Open Access
4 June 2021 Starshade Rendezvous: exoplanet orbit constraints from multi-epoch direct imaging
Andrew Romero-Wolf, Geoffrey Bryden, Greg Agnes, Jonathan W. Arenberg, Samuel Case Bradford, Simone D'Amico, John Debes, Matt Greenhouse, Renyu Hu, Steve Matousek, Jason Rhodes, John Ziemer
Author Affiliations +
Abstract

The addition of an external starshade to the Nancy Grace Roman Space Telescope will enable the direct imaging of Earth-radius planets orbiting at ∼1 AU. Classification of any detected planets as Earth-like requires both spectroscopy to characterize their atmospheres and multi-epoch imaging to trace their orbits. We consider here the ability of the Starshade Rendezvous Probe to constrain the orbits of directly imaged Earth-like planets. The target list for this proposed mission consists of the 16 nearby stars best suited for direct imaging, around which ∼10 to 15 planets are expected to be discovered. Of these planets, ∼1 to 2 will be Earth-like in mass and temperature. The field of regard for the starshade mission is constrained by solar exclusion angles, resulting in four observing windows during a two-year mission. We find that for Earth-like planets that are detected at least three times during the four viewing opportunities, their semi-major axes are measured with a median precision of 7 mas, or a median fractional precision of 3%. Habitable-zone planets can be correctly identified as such 96.7% of the time, with a false positive rate of 2.8%. If a more conservative criteria are used for habitable-zone classification (95% probability), the false positive rate drops close to zero, but with only 81% of the truly Earth-like planets correctly classified as residing in the habitable zone.

1.

Introduction

The Starshade Rendezvous probe (SRP) mission concept proposes adding a Starshade to the Nancy Grace Roman Space Telescope enabling the detection of habitable zone exoplanets and characterization of their atmospheres.1 The detailed technical basis for the SRP study report1 was presented in Ref. 2 along with the simulations used to estimate sensitivity of the observatory. The corresponding software is publicly available for reproduction of the results below and for comparison with similar simulations ( https://github.com/afromero/Starshade_Rendezvous_Probe_sims). The main result of these studies is that SRP is capable of discovering Earth-size planets in the habitable zones of nearby stars using the relatively moderate aperture of the Roman space telescope2 along with the Coronagraph Instrument.

While the SRP science objectives include quantifying the amount of habitable zone dust around nearby stars and measuring the metallicity of known gas giant planets, the primary driver is the detection and characterization of Earth-like planets. The overall strategy, described in more detail in Ref. 2, involves three main steps: (1) initial detection via direct imaging, (2) habitable zone determination via orbit tracing, and (3) atmosphere characterization via spectroscopy. The integration times necessary to image and to take spectra of Earth-like planets (steps 1 and 3) were taken into account with a model of the observatory. However, step 2 is more complicated since the observatory field of regard is constrained by solar exclusion angles, typically limiting the target observing windows to two 30 day blocks per year—a total of four observing opportunities during the assumed 2-year mission lifetime. Depending on the orientation and phase of a planet’s orbit, it may or may not be visible during each of these four observing windows. In Ref. 2, we assumed that detecting the planet during at least three of the four epochs would be sufficient to determine if a planet lies within its parent star’s habitable zone. In this paper, we consider this step in more detail, performing multi-epoch orbit fitting for the target list and expected signal-to-noise ratio (SNR) given by the observatory model used in Ref. 2.

Measurement of planetary orbits has been previously modeled for several types of observations—radial velocity, astrometric wobble, and coronagraphic direct imaging. Examples include: (1) combining radial velocity measurements with direct imaging upper limits to improve the orbital fit for the planet eps Eri b,3 (2) quantifying the robustness of planetary orbit determination via the stellar astrometric signal, concentrating on the difficulty posed by multi-planet systems,4 and (3) considering simultaneous stellar astrometry and direct imaging, finding that Earth-like planets can be characterized in just a few observations.5 Among the studies that, such as this paper, concentrate on direct imaging, Ref. 6 performed a detailed analysis of orbital constraints based on the small fraction of an orbit that is traced by known long-period direct-imaged planets. For theoretical cases where the observations span at least half an orbital period, Ref. 7 finds that three equally spaced observations with SNR10 can measure the semi-major axis and eccentricity to 10%. Reference 8 also considered direct imaging of shorter period planets, finding that a habitable zone planet’s semi-major axis can be measured to within 5% if it is observed with precision 3.5 mas over three epochs each spaced by at least 90 days apart.

