0.4)8/3 = 163 times shorter than the collision timescale of particles within the belt. This value has significant uncertainties that depend on how the grain size distribution of the planetesimal cloud differs from that of the belt.

Counterbalancing this is the fact that Fomalhaut b spends only ∼200 yr in the belt per orbital period, which is ∼15% of the orbital period of a belt particle at 150 AU. Therefore the collision timescale of Fomalhaut b is one order of magnitude shorter than a belt particles instead of two orders of magnitude.

This estimate for the scaling factor suggests that the catastrophic collision timescale for a 10 μm grain bound to Fomalhaut b is 104 yr ⩽ tcc ⩽ 105 yr. A Fomalhaut b dust cloud would survive belt passages for 10–100 orbital periods.

The survival of a Fomalhaut b dust cloud after many crossings through the main belt appears counterintuitive. We therefore conduct an order of magnitude check based on the observables in the optical data, rather than the above extrapolation from the analytical analysis given by Wyatt & Dent (2002). We begin by assuming that the lifetime of Fomalhaut b as a dust cloud is roughly equal to the timescale for intercepting its own mass in main belt dust grains. For this simple scenario we assume that in fact all of the dust cloud interacts with main belt material, and that both cloud and belt have a uniform number density of objects for any given grain size.

Regardless of whether or not the scattering grains are in a small cloud, large cloud, ring, or any other geometry, the optical photometry constrains the geometric scattering cross sections of grains that comprise the structure. We use the relationship derived by Kalas et al. (2008),

where σp is the projected geometric surface area of scattering grains in m2, and Qs is a scattering efficiency factor, such as the geometric albedo. Observations give mp ∼ 25 mag in the optical. Therefore if the cloud material has a relatively low albedo such that Qs = 0.1, then the projected geometric surface area of grains within the Fomalhaut b cloud is σp = 1.2 × 1019 m2. Turning now to the main belt, Kalas et al. (2005) give the model dependent total grain scattering cross section σmb = 5.5 × 1022 m2, given an assumed albedo = 0.1. The planetesimal cloud therefore has a total grain cross section that is 2.2 × 10−4 of the entire optically detected main belt.

As an aside, these results are testable in the submillimeter, where the flux of the entire main belt is 81 mJy (Holland et al. 1998), yielding a predicted flux from Fomalhaut b of ∼18 μJy (if the circumplanetary grains have similar properties as grains in the main belt). The current ALMA data have an rms noise of ∼60 μJy beam−1 (Boley et al. 2012), which to first order excludes a circumplanetary dust cloud significantly more massive than in our prediction above. Vetting various dust-cloud models will require additional future observations with ALMA.

For a single belt crossing, the cloud will encounter only a fraction of the main belt volume. If Fomalhaut b requires 100 yr to cross the belt, then the volume of the main belt that is encountered by the grains within the cloud is 1.1 × 1032 m3. Adopting the assumptions from Wyatt & Dent (2002) that the belt has an inner and outer radius of 125 and 175 AU, respectively, and adopting a fixed vertical width of ∼10 AU, the volume of the model belt is 1.6 × 1039 m3. Thus one encounter volume is 6.9 × 10−8 of the total main belt volume. Multiplying σmb by this factor, the geometric surface area of main belt grains encountered by cloud grains in a single belt crossing is 3.8 × 1015 m2. For the dust-cloud grains to intercept an equal surface area of belt grains requires 3 × 103 belt crossings. Given two belt crossings per 1700 yr orbital period, the lifetime of the cloud is ∼106 yr.

To summarize, both the empirically based, order-of-magnitude approach and the analytic approach of Wyatt & Dent (2002) confirm that a dust cloud is not destroyed by a single belt crossing, and in fact survives for a minimum of 10 orbital periods.

Is there a shorter timescale by which the cohesiveness of the dust cloud could be lost? If the cloud is not gravitationally bound to itself, then two possibilities are orbital shearing between the portions of the cloud closest and farthest from the star, and the velocity dispersion of grains. For shearing, if the semi-major axis of one side of the cloud differs from the opposite side by 2 AU (260 mas; 5 STIS pixels) then the orbital velocities differ by 10−2 km s−1. In 10 yr the separation along the direction of orbital motion increases by ∼0.02 AU (3 mas), which is not detectable. However, in one orbital period (∼2000 yr) the cloud shears by 4.6 AU (600 mas). A spherically symmetric cloud of grains with no self-gravity therefore spreads into a triaxial spheroid and eventually into a trail within a few orbital periods.

Figure 28 demonstrates the shearing of a spherical cloud that begins with radius 0.05 AU at a position −90° from periapse. We use the AstroGrav numerical model to study the evolution of a dust cloud composed of 1000 massless particles, with a stellocentric motion that follows the nominal Keplerian orbit of Fomalhaut b (a = 177 AU, e = 0.8, i = 0°). Shearing at periapse extends the structure such that by 200 yr when it is in Fomalhaut b's position near belt crossing, the diameter is 0.5 AU. This corresponds to 65 mas and would be unresolved by the observations. However, after the second periapse passage the structure is 9 AU in length, resembling a trail of material along the path of the orbit. Thus, even though Fomalhaut b as a dust cloud could survive many belt crossings, it shears into a trail of particles in one orbit if it is not gravitationally bound to a more massive object. We therefore consider it unlikely that Fomalhaut b is only a dust cloud, because it requires a fortuitous timing in discovering it a few centuries after it was created.

Figure 28.

Figure 28. Numerical model of a massless dust cloud which at t = 0 has radius 0.05 AU and is located east of Fomalhaut A. At t = 50 yr it is near periapse and after periapse the structure resembles a triaxial spheroid oriented roughly north–south in our line of sight. By t = 200 yr, the structure is 0.5 AU in diameter (each inset box is 0.82 AU on a side).

Standard image High-resolution image

How massive does a central object have to be so that an 0.05 AU radius cloud is not disrupted by shearing at periapse? Using the numerical simulation, we found a mass of 1.0 × 1023 kg is sufficient (1.4 lunar mass). The cloud can be smaller, since Kalas et al. (2008) suggested that a circumplanetary disk with ∼30 RJ (2.1 × 109 m or 0.014 AU) is consistent with the detected optical flux. For this smaller radius, the cloud is stable against shearing if the central object has mass 5 × 1021 kg (5× Ceres mass). Therefore a dwarf planet between Ceres and Pluto in mass may retain a system of satellite dust and moons that is stable against shearing.

The dust-cloud model in Kalas et al. (2008) specifies that the dust cloud's scattering surface area corresponds to the disruption of a minimum 10 km radius object. Since dwarf planets are much larger, ∼500 km radius objects, the mass in the dust cloud could have been launched from a single cratering impact.

Finally, given a scenario that Fomalhaut b is a dwarf planet, is it possible that its mass is still in the process of increasing significantly due to the accretion of belt material during belt passages? The mass accretion rate (kg s−1) enhanced by gravitational focusing is (Kennedy & Wyatt 2011)

where the main belt total mass and volume are estimated as Mmb = 75 M and Vmb = 1039 m3, and the dwarf planet radius and escape velocity are Rb = 106 m and vesc = 1000 m s−1, respectively. Therefore for a relative velocity vrel = 2700 m s−1, we find dM/dt = 1.4 × 1011 kg yr−1. Since each orbital period consists of only 200 yr spent within the belt, the mass accretion rate is equivalently expressed as 2.8 × 1013 kg orbit−1. If we assume that 90% of this accreted material adds to the mass of the central object and 10% adds to the mass of the planetesimal cloud, then we have the following results: in ∼3000 orbits (5 × 106 yr) Fomalhaut b has accreted ∼1016 kg of additional mass into the planetesimal cloud, which is equivalent to the 10 km sized object that was originally envisioned to explain the grain scattering cross section. However, the central mass has increased by only a factor of 10−5. The 5 × 106 yr timescale is of order the lifetime we might expect for the coplanar case where Fomalhaut b crosses through the planetary region, leading to an eventual strong scattering event with another planet. We therefore conclude that even though the coplanar, belt-crossing orbit is most likely short-lived, it is long enough for a dwarf planet to capture a surrounding cloud, but not long enough for the dwarf planet to increase its mass significantly.

9.3.2. Planet with a Satellite System

The circumplanetary dust disk hypothesis presented by Kalas et al. (2008) received a measure of plausibility with the discovery of Saturn's Phoebe ring at >200 Rp (Verbiscer et al. 2009). The basic physical mechanism is that the surface of a small (radius ∼100 km), distant (a = 215 Rp) planetary moon is bombarded by interplanetary meteoroids, launching ejecta that spirals toward the planet due to Poynting–Robertson drag. Verbiscer et al. (2009) estimate that the normal optical depth of the Phoebe ring is ∼2 × 10−8, which could be attributed to material ejected from a single, 1 km diameter crater.

One consequence of the highly eccentric orbit for Fomalhaut b is that it is less likely to capture additional outer satellites compared to nested planets such as Saturn because its velocity relative to nested orbiting objects is high, and stellar gravitational shearing forces at periastron are significant. However, the Phoebe ring scenario only requires the existence of a single distant satellite, and certainly the prospect that Fomalhaut b previously had a lower eccentricity orbit is not ruled out. Moreover, a planet–planet scattering event that could account for Fomalhaut b's high eccentricity does not necessarily lead to the loss of the moon orbiting the scattered planet (Debes & Sigurdsson 2007; Nesvorný et al. 2007).

Instead of a single Phoebe-like satellite, Kennedy & Wyatt (2011) study the capture of many irregular satellites around Fomalhaut b that collisionally produce a circumplanetary dust cloud. This collision concept is different because the Phoebe ring is produced when a satellite is "stranded" far from the next innermost satellite, and a large fraction of impactors that strike the moon originate from outside the system. If there are numerous irregular satellites like Phoebe, then this "swarm" is self-eroding via many mutual collisions. The morphology of the dust swarm resembles an hourglass instead of a ring or torus due to the instability of high inclination moons (Hamilton & Burns 1991; Nesvorný et al. 2003). Krivov et al. (2002) present evidence that Jupiter is surrounded by a cloud of dust particles between 50 and 300 RJ, except that given the relatively few irregular satellites at this late epoch, the erosion mechanism is mainly the external meteoroid flux, as with Phoebe, instead of self-erosion.

The e ∼ 0.8, a = 177 AU orbit for Fomalhaut b changes the Kennedy & Wyatt (2011) scenario in that the Hill radius is effectively three times smaller at periastron, and so too is the region of satellite stability. The Hill radius depends on the planet and stellar masses, and the planet semi-major axis, apl:

However, in the case of an eccentric orbit, the instantaneous star–planet separation, ρ, should be adopted instead of apl. At an assumed periastron of q = ρ = 32 AU, and with mpl = 1 MJ and M = 2 × 103 MJ, the Hill radius is 1.76 AU. For comparison, the Hill radius for our Jupiter and Neptune are 0.33 AU and 0.77 AU, respectively, and a perihelion scaling reduces these values by only a few percent. As Fomalhaut b intersects the belt at ∼150 AU in the coplanar case, the instantaneous Hill radius is 8.24 AU.

Fomalhaut b therefore represents an interesting case study for the dynamical evolution of moons and the observational consequences when the host planet has a highly eccentric orbit. To gain some rough insights, we used the N-body package AstroGrav to study the evolution of 500 moons, randomly assigned 0.01 AU ⩽ amoon ⩽ 10 AU, 0.0 ⩽ emoon ⩽ 0.1, with a spherical distribution of orbits around a Jupiter-mass planet that has Fomalhaut b's orbital properties. After 2 × 105 yr, approximately 50 moons remain bound to the planet with 0.02 ⩽ amoon ⩽ 0.91 AU, with median values a = 0.37 AU. The maximum value of 0.91 AU is 52% of the Hill radius calculated at periastron. This result is consistent with previous observational and theoretical studies concerning the dynamical evolution of distant satellites orbiting asteroids and planets (Hamilton & Burns 1992; Jewitt & Haghighipour 2007; Shen & Tremaine 2008)

The majority of moons lost near periastron orbit the star at a reduced semi-major axis and eccentricity, forming an eccentric disk interior to the main belt, and apsidally aligned with Fomalhaut b. It is possible that such lost moons exist if Fomalhaut b's orbit before the dynamical instability had a larger Hill sphere because the orbit was initially farther from the star. An instability that subsequently reduces the star–planet separation would then result in a smaller Hill sphere and lost moons orbiting the star instead of the planet. The implication is that the interbelt dust disk may consist of material with different origins: (1) bodies that formed there, (2) moons lost by planet instabilities, and (3) material perturbed inward from the outer disk (cf. Section 9.3.3). This scenario also invokes the possibility that if the perturbed planet is not coplanar with the main belt, the eccentric disk of lost moons would also orbit and collisionally evolve in the planet's orbital plane and not the main belt plane. The system would therefore appear to have two inclined debris disks. The β Pic system shows evidence for a secondary, inclined disk that is significantly less massive and prominent than the primary disk (Heap et al. 2000; Golimowski et al. 2006). The concept of lost moons after a planetary dynamical instability could serve as an alternate model to the current paradigm that inserts an inclined planet into a pre-existing disk, creating a vertical disk warp that propagates outward (Mouillet et al. 1997).