Here, we consider the results obtainable by a specific mission concept—the SRP. Unlike previous work, this includes: (1) a realistic SNR calculation as a function of stellar illumination, rather than an assumed astrometric precision, (2) a starshade-specific inner working angle (IWA) that obscures planet images close to the star, and (3) observing windows based on known target sky positions and observatory pointing constraints, not just arbitrarily spaced observations. Furthermore, we focus here not on general orbit fitting results, but on a specific science question—whether or not we can determine if a planet lies in its parent star’s habitable zone.

In the following, we first summarize the integration time-based observatory model from Ref. 2 used to calculate the SNR for each planet image Sec. 2. For many sets of simulated observations, we apply a parametric SNR-based estimate of astrometric precision to then extract orbital elements for each injected planets Sec. 3. We give the resulting precision for the orbital fits in Sec. 4 and summarize in Sec. 6.

2.

Observing Model

We briefly summarize the models used for planet properties, orbit propagation, and the observatory. A detailed presentation of these models can be found in Ref. 2.

2.1.

Targets

Planet sizes and orbital periods are drawn randomly from these defined ranges for Earth-like planets, based on the distribution defined by 13th NASA Exoplanet Program’s Study Analysis Group (SAG-13)9 and modified by the HabEx mission concept10 to include the dependence of the orbital semi-major axis on the lower limit of planet radii. The orbital period P defines the orbital radius ap by way of the stellar mass M using Kepler’s third law. For sampling of Earth-like exoplanets, the orbits are assumed to be circular, consistent with most previous studies, e.g., Ref. 11. However, when fitting, we allow eccentricity to be a free parameter (see Sec. 3).

The orbital semi-major axes are sampled over a range from inside the inner habitable zone (defined as 0.95L AU, where L is the stellar luminosity relative to the Sun) to outside the outer habitable zone (defined as 1.67L AU ; Ref. 12). These orbital radii correspond to planet equilibrium temperatures from 261 to 197 K, for an assumed geometric albedo of 0.2.

The range of planet radii considered is bounded above at rpl1.4 r, based on evidence suggesting that planets with radii below this value are predominantly rocky.13 The lower bound on terrestrial planet radii depends on the planet’s ability to retain an appreciable atmosphere, which in turn depends on its stellar illumination. This results in a dependence on the planet’s semi-major axis ap, modified by the stellar luminosity, to give14 rpl/r0.8ap1/2L1/4.

We model Earth-like exoplanets assuming they scatter light isotropically using a Lambertian illumination phase function with a geometric albedo of 0.2 (Ref. 15). We model the star as a blackbody radiator with the parameters provided in ExoCat.16 We include obscuring dust, both in the target system (exozodiacal dust with a fiducial value of 4.5 zodi17) and locally in the Solar System (zodiacal dust using the model of Ref. 18). While some systems may have exozodiacal dust above this nominal level (e.g., Fomalhaut), the mission’s observing plan includes a decision tree to efficiently remove systems from the target list, should their observed dust level preclude Earth-like planet imaging.2 Assumed planet and dust characteristics are summarized in Table 1.

Table 1

Planet and dust parameters.

ParameterValue [or(range)]
Earth-like planet geometric albedoa0.2
Earth-like planet radius[0.8ap1/2L1/41.4]r
Habitable zone[0.951.67]L AU
Zodiacal dust brightnessb23.3  mag/arcsec2
Exozodi dust brightnessc4.5 zodi

aFor the assumed isotropic scattering, this geometric albedo is equivalent to 0.3 spherical albedo.

bThe zodiacal light varies with position on sky. The given value corresponds to the minimum at the ecliptic poles (Ref. 18).

cThe unit of 1 zodi is equivalent to 22 mag/arcsec2.

While the use of an occulting starshade does allow for detection of fainter and closer planets, it also puts constraints on the allowed times of observation. The telescope-starshade system has a region of allowed Sun angles over which it can operate with the lower limit defined by the exclusion angle from the baffle of the telescope and the outer limit defined by reflection and scattering of sunlight off the starshade into the telescope baffle (Table 2).

Table 2

Mission parameters.

ParameterAssumed performance
Mission lifetime2 years
Telescope primary mirror2.4 m
Imaging resolution65 mas at 750 nm
Imaging bandpass615 to 800 nm
Imaging end-to-end efficiency0.035
Solar exclusion angle (min)54 deg
Solar exclusion angle (max)83 deg
IWA100 mas
Instrument contrast4×1011
Imaging integration time1 to 6 days

2.2.