Since Fomalhaut b and the lost moons continue to have a similar periapse, a bottleneck of orbits is evident near periapse. Even though Fomalhaut b's Hill radius is at a minimum here, the volume number density of lost moons is greatest near periapse. We find that moons lost at periastron may be recaptured near periastron (ejection from the system or collision with the planet are also possible). In fact, the recapture epoch begins after periastron when the Hill sphere of the planet is expanding but the bottleneck is still providing a relatively high volume number density of objects. Recaptured moons tend to be captured into eccentric (e ≳ 0.3) orbits around the planet, which means that they are loosely bound and lost again as the planet approaches the next periastron passage.

Even if the recapture of moons does not occur, the outer moons still bound to Fomalhaut b are dynamically heated by periastron passage. The eccentricities found in the surviving, bound moons are 0.0 ⩽ emoon ⩽ 0.9, with median emoon = 0.1. Thus there are at least three mechanisms that could increase collisional dust production surrounding Fomalhaut b at periastron and soon after periastron: objects have energetic collisions with Fomalhaut b at periastron, the collisional grinding of moons bound to Fomalhaut b is enhanced at periastron, and soon after periastron additional moons may be recaptured on highly eccentric orbits that would collide with the bound moon system. Fomalhaut b is currently observed ∼120 yr after periastron, and the simulation supports the concept that moon capture may have recently activated the collisional erosion of a planetary moon system.

Two more epochs of enhanced collisionally activity may occur at the ascending and descending nodes relative to the belt plane. When Fomalhaut b crosses the orbital plane of the belt or the interbelt dust disk, the external meteoroid flux is enhanced again.

Observationally, the optical depth of the Fomalhaut b ring/cloud system will increase during enhanced erosion, the grain size distribution will shift temporarily to small sizes as fresh dust below the radiation pressure blowout radius is released, and as a result we might observe a brighter and bluer scattered light signature from Fomalhaut b. On the other hand, a dust cloud could also become optically thick, which would make light scattered toward the observer sensitive to viewing geometry. In other words, shadowing due to too rapid dust production could decreases the brightness.

Though this discussion focuses on the erosion of satellites from both circumplanetary and circumstellar impactors, generating fresh circumplanetary dust that is observable in reflected light, other processes may be at work that have distinct observable signatures. When the bound moons have their eccentricities pumped by periastron passage, this could increase the planetary tidal heating of the hypothetical moons (Peale & Cassen 1978; Peale et al. 1979; Cassen et al. 1979). Tidal heating and melting has an infrared signature (e.g., Peters & Turner 2013 and references therein). A clone of Jupiter's regular moon Io would also generate an optically detectable sodium cloud. Jupiter's sodium cloud has been detected in Sun scattered light to at least 400 RJ (0.2 AU; Mendillo et al. 1990). Thus, in addition to dust scattered light and Hα emission as possible explanations for Fomalhaut b's anomalously high optical flux (Kalas et al. 2008), a sodium cloud could also contribute at 0.59 μm. This lies in the F606W bandpass of the 2004 and 2006 HST/ACS observations.

Finding definitive evidence for such a cloud would have many significant implications, such as showing that Fomalhaut b has a magnetic field similar to Jupiter's. Clearly a spectrum of Fomalhaut b is required, but the issue can also be examined via imaging. For example, Jupiter's sodium cloud is variable in size and brightness. These variations are correlated to the volcanic activity of Io (Mendillo et al. 2004). Therefore, the characteristic timescales of variability from imaging Fomalhaut b may be more geophysical than astrophysical—Fomalhaut b may episodically appear brighter and more extended on timescales measured in months.

We have argued that even though variability and extended morphology exist in the HST optical data (Figures 2 and 4), they can also be attributed to instrumental noise (Section 2.2; Figure 6). However, Galicher et al. (2013) claim that the extended morphology of Fomalhaut b in the 2006 ACS/HRC data is not instrumental in the F814W image. This is puzzling because the F606W image taken at the same epoch and with slightly greater sensitivity does not appear to be extended. This is difficult to reconcile with a model where optical light arises from grain scattering—the F606W image should also show extended morphology. However, the more complicated model involving the evolution of atomic and molecular species from moon volcanism to cirumplanetary magnetospheres could yield a solution. For example, Io is also the source of a circumplanetary potassium cloud that emits at 0.77 μm (Trafton 1975), and this lies within the F814W bandpass. Jupiter's potassium cloud is significantly weaker and less extended than the sodium cloud, but a different geochemistry for the hypothetical Fomalhaut b moon could conceivably produce a potassium cloud that adds a halo of extended light in the F814W images.

9.3.3. Disruption of the Main Belt by Planet Crossings

A key question is whether or not a planet mass passing through the belt disrupts the morphology of the belt. The maximum radius of gravitational influence could be taken as approximately three times the Hill radius (Section 9.3.2) at 150 AU. For a Jupiter mass this corresponds to 25 AU (6farcs5) diameter. This size is 3.2 times smaller for a 10 M planet, but the corresponding 2farcs0 angular scale is still resolvable with current instrumentation. In principle, the local dynamical stirring should enhance dust production, shifting the grain size distribution to favor smaller grains with larger surface area, and produce a transient brightening of the belt in scattered light (e.g., Kenyon & Bromley 2001; Dominik & Decin 2003).

To explore the cumulative effects of Fomalhaut b's dynamical perturbations on belt parent bodies over many belt crossings and under a variety of assumptions, we use the three-dimensional, N-body simulator AstroGrav. Our model of the Fomalhaut system begins with a nested Jupiter-mass planet (Fomalhaut c) with a = 120 AU, e = 0.10, i = 0 and ϖ = 178°. We add two populations of main belt objects. First, an effectively massless population of 8000 objects are randomly assigned 140 AU ⩽ a ⩽ 160 AU, 0.09 ⩽ e ⩽ 0.11, and −1fdg5 ⩽ i ⩽ 1fdg5. The orbits are apsidally aligned with the nested planet and randomly distributed in orbital phase. The second population is 200 objects between Ceres and Pluto in mass and radius (total mass = 0.057 M), distributed randomly throughout the belt as in the first population. However, the masses are not negligible and will gravitationally perturb other objects. Collisions are treated as mergers.

The first part of the simulation does not contain Fomalhaut b (assumed to be on a circular orbit well within the orbit of Fomalhaut c and dynamically negligible). The goal is to reach a quasi-steady state with respect to the dynamical sculpting of the belt's inner edge by Fomalhaut c. The entire system is coplanar and integrated for 1.5 × 105 yr (∼1500 orbits). The model qualitatively reproduces the numerical experiment of Chiang et al. (2009) where Fomalhaut c maintains a sharp inner edge. A key difference is that Chiang et al. (2009) proceeded to model the dynamics of dust particles in addition to the parent bodies, where radiation pressure instantaneously increases the eccentricities of the observed dust population. Since we are studying only the parent bodies, the timescales given below may be considered upper limits to the eccentricity evolution of observable particles. Another difference is that particles approaching within a few times the Hill radius are not removed from our simulation, and therefore we observe the capture (and loss) of satellites.

The second part of the simulation assumes that after 1.5 × 105 yr, Fomalhaut b is strongly perturbed from the inner part of the system (by interaction with a third planet) and appears as a rogue planet with a = 177 AU, e = 0.8, i = 0°, and apsidally aligned with Fomalhaut c. We assume three cases where the rogue planet has a Jupiter, Saturn, and Neptune mass. We use the "roque" terminology to designate bound planets with large a and e such that they cross the orbits of other planets and belts in the system.

The impulse imparted on belt material by a planet crossing through the belt does not create a visually noticeable disruption of the overall belt morphology. Inspection of the belt particle orbital elements shows no statistically significant difference before and after planet crossing. For example, after the nested Jupiter-mass Fomalhaut c has been sculpting the belt for the first 1.5 × 105 yr, the mean eccentricity distribution of belt particles has evolved to e = 0.1077 ± 0.0747. Fomalhaut b as a rogue Jupiter mass would have the most significant dynamical effect, yet after a single crossing the eccentricity distribution is e = 0.1072 ± 0.0746. Measured another way, before the belt crossing, 8.07% of belt particles have e > 0.20. After the single crossing of a Jupiter, 8.17% have e > 0.20.

The cumulative effect of many belt crossings is to gradually spread the belt radially. Over 102–104 belt crossings, the only belt morphology that is noticeably "disrupted" is the sharpness of the belt boundaries. The belt becomes a disk.

Kalas et al. (2006) noted that debris disks appear either as extended disks or narrow belts. The numerical models studied for the Fomalhaut system suggest that when a planet transitions from a nested to a rogue orbit, belts become disks. The rogue can subsequently evolve away from a belt-crossing orbit, and the disk will then resemble a belt again by interaction with the nested planets, though with reduced mass due to the scattering of objects during the rogue phase. The Fomalhaut system could be in transition from belt to disk, depending on the mass and orbit of Fomalhaut b.

The rogue planet in our model competes with the nested planet in shaping the inner edge (Figure 29). The qualitative result is that Fomalhaut b as a coplanar Saturn erodes the belt edge significantly after only 105 yr (Figure 30). A coplanar Neptune mass Fomalhaut b, on the other hand, does not erode the belt inner edge on timescales approaching 106 yr. One reason we stop the simulations before ∼106 yr is that the coplanar geometry leads to a significant planet–planet scattering that alters the orbit of Fomalhaut b and the belt edge erosion timescales have to be reconsidered. For example, some encounters reduce Fomalhaut b's apoastron so that it resides in the belt. With Fomalhaut b spending a greater fraction of time in the belt, a Neptune mass becomes a significant disruptor of pre-existing morphology and the belt becomes a disk in <106 yr.

Figure 29.

Figure 29. Numerical integration of a Fomalhaut system consisting of a nested Jupiter at 120 AU (white circle), a 20 AU wide main belt, and a coplanar, belt-crossing Fomalhaut b (marked with a cross and white ellipse) that has a Neptune mass (left) and a Saturn mass (right). The orbits of the nested and rogue planets are traced by thick and thin white lines, respectively. The rogue Neptune planet has not eroded the inner belt edge 300 kyr after being introduced in a belt-crossing orbit (450 kyr in the simulation). Conversely, after only 75 kyr (225 kyr in the simulation), the rogue, coplanar Saturn has spread the belt radially, eroding both the inner and outer edges. The roque Saturn (right) has a = 285 AU (e = 0.82) due to a close encounter with the nested Jupiter (which also has a modified orbit) just before this snapshot. With the longer orbital period (3500 yr), the Saturn simulation will erode the belt on a 75% longer timescale. When the rogue Saturn has an initial 20° inclination relative to the belt, the belt spreads as in the right panel after ∼500 kyr.

Standard image High-resolution image
Figure 30.

Figure 30. Radial profiles of the model main belt which started as 8000 particles contained in the region 140–160 AU. A Jupiter is placed at a = 120 AU and after 150 kyr the perturbations widen the belt's width (black solid line), except that Jupiter maintains a sharp inner edge. A fraction of particles (6%) cross inward of Jupiter to produce the interbelt disk (e.g., shown here between 60 and 120 AU). We add a rogue, coplanar Neptune (a = 177 AU, e = 0.8 and integrate the belt for another 300 kyr. By 450 kyr (blue dotted line), perturbations from both Neptune and Jupiter spread the belt more, but Jupiter continues to maintain the sharpness of the inner edge. Instead of Neptune, we add a rogue Saturn with the same orbit as the Neptune case. By 225 kyr (75 kyr after it was added into the simulation) the sharpness of the belt inner edge has been significantly eroded.