Observatory

The integration times are based on the Starshade/ Roman system parameters provided in Table 2. Roman has a telescope diameter of -m resulting in a point spread function of 65 mas at 750-nm wavelength. We assume observations in the 615- to 800-nm band with an end-to-end efficiency, including optical throughput and detector efficiency, of 3.5% in imaging mode. The Starshade has an IWA of 100 mas, below which planets are assumed to not be observable. This is treated as a sharp cutoff, although in reality there is some throughput below the IWA.19 The instrument contrast at the IWA and above is assumed to be 4×1011, as calculated by Ref. 1. Integration times vary between 1 and 6.3 days, depending on the target (explained below).

In addition to the sensitivity, determined by the parameters given above, the main constraints to orbit reconstruction are the nominal mission lifetime of 2 years and the solar exclusion angles. The minimum solar exclusion angle of 54 deg is determined by the telescope baffle while the maximum solar exclusion angle of 83 deg is determined by scattering off the edge of the Starshade. For Roman’s orbit around the L2 Sun-Earth Lagrange point, the calculated observing windows are shown in Fig. 1. During the 2-year lifetime of the mission, there are generally four opportunities to observe each target. It is assumed that targets with Earth-like planet candidates, following the decision tree laid out in Ref. 2, will be visited once in each of the 4 observing windows available. While spectral characterization can be performed with only a single visit with favorable illumination phase, multiple epochs are needed to constrain the planet’s orbit, in particular its semi-major axis, which determines whether the planet is in the habitable zone. To best trace the orbit, the timing of the observations should be as evenly spaced as possible over the orbital period, given the limited observing windows and mission lifetime. The dots in the observing windows in Fig. 1 show the observations times assumed in this study. The retargeting strategy in Ref. 2 assumed that observations can be made within a number of days of the marked dots that would allow for a spectral measurement, which varies from target to target depending on the expected Earth-like exoplanet brightness and observing window length (typically 10 days). It is worth noting that not all observations are required as targets are re-prioritized based on observations. See Ref. 2 for more details.

Fig. 1

Target star observing windows (reproduced from Ref. 2) resulting from the telescope and starshade solar exclusion angles. Each target typically has two 30-day-long observing windows per year. Targets at high ecliptic latitude can have longer observing windows per year. The black dots mark the four desired observation start times in a two-year period. This is driven by the need to allow for sufficient time to spectrally characterize a planet if it is bright enough.

JATIS_7_2_021219_f001.png

The properties of the target stars are shown in Table 3. In Ref. 2, we defined the single-visit completeness as the probability that an Earth-like planet would be detected during one target observation at a random time. The imaging integration time for each target is set by identifying what is required to reach a single-visit completeness of at least 50% with a minimum value of 1 day and a maximum value of 6.3 days (see Table 3). We also defined the orbital completeness as the probability that a randomly selected orbit will be detectable (S/N 7) during at least 3 of its 4 observing epochs. For most of the targets about half of the planet orbits meet this criterion, but the orbital completeness can sometimes fall below 20% for systems where the planet signal is relatively weak (again, see Table 3). While the orbital completeness is limited by the orbit geometry (sometimes a planet is only illuminated as a crescent phase and sometimes its projected location is too close to its parent star), it is possible to increase the completeness somewhat with longer integration times for imaging. However, this would limit the time available for characterizing the planet’s spectrum, particularly when the observing window is only 30 days long. With the nominal integration times, about half of the detected Earth-like planets will have sufficient SNR spectra to characterize their atmospheres.2

Table 3

Astrometric precision for Earth-like planets.