Standard image High-resolution image

These timescales are likely lower limits because we have been assuming that Fomalhaut b is coplanar with both the main belt and with Fomalhaut c, maximizing the probability of strong interactions. When Fomalhaut b's orbital plane is inclined by 20° relative to Fomalhaut c and the main belt, the belt still spreads radially in the Saturn case, but the timescale is 500 kyr instead of the 75 kyr observed in the coplanar case (Figure 29). Inclining a Neptune-mass Fomalhaut b by 20° decreases the probability of a strong interaction with a nested Jupiter. However, many weak interactions over ∼106 yr timescales will evolve the Neptune orbit. For example, by 6 Myr Fomalhaut b's inclination has increased to ∼39° and eccentricity has decreased to ∼0.5. Kozai-like oscillations between inclination and eccentricity means that the belt tends to be protected from significant interactions with Fomalhaut b: high I, low e excursions mean that Fomalhaut b enters the system only near periastron in the inner regions of the system that miss the belt, whereas low I, high e means that Fomalhaut b passes closer to the main belt, but with a shorter timescale due to the high e.

A more thorough exploration of the orbit and mass parameter space is required and could establish a lower limit to the age of the current orbital configuration. Certainly any model must also test the origin of Fomalhaut b's eccentric orbit, with the possibility that multiple massive planets are located in the system. Future observations can also search for evidence that the belt was disturbed by the previous belt crossing east of the star.

In summary, a planet mass for Fomalhaut b is not excluded by arguments concerning belt disruption because (1) belts are not "wiped out" by a single belt crossing of a Jupiter-mass planet, and we do not know how many belt crossings have already occurred; (2) Saturn masses and below can cross through the belt hundreds of times before the belt edges are eroded; (3) belt crossings erode the sharpness of belt edges, but nested planets may compensate by maintaining the edges sharp; and (4) a mutual inclination of Fomalhaut b relative to the main belt increases belt edge erosion timescales significantly.

9.3.4. Impacts on the Planet

One indirect way to infer the mass of Fomalhaut b is if impacts become evident during a belt crossing. If Fomalhaut b has the mass of a gas giant, then we could expect phenomena similar to the SL9 (D/1993F2) impacts with Jupiter (e.g., Graham et al. 1995; Zahnle & Mac Low 1995; Anic et al. 2007). The greater energies involved would manifest as significant optical and infrared variability, with a different characteristic spectral energy distribution and time dependence than impacts on a dwarf planet. Careful analysis could even yield information concerning its atmosphere and composition (e.g., Bjoraker et al. 1996).

For the SL9 events, Carlson et al. (1997) report ∼0.5 × 1025 erg from the G-impact fireball in a 60 s interval. Temperatures are 8000 K at the beginning, cooling to ∼1000 K after 80 s, giving roughly 8 × 1022 erg s−1 = 8 × 1015 W. If this event were located at the heliocentric distance, D, for Fomalhaut, then

Relative to the luminosity of Fomalhaut A:

Equation (3)

Here we assume that the received flux from Fomalhaut A is f = 8.9 × 10−9 W m−2, and N is some tuning factor, such as N fireballs. For N > 100, Δmag < 24.8 mag and Fomalhaut b would appear significantly brighter. However, the peak of emission quickly (i.e., in seconds) shifts from an optical flash to a relatively long-lived emission at near-infrared wavelengths.

Are N > 100 fireballs plausible for Fomalhaut b's encounter with the belt? In Section 9.3.1 we calculated the accretion rate with gravitational focusing onto a dwarf planet. If we instead assume a Jupiter mass, with Rb = 6.99 × 107 m and vesc = 159 km s−1, then dM/dt = 6.4 × 1010 kg s−1, or about one comet Halley per hour.

This rough calculation suggests that optical flashes may be observable as Fomalhaut b crosses through the belt. The energies involved will help constrain the mass of Fomalhaut b, but its atmosphere will also be heated and excavated. The infrared luminosity would therefore rise and molecular features in a spectrum may become observable and display variability as conditions change on the planet.

9.3.5. Recent Giant Impact as the Origin of the Main Belt

Extending the impact theme even further, is it possible that Fomalhaut b collided with a hypothetical second planet, Fomalhaut c, and the main belt is now the remnant debris of Fomalhaut c? Giant impacts that can produce transient circumstellar dust rings have recently been invoked to explain the properties of several debris disks (e.g., Lisse et al. 2009; Melis et al. 2012). Such a hypothesis is exceedingly unlikely at the large stellocentric distances represented in the Fomalhaut system. On the other hand, the hypothesis could account for several observed properties. The stellocentric offset of the belt is explained as the elliptical orbit of the precursor object, Fomalhaut c. The belt is narrow because it is recently created and has not had time to collisionally evolve and spread radially. Fomalhaut b is belt crossing because the belt would not exist otherwise. While most of the mass in Fomalhaut c is dispersed along its orbit, most of Fomalhaut b's mass is retained in a circumplanetary disk in the process of reaccreting onto the planet or forming moons, but temporarily making it bright in reflected light. The fraction of Fomalhaut b's mass that has been lost comprises a more tenuous stream of co-orbital material that manifests as the interbelt 141° arc of 450 μm emission. In the next section we also study whether or not the main belt gap could be explained by this model.

The critical problem given by Boley et al. (2012) is that the impact speed for collisional erosion has to be significantly greater than the mutual escape speed of the two bodies, but the orbital velocities are small at great distances (150 AU) from the star. Erosive impacts, require vi/vesc ⩾ 1.5 (Asphaug 2009; Marcus et al. 2010; Stewart & Leinhardt 2012; Leinhardt & Stewart 2012), where the impact velocity is $v_i = \sqrt{\vphantom{A^A}\smash{{{ v_{{\rm esc}}^2+v_{{\rm rel}}^2}}}}$.

For a Moon mass, vesc = 2.4 km s−1 and the 45° entry of Fomalhaut b into the belt in the prograde sense gives vrel = 2.7 km s−1. Therefore, vi = 3.6 km s−1, which is a factor of 1.3 greater than vrel. If the collision is in the retrograde sense, then vrel = 6.6 km s−1, and vi/vesc = 2.6.

Therefore the collision of two Moon mass objects would be an erosive event in the retrograde sense. For lower mass objects (e.g., Pluto) the prograde collision becomes erosive (vi/vesc ≈ 3) but the objects do not represent enough mass to account for the main belt mass. For higher mass objects, vesc becomes too large to be consistent with erosive impacts. Therefore the likely mass ranges of the colliding objects are in the Moon regime and a retrograde collision may be necessary.

The timescale for the debris to spread into a circumstellar belt is given by Wyatt & Dent (2002): $\Delta t = 2 \pi / (2 \sqrt{3} v_\infty a)$. For a = 150 AU and the range 10 m s−1 <v < 100 m s−1, we find 104 yr < Δt < 103 yr. Thus the collision may have occurred relatively recently. Since the collision lifetime for 10 μm sized grains is 105 yr (Wyatt & Dent 2002), the giant impact scenario allows stripped material from Fomalhaut c to evolve into a belt on shorter timescales than the collision lifetime of grains.

9.3.6. Origin of the Main Belt 331° Gap

The current snapshot that Fomalhaut b is about to cross through the belt near the 331° gap invokes the idea that the gap is caused by material scattered away from the belt when the planet crosses through. In the previous sections we argued that this is not true in the case of Fomalhaut, though other astrophysical disks that are thin, gaseous, self-gravitating, and/or shadowed by optically thick material closer to the star may display such morphology (e.g., the azimuthal gap in the circumbinary belt surrounding the pre-main sequence system GG Tau; Roddier et al. 1996; Krist et al. 2005).

One possibility for explaining Fomalhaut's main belt gap is that a planet orbits within the belt, and the gap represents tadpole or horseshoe orbits of the co-orbital planetesimals and dust. The analogy is to the dynamics of the giant planet Trojan populations observed in our solar system (Chiang & Lithwick 2005; Nesvorný & Vokrouhlický 2009; Lykawka et al. 2011). The Earth is also known to trap in-spiralling dust grains near a 1:1 resonance, producing a ∼1 AU radius dust ring orbiting the Sun, except near the planet, where grain dynamical lifetimes are short (Jackson & Zook 1992; Dermott et al. 1994).

A second possibility is that the gap is related to the giant impact scenario (Section 9.3.5). For ∼103 orbital periods all particles created at the collision origin point return back to the collision origin point. A snapshot of debris particle location reveals a gradual pinching of the belt toward and away from the collision origin point (Jackson & Wyatt 2012). This tapered morphology resembles the main belt gap. Fomalhaut b would return to the collision origin point if a grazing, hit-and-run collision ejects mantle material, but the planet core stays close to the pre-encounter orbit. Therefore, one of Fomalhaut b's two apparent belt crossing per orbital period should occur near this point, which should coincide with the radially tapered section of the belt. The recent, giant impact thereby ties together the apparent proximity of Fomalhaut b to the main belt gap to the north of it.

Contradicting this scenario is that even though the orbital paths for debris tapers toward and away from the collision origin point, the quantity of dust is not changing. This scenario therefore does not account for diminishing the scattered light in the belt gap region, though more complex effects mentioned below may come into play. Another significant problem is that the belt pinch requires a low velocity dispersion of debris (100–150 m s−1), which counters the relatively high velocities needed to disrupt the required precursor mass (A. Jackson 2012, private communication).

One general solution is to suppose that the giant impact occurred recently (<104 yr) and the debris is still spreading along the orbit of the precursor object—more time is required to create an azimuthally uniform ring. We cannot rule this out, and certainly future observations should search for other structure in the belt consistent with a debris field that is dynamically young. A second solution is that the gap arises from a combination of a relatively young ring (∼104 yr) and a period commensurability between the planet and the debris (see Figure 6 in Jackson & Wyatt 2012).

We propose a third possibility concerning a geometrical effect that results if apsidal alignment is accompanied by nodal alignment, presumably because of dynamical interactions with a low-eccentricity planet. If a large ensemble of belt particles have both non-zero orbital inclinations and nodal alignment, the belt pinches vertically toward the midplane at the ascending and descending nodes. Figure 31 demonstrates this hypothetical configuration. We note that the 331° is roughly 180° away from our estimate for the ascending node. In other words, the belt gap is near the descending node where one of the two pinch areas occurs. We find that due to projection effects, the pinch area is obscured at the ascending node due to foreground and background material contained in the line of sight to the ascending node. The orbital configuration of particles is similar to the giant impact scenario, except that there are two pinch points.

Figure 31.

Figure 31. Examination of projected belt geometry when a belt has some vertical thickness and all of the belt particles are apsidally and nodally aligned. Here we have taken our N-body model for the main belt and exagerrated the vertical thickness by a factor of five to emphasize the projected morphologies. In the near edge-on view (left panel) the apsidal+nodal alignment produces vertical pinching. If we were to rotate our viewpoint 90° to the right or left, the pinch points would not be evident because they are projected within the ansae of the belt. The right panel shows the same model belt rotated according to the geometric elements that we derive for the belt. The ascending node is to the lower left but the descending node to the upper right has a tapered morphology as seen in projection.

Standard image High-resolution image

As with the giant impact debris field, the problem with this scenario is that material is confined, but not necessarily removed from the pinch area. Therefore the surface brightness should not diminish significantly. On the other hand, the material at a vertical pinch point has a very flat spatial distribution, so that belt particles are more likely to be self-shadowing. Self-shadowing is invoked to explain why Uranus' epsilon ring is fainter at periapse than at apoapse—the ring optical thickness increases at periapse and self-shadows (Karkoschka 1997).

Future work is needed to quantitatively study the cumulative effect of these factors on the scattered light appearance of debris belts near pinchpoints. The theory of apsidally and nodally locked planetary rings depends on the interplay between collisions, self-gravity, and the quadropole field of the planet (Chiang & Culter 2003). For a debris belt, collisions are important, but self-gravity is assumed to be insignificant, and planet–belt dynamics include significant secular effects. Observationally, the origin of Fomalhaut's 331° main belt gap could be explored by ALMA observations. Self-shadowing would be irrelevant at mm wavelengths such that a belt gap in a mm map would indicate an absence of grains, thereby supporting the horseshoe/tadpole orbit hypothesis.

9.4. Comparison to the Solar System

Figure 32 plots the semi-major axes and eccentricities of classical Kuiper Belt Objects (KBOs), scattered Kuiper Belt Objects (sKBOs), and Centaurs cataloged by the Minor Planet Center. Fomalhaut b's orbit lies in a region of ae parameter space occupied by sKBOs (Luu et al. 1997; Levison & Duncan 1997; Trujillo et al. 2000; Brown et al. 2005). A key reason for the overlap is that the perihelia of KBOs do not cross inward of Neptune's 30 AU semi-major axis (producing the upward curved boundary on the right side of the cluster of points), and Fomalhaut b's periastron also happens to be near 30 AU. Therefore, the plot merely emphasizes that in terms of eccentricity, Fomalhaut b is in the domain of the scattered Kuiper Belt instead of the classical Kuiper Belt. However, because Fomalhaut b's periastron is significantly smaller than particles in Fomalhaut's main belt, a more apt comparison is to a few known Centaurs with high eccentricity that cross Neptune's orbit.