Habitable zoneAstrom. precisionaHabitable zone classification
Star nameDistance (pc)V (mag)L⋆ (L⊙)M⋆ (M⊙)ap (mas)Period (years)Int. Time (days)Completenessper epoch (mas)ap (mas)50% Threshold95% Threshold
Single-visitOrbitalbFalse Pos.False Neg.False Pos.False Neg.
tau Cetic,d3.73.50.520.80187 to 3290.64 to 1.51.00.670.553.25.72.2%2.0%0.0%12.6%
Procyone3.50.47.11.49722 to 12703.3 to 7.71.00.650.543.228.22.2%1.5%0.0%14.1%
eps Indd,e3.64.70.230.68124 to 2190.36 to 0.842.50.670.542.33.51.7%0.5%0.0%7.7%
Siriuse2.6−1.430.52.401993 to 35037.6 to 17.81.00.580.533.972.53.5%2.1%0.0%16.9%
omi 2 Erid5.04.40.420.81124 to 2180.54 to 1.36.30.650.522.54.70.8%1.5%0.2%9.1%
Altair5.10.810.71.83605 to 10644.0 to 9.42.50.580.513.936.33.5%3.9%0.0%24.3%
del Pav6.13.51.30.99179 to 3151.2 to 2.76.30.640.553.54.92.2%2.2%0.0%12.7%
Erid6.04.30.690.85130 to 2290.75 to 1.76.30.600.403.14.41.2%2.4%0.2%11.1%
sig Dra5.84.70.440.80109 to 1810.56 to 1.36.30.550.423.14.51.7%1.1%0.0%7.2%
bet Hyi7.52.83.71.14245 to 4302.3 to 5.46.30.580.403.820.94.6%5.2%0.2%31.7%
bet CVne8.44.21.31.03126 to 2221.1 to 2.56.30.430.144.44.42.2%2.3%0.0%10.5%
Ori8.13.23.01.24203 to 3581.9 to 4.46.30.500.314.99.34.1%5.5%0.0%21.6%
Fomalhautc,e7.71.216.52.05500 to 8795.3 to 12.36.30.460.444.950.54.9%4.0%0.0%40.7%
del Eri9.03.53.41.19193 to 3392.1 to 4.96.30.460.265.311.65.7%6.0%0.2%30.2%
gam Lep8.93.62.51.27168 to 2961.6 to 3.86.30.440.215.98.91.9%7.2%0.0%30.1%
zet Tuc8.64.21.31.01127 to 2241.1 to 2.76.30.420.165.44.92.1%5.4%0.0%24.9%

aAstrometric precisions are medians of many sampled planetary orbits

bOrbital completeness requires three detections with S/N≥7

cKnown debris disk

dKnown to have planet (s)

eBinary

In the next section, we calculate whether at least three detections in four visits (our definition of orbital completeness) is sufficient to determine whether a planet lies within its habitable zone.

3.

Orbit Reconstruction

Having identified the best targets for detection of Earth-like planets with their observation availability windows, we now describe our approach to orbit reconstruction. We assume the planet is observed at the beginning of each window as shown in Fig. 1, which provides four observing epochs per target in most cases. In cases where the star has a single long availability window per year, we have set the observing times to the beginning and middle of that window.

For each Monte Carlo sampled planet, we propagate its circular orbit to each of the observing epochs and calculate its illumination phase (see Sec. 2). We apply the observatory model to estimate the planet SNR. Observations with SNR7 are considered to be detections. Otherwise, the observation is rejected as a non-detection. The one-dimensional astrometric uncertainty is approximated according to Ref. 20 δθ=(65  mas)/SNR. The median astrometric precision for each star ranges from 2.3 to 5.9 mas (see Table 3). The simulated data are created by taking the true position of the planet and adding two-dimensional Gaussian scatter based on the astrometric precision, δθ.

For the orbit reconstruction, we implemented a forward modeling of Kepler’s laws, as described in Ref. 21, into our own software package. We sample all six Keplerian parameters, also including uncertainties in the star’s mass and distance, treating them as nuisance parameters. Table 4 lists the parameters that we fit for each orbit, along with their assumed ranges and prior constraints. We use the emcee Markov chain Monte Carlo (MCMC) software package22 to fit the orbit. The MCMC fitting procedure calculates the quality of fit for a series of parameter values, not only converging toward the best set of values but also finding the full parameter ranges that are consistent with the data. Periodic orbital elements (ω, Ω, T0, and i) are modulated to stay within their prescribed bounds.

Table 4

Orbit fitting parameters.

ParameterDescriptionBoundsPrior constraint
aSemi-major axis0 to 10 AUUniform
eEccentricity0 to 1Uniform
ωArgument of periastron0 deg to 360 degUniform
iInclinationa0 deg to 90 degsin(i)
ΩLongitude of the ascending node0 deg to 360 degUniform
T0Periastron phase0 deg to 360 degUniform
MStellar mass0 to 5 MObserved value with 10% uncertainty
dDistance0 to 20 pcObserved value with 1% uncertainty

ai=90  deg corresponds to edge-on.

Three examples of the orbit reconstruction simulations are shown in Fig. 2. The first panel shows Procyon, a relatively luminous star (7.1 L), meaning that its habitable zone is relatively distant both in angular scale (0.7″ – 1.3″) and in physical space (2.5 to 4.5 AU). With a mass of 1.5 M, planets in the habitable zone have relatively long periods. The randomly selected planet orbiting Procyon is detected in all four observations, but because of the long period, only a fraction of the orbit is traced. The second panel shows a planet orbiting tau Ceti that is only detected in three of the four observing epochs; during the first observation, the planet falls behind the starshade mask (the gray circle in the center of each panel), a common occurrence for planets on inclined orbits. In the third panel (sigma Dra), there are again only three successful observations, but in this case the planet is too faint to be detected during the first epoch due to unfavorable illumination phase. Also, because sigma Dra is near the ecliptic north pole, it has only one observing window per year, although of significantly longer duration (see Fig. 1), and its orbital phase coverage is limited (epoch pairs 1/2 and 3/4 are within the same window). With only three closely spaced epochs, the fit is relatively poorly constrained.

Fig. 2

Example fits are shown for three different stars—Procyon, tau Ceti, and sigma Dra. In each case, the planet’s semi-major axis lies within the habitable zone (dashed lines). The true orbit is shown in orange, with true positions marked as orange circles and simulated observations shown with error bars shown in red only if detected. The numbers indicate the visit number for each of the four observations. Sample best-fit orbits are shown as thin black lines. The region masked by the starshade is shown as a gray circle.

JATIS_7_2_021219_f002.png

An example of the MCMC posterior distributions is shown in Fig. 3. For this inclined orbit, the inclination (i) and longitude of ascending node (Ω) are well determined and are accurately retrieved. The retrieved eccentricity (e) is necessarily larger than the assumed circular orbit, but is still close to zero (0.03±0.02). Given the circular orbit, the true argument of periastron (ω) is undefined and the retrieved value is only loosely constrained. Most importantly, the semi-major axis (a) is well constrained by the observations, enabling us to determine that the planet lies well within the habitable zone.

Fig. 3

Best-fit parameters for a random planet orbiting tau Ceti, showing the marginalized probability distributions for 3 of the 8 model parameters (top histograms) and the correlations between pairs of parameters (central panels, with 1-, 2-, and 3σ contours). The retrieved parameters are consistent with the true values (blue lines), although the fit eccentricity is necessarily larger than the true circular orbit’s. For this example, the planet lies unambiguously inside the habitable zone; the distribution of semi-major axes falls entirely within tau Ceti’s 187 to 329 mas range.

JATIS_7_2_021219_f003.png

4.

Results

For each of the 16 target stars listed in Table 3, we simulate 1000 random orbits and then extract orbital parameters as described above. Table 5 lists the number of orbital calculations for the full set of simulations (see Ref. 22 for details on MCMC parameters).

Table 5

Simulation parameters.

ParameterQuantity
Target stars16
Random orbits per star1000
MCMC iterations per orbit5000
MCMC walkers100

Our first objective is to accurately determine each planet’s semi-major axis. The ability to make this measurement depends not only on the astrometric precision for individual observations (2.8 to 5.9 mas; Table 3), but also depends critically on the orbital sampling. If planets do not trace out their full orbit during the two-year observing window, the quality of the fit is reduced. This is particularly true for stars with high luminosity, which translates to a more distant habitable zone and hence longer orbital periods. Sirius, the most luminous star in our sample (30.5 L), has the worst precision in its orbit fitting (72.5 mas), whereas eps Ind, the least luminous star (0.23 L), has the best determined orbit (3.5 mas). Table 3 lists the median semi-major axis precision obtained for the other target stars.

Our ultimate objective is to determine whether a planet lies within its parent star’s habitable zone. The key metric for this determination is not the absolute precision, but rather the fractional precision on a planet’s semi-major axis. While the absolute precision varies between Sirius and eps Ind by a factor of 20, the fractional precision for the two is comparable (2.7% and 2.1%, respectively) since Sirius’ habitable zone is a factor of 16 larger than eps Ind’s. For the overall sample, the median fractional precision varies from 2.1% up to 7.6% for Fomalhaut.