Figure 32.

Figure 32. Distribution of a vs. e for classical Kuiper Belt Objects (red diamonds), scattered Kuiper Belt Objects and Centaurs (blue crosses), and Fomalhaut b (blue shaded region).

Standard image High-resolution image

One example of such a Centaur is 2001 XA255 with a = 30 AU, e = 0.7, i = 13° (Jewitt et al. 2002; de la Fuente Marcos & de la Fuente Marcos 2012). This object crosses the orbits of three planets (Saturn, Uranus and Neptune) and the dynamical evolution is chaotic and short (<108 yr; Dones et al. 1996; di Sisto & Brunini 2007; Bailey & Malhotra 2009).

This suggests that as an alternative to planet–planet scattering scenarios, the dynamical mechanisms that produced the sKBOs and Centaurs may be active in the Fomalhaut system. We calculate the Tisserand parameter of Fomalhaut b relative to a hypothetical, significantly more massive Fomalhaut c that serves as the perturber of the main belt inner edge (ac = 120 AU), or a hypothetical Fomalhaut d located near Fomalhaut b's periastron (ad = 30 AU). For example, in the case of an interaction with Fomalhaut c, the Tisserand equation is

Equation (4)

The mutual inclination of the two objects is Ibc. If ac = ab, eb = 0, and Ibc = 0°, then T = 3. The Tisserand parameter is approximately conserved for dynamical interactions in the restricted, three-body problem. Therefore, instead of classifying objects in terms of their present-epoch orbital parameters, the Tisserand parameter is a more useful standard because it tends to be invariant over the many dynamical interactions with planets that vary a minor body's orbital parameters over time (Levison & Duncan 1997).

In principle, slow, strong encounters have T ≲ 3 and the perturbed object is dynamically coupled to the massive planet. An Oort cloud comet with high mutual inclination has T < 2. Objects in the solar system with 2 < T < 3 are dynamically coupled to the planets. If T ≳ 3, then the object is not dynamically coupled to the massive planet.

In the coplanar case Tbd = 3.0 ± 0.1 and Tbc = 2.1 ± 0.3 (Figure 33). Therefore Fomalhaut b may be dynamically coupled to both planets in the coplanar case. If we take Ibc = 36° as the maximum value for mutual inclination (Figure 22), then Fomalhaut d can have ad as small as 20 AU for T = 3.0. If the object that scattered Fomalhaut b resides within the main belt (e.g., a planet responsible for the 331° gap), then for ac = 140 AU, Tb = 2.1 ± 0.3.

Figure 33.

Figure 33. Tisserand parameter for Fomalhaut b assuming that it was scattered by a Fomalhaut c that has ad = 30 AU (blue histogram) or ac = 120 AU (red histogram).

Standard image High-resolution image

Another potential comparison to the Kuiper Belt concerns mean motion resonances. If ab = 177 AU, then Pb = 1700 yr, whereas the main belt at a = 133, 143, 153 AU has P = 1106, 1234 and 1366 yr, respectively. Therefore a 3:2 resonance is apparent between Fomalhaut b and the inner edge of the belt at 133 AU. This is probably physically irrelevant given that Fomalhaut's main belt has significant width, and Fomalhaut b's semi-major axis is sufficiently uncertain, that at least one mean motion resonance can be identified by chance. Moreover, if Fomalhaut b's orbit is a relatively recent outcome from a planet–planet scattering, then the important resonances over the age of the system are those that Fomalhaut b had before it was scattered.

Given the similarity of Fomalhaut b to the dynamics of our solar system's scattered disk KBOs and Centaurs, is Fomalhaut b better described as an extrasolar dwarf planet rather than an extrasolar planet? The main observational constraints aside from the astrometry that gives the orbit, is the optical brightness that could be reflection from material with a large surface area. Therefore the analogy to a KBO is possible if we invoke Pluto, during the brief (∼100 yr) epoch when its moons were being assembled from a circum-Pluto disk, or the Haumea family of KBOs (Brown et al. 2007), at the epoch when a significant collision disrupted the precursor object. Both scenarios invoke a "giant impact" between dwarf planets, which we considered in Section 9.3.5. The Charon-forming giant impact accounts for the small mass ratio between Pluto and Charon (Canup 2005) and the formation of Nix and Hydra (Canup 2011). The simulations show that low relative velocities (∼1 km s−1) can form the Pluto system if the collisions are oblique and the precursor object is partially differentiated and has an ice shell. However, other collision scenarios do not necessarily produce moons but yield circumplanetary disks of material with mass ∼1020 kg. The Haumea collisional family currently consists of one dozen members with moderate eccentricity (0.1–0.2), a relatively large inclination (24°–29°) and a semi-major axes near 43 AU (Lykawka et al. 2012). The various collision models (Volk & Malhotra 2012; Schlichting & Sari 2009; Leinhardt et al. 2010) are broadly similar to the discussion in Section 9.3.5.

Given the plausibility of these hypothetical scenarios that address observations of our own solar system, it is difficult to rule out the hypothesis that Fomalhaut b is an extrasolar dwarf planet.

Finally, we briefly consider if the solar system could presently contain an object like Fomalhaut b. The geometric surface area of grains representing Fomalhaut b, σp = 1.2 × 1019 m2, is ∼6 × 103 greater than the projected surface area of Neptune. Neptune's apparent visual magnitude is +7.8 mag, which means that something with a factor of 6000 times greater surface area and the same albedo as Neptune would be one of the brightest objects in the night sky at −1.6 mag. Consider now a factor of ∼10 increase in heliocentric distance corresponding to Fomalhaut b's apastron at ∼300 AU. Fomalhaut b is fainter by a factor of 104 as viewed from Earth, giving it an apparent magnitude of +8.4. Clearly no clone of Fomalhaut b exists at the current epoch in our solar system.

10. SUMMARY: IS IT A PLANET?

Our finding of a likely periastron passage near 30 AU radius now confers to Fomalhaut b a direct physical connection to the region where planetesimals grow to planets because the dynamical timescales are shorter and the primordial disk is denser closer to the star. On the other hand, compared to the present-day dynamics of the solar system, the orbit of Fomalhaut b is similar to Centaurs in the solar system. How do we distinguish between a dwarf planet and a planet in the case of Fomalhaut b?