Figure 4 shows semi-major axis measurement precision versus true semi-major axis for each of the 1000 simulated planets around three of our target stars. 82 Eri (left panel) has one of the best precisions (4.4 mas median), although the performance degrades significantly for more distant orbits, where the planets are relatively faint. Fomalhaut (central panel) has one of the worst precisions (50.5 mas median), primarily because the planets in its habitable zone around this A-type star have periods considerably longer than our 2-year mission lifetime (5 to 12 years), such that only a fraction of each orbit is traced. The effect of limited phase coverage can be seen in Figure 5, which plots semi-major axis precision as a function of orbital period in the center of the habitable zone. Planets with periods less than our 2-year mission lifetime are well constrained, but those farther out have semi-major axis precision increasing roughly linearly with the period.

Fig. 4

For each star (three examples shown here—82 Eri, Fomalhaut, and sigma Dra), the orbital parameters and their uncertainties are retrieved for 1000 random planet orbits, each of which is directly imaged at least three times. The precision for measuring the semi-major axis of each planet is shown here as a function of the true semi-major axis. The habitable zone is interior to the dashed lines. The starshade masks all orbits inside of 100 mas. Hash marks at the bottom of each panel correspond to orbital periods equal to the spacing between observing epochs, with a 1-year period highlighted in red.

JATIS_7_2_021219_f004.png

Fig. 5

Our ability to pin down a planet’s semi-major axis depends on the fraction of its orbit that is traced by the observations. Periods less than the mission lifetime (2 years) are well sampled, whereas those with longer periods are only observed for a partial arc, resulting in lower precision in determining the orbit.

JATIS_7_2_021219_f005.png

While the semi-major precisions for other target stars (shown in Fig. 8 in the Appendix) follow a similar pattern of smoothly decreasing precision with increasing ap, sigma Dra (right panel in Fig. 4) has an unusual bump around ap160  mas, corresponding to orbital periods of 1  year. The decrease in performance is due to the sampling being repeated on 1-year cycles, where observations taken during the second year of the mission have about the same orbital anomaly as those taken in the first year of the mission (i.e., there is a 1-year aliasing). The resulting small range of orbital-anomaly coverage (see the right panel of Fig. 2) makes it more difficult to fit the orbit.

While the 1-year aliases just discussed is most pronounced for sigma Dra, several other systems exhibit a similar effect at orbital periods that match the phasing of the observations. For all of the precision plots (Figs. 4 and 8), the semi-major axis corresponding to a 1-year period is indicated by a red hash mark; the other black marks correspond to other time differences between observing epochs (e.g., for 82 Eri, the first and second observations are separated by 40 days, the second and third by 325 days, and the first and fourth by 405 days; see the observing windows in Fig. 1). While the 1-year aliasing generally causes the strongest effect, other orbital period/observing period alignments can also degrade performance.

Our MCMC fitting procedure calculates a (non-Gaussian) posterior distribution for each orbital parameter. From these distributions we derive the probability that each planet lies within its star’s habitable zone. Figure 6 shows the results for the same three target stars as in Fig. 4. For 82 Eri (left panel), the orbit fitting is fairly deterministic—planets well inside the habitable zone are correctly identified as such with high probability (>99%), whereas those well outside are ruled out (probability <1%). As one would expect, there is some ambiguity near the edges of the habitable zone, but for the overall sample there is just a 2.4% chance of a habitable zone planet being falsely classified as falling outside the habitable zone, whereas 98.8% of the planets classified as residing in the habitable zone are truly habitable zone planets (i.e., a false positive rate of 1.2%).

Fig. 6

As in Fig. 4, orbital parameters are measured for 1000 random planetary orbits around each of three stars—82 Eri, Fomalhaut, and sigma Dra. The derived probability of each planet residing in its star’s habitable zone is shown as a function of its true semi-major axis. The habitable zone is indicated by the dashed lines.

JATIS_7_2_021219_f006.png

These rates are based on a nominal classification threshold, where planets with habitable zone probability >50% are categorized as habitable zone planets. If a more conservative approach is desired, less planets can be included as habitable zone. If a 95% threshold is used for classification, for example, then 82 Eri will have only 0.2% false detections. However, only 88.9% of the true habitable zone planets will be included (11.1% false negative rate).

For Fomalhaut (middle panel of Fig. 6), the worse orbital precision translates to much more scatter in the plotted probabilities and less certainty for determining whether a planet lies in its habitable zone. Still, there is only 4% probability of a habitable zone planet being misclassified, and only a 4.9% chance of a habitable-zone-classified planet not being truly in the habitable zone. For the conservative (95%) classification threshold, the false positive rate falls to zero, but at the expense of only 59% of the true habitable zone planets being included (i.e. a false negative rate of 41%).