One a priori argument against a planet mass for Fomalhaut b is that by crossing through the belt, it would dynamically disrupt the belt. We give several important reasons why this argument is not definitive.

  • 1.  
    We find that the mutual inclination between Fomalhaut b and the main belt is 17° ± 12°, with ∼10% of possible ascending and descending nodes crossing through the main belt. Therefore it is unlikely, but not ruled out, that Fomalhaut b crosses through the belt at the present epoch, though belt crossing may happen at other epochs due to orbital evolution. For example, our initial N-body tests indicate that Fomalhaut b may evolve into a significantly different configuration on 106–108 yr timescales if a nested, Jupiter-mass planet orbits within the belt perimeter.
  • 2.  
    We present N-body simulations that show a planet crossing through a belt does not destroy it, but instead erodes the edges of a belt on timescales that depend on the assumed mass of the planet (e.g., a Saturn mass requires ∼105 yr or 102 crossings). Since we do not know how long the present orbital geometry has existed in the Fomalhaut system, Fomalhaut b could be a gas giant planet.
  • 3.  
    A corollary to (2) is that due to Fomalhaut b's large eccentricity, it passes through the belt quickly (∼100 yr) in the coplanar case. Therefore the effectiveness of a belt-crossing planet in modifying belt dynamics is diminished.
  • 4.  
    We show two new features in the main belt that in fact suggest the dynamics of the system are more complex than previously established. First we identify a ∼50 AU wide gap in the azimuthal structure of the belt north of Fomalhaut b. Second, the outer edge of the belt is extended to at least 209 AU, and appears warped beyond this radius.

To summarize, the potential belt passage of Fomalhaut b does not exclude any masses up to the 1 MJ mass limit determined by infrared imaging surveys. We therefore considered several aspects of a lower mass limit, establishing the following principles.

  • 5.  
    Assuming the observed optical light from Fomalhaut b is reflected from dust grains, the mass of the required grains implies a precursor object >10 km in size. The collision physics of objects this large lies in the gravity dominated regime, which means that Fomalhaut b consists of a central planetesimal or family of planetesimals surrounded by a bound dust cloud of greater extent and surface area.
  • 6.  
    If the mass of the central planetesimal is too small, a dust cloud surrounding it is sheared away during periastron passages. We show that for the planetsimal cloud to be stable against shearing, the minimum mass of Fomalhaut b is 5 × 1021 kg, comparable to 5× Ceres' mass.

Fomalhaut b's mass is therefore in the range between our solar system's dwarf planets and Jupiter. Unless new dynamical simulations can show otherwise, the main belt's inner edge and stellocentric offset are not definitively linked to Fomalhaut b alone. However, our current orbit determination shows that Fomalhaut b is apsidally aligned with the main belt.

To explain all of the various observed features in the system, additional, low-eccentricity "nested" planets may be necessary. A comprehensive analytic and numerical study of the possible parameter space is required. We considered the possibilities that Fomalhaut b was scattered from a nested planet located near its periastron at ∼30 AU, that a planet at 120 AU sculpts the inner edge of the main belt, and that a planet orbiting within the main belt could account for the azimuthal gap. Current direct imaging studies of the system are inadequate to exclude any planets less than a Jupiter mass in these regions.

Several additional points are:

  • 7.  
    Fomalhaut b's orbital parameters are similar to scattered KBOs and Centaurs, which suggests the dynamical mechanisms operating in the Fomalhaut system could further our knowledge of the early solar system's dynamical history.
  • 8.  
    For a coplanar orbit, Fomalhaut b will collide with the main belt two decades from now. Monitoring for transient phenomena similar to SL-9 impacts on Jupiter would elucidate the mass and composition of Fomalhaut b, and perhaps lead to unique insights of exoplanet atmospheres as main belt planetesimals excavate and heat the atmosphere.
  • 9.  
    We also considered an unlikely scenario where the main belt is the partial remnant of a Fomalhaut c that suffered a recent collision with Fomalhaut b. Due to the low relative velocities at ∼140 AU, and the requisite mass of material in the belt, this scenario is more credible for a head on collision (the belt has a retrograde orbit relative to Fomalhaut b).
  • 10.  
    We speculate that if Fomalhaut b has a satellite system that was dynamically disturbed by the recent periastron passage, then the moons may also be tidally heated by the central planet. By analogy to Io and Jupiter, enhanced volcanic activity could lead to large sodium or potassium clouds around Fomalhaut b that would explain puzzling aspects of the optical data.
  • 11.  
    We studied the possibility that the star Fomalhaut B (TW PsA) may also have a role in the dynamics of the system, but little is known about its orbit except that the period is likely >8 Myr.

Future observations and theoretical investigations can therefore address several important open questions. (1) Does the orbit of Fomalhaut b pass through the belt? (2) Is the spectrum of Fomalhaut b consistent with reflected light, and are there any features indicating composition? (3) What is the interconnection between the apsidal alignment of Fomalhaut b with the main belt and the azimuthal belt gap? (4) Where is Fomalhaut c?

Based on observations with the NASA/ESA Hubble Space Telescope, obtained at STScI, which is operated by AURA under NASA contract NAS5-26555. This work received support from the following: GO-11818 and GO-12576 provided by NASA through a grant from STScI under NASA contract NAS5-26555; NSF AST-0909188; and the University of California LFRP-118057. We are grateful to G. Soutchkova, J. Debes, G. Schneider, J. Duval, A. Aloisi and C. Proffit for assistance in the observations. We thank H. Beust for sharing his insights on the Markov chain orbital element analysis, and A. Jackson, G. Kennedy, H. Levison and M. Wyatt for reviewing the draft manuscript. We also thank E. Chiang, T. Currie, R. Dawson, G. Duchene, E. Ford, B. Hansen, M. Hughes, C. Lisse, R. Malhotra, D. Nesvorny, S. Stewart, D. Tamayo for helpful discussions

Facility: HST (STIS) - Hubble Space Telescope satellite

Footnotes

  • Our analysis of the background star in the lower frame of the 2012 STIS detection (Section 5) shows that its point-spread function is broader than that generated by TinyTim (Krist et al. 2011). We use the image of this source to facilitate estimation of the fraction of energy encircled in the small 2.4 pixel diameter aperture used in both STIS datasets.

Please wait… references are loading.
10.1088/0004-637X/775/1/56