The false positive and false negative rates for 82 Eri, Fomalhaut, and sigma Dra are shown in Fig. 7, as a function of the classification threshold. These plots also display an overall success metric—the F1 score, the harmonic mean of (1—false positive rate) and (1—false negative rate). While a nominal threshold of 50% results in an optimal balance between these two factors (i.e., it give maximum F1 score), an emphasis on avoiding false detections would warrant a more conservative approach.

Fig. 7

False positive and false negative rates for 82 Eri, Fomalhaut, and sigma Dra, shown as a function of the habitable-zone probability threshold. A more liberal threshold (lower probability) reduces the false negative rate, but increases the false positive rate. A very conservative threshold (to the right of each panel) can ensure that no false detections are made, but misses a significant number of true habitable-zone planets. A performance metric combining the two rates is given by the F1 score, the harmonic mean of the precision (the fraction of detected habitable zone planets that truly are located in the habitable zone) and recall (the fraction of habitable zone planets that are correctly classified as such).

JATIS_7_2_021219_f007.png

The false positive/false negative rates for each star are listed in Table 3 for both the nominal classification threshold (50%) and a conservative classification threshold (95%). For the nominal threshold, the average performance for the overall sample is a 2.8% false positive rate and a 3.3% false negative rate. For the conservative threshold, the average false positive rate is just 0.05%, but the average false negative rate goes up to 19%.

5.

Discussion

The results of this study should be considered as a lower bound on the sensitivity since it is based on a simple SNR estimation and has not included the difficulties of image processing for handling non-ideal behaviors. These include stray light from the starshade, solar glint, starlight leakage, fluctuating position errors of the starshade, structure, and gradients in the exozodiacal light, and overlapping exoplanet images in complex systems. These issues are best addressed using imaging simulations such as SISTER23 and will be the subject of future work. Many of these issues will be addressed with a community-based data challenge.24 A number of these non-ideal behaviors have been shown in Ref. 25. Diffraction of sources near or below the IWA could introduce additional astrometric position errors and will need to be studied further in future imaging simulation work. While the upcoming data challenge will focus on issues specific to starshade imaging, it will also consider the more generic problem of interference between multiple planets in a single system. The models and expected performance presented in this work and Ref. 2 will serve as a useful metric to compare against the results of the Starshade data challenge and can provide a base model to estimate the magnitude of systematic uncertainties due to these non-ideal behaviors.

6.

Conclusions

Based on a model for the SRP mission concept with target-specific observing windows and SNR calculations dependent on the planet illumination during each window, we have quantified the ability of SRP to identify habitable zone planets. We find that detection of a planet in at least 3 out of the 4 observing epochs will adequately measure the planet’s semi-major axis. For a 16 star sample observed with this strategy, we find that habitable zone planets are correctly identified as such 96.7% of the time, with 2.8% contamination by false classifications. Including the full range of planet masses, the mission is expected to detect 10 planets in the vicinity of the habitable zone,2 such that a very small number of planets (<1) are expected to be misclassified.

7.

Appendix A: Additional Figures

In this appendix we provide figures for astrometric precision and the probability that an exoplanet is correctly identified as being a habitable zone exoplanet as a function of semi-major axis for all targets considered in this study (Figs. 8 and 9).

Fig. 8

For each of our 16 target stars, the orbital parameters are retrieved for 1000 random planet orbits, each of which is directly imaged at least three times. The precision for measuring each planet’s semi-major axis is shown here.

JATIS_7_2_021219_f008.png

Fig. 9

For each target star, orbital parameters and their uncertainties are measured for 1000 random planetary orbits. The probability of each planet residing in its star’s habitable zone is shown as a function of its true semi-major axis.

JATIS_7_2_021219_f009.png

Acknowledgments

We thank the reviewers for many insightful comments that have improved the quality of this paper. Part of this work was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. © 2021. All rights reserved. This research has made use of (1) the NASA Exoplanet Archive, which is operated by the California Institute of Technology, under contract with the National Aeronautics and Space Administration under the Exoplanet Exploration Program and (2) the SIMBAD database, operated at CDS, Strasbourg, France.

References

2. 

A. Romero-Wolf et al., “Starshade Rendezvous: exoplanet sensitivity and observing strategy,” J. Astron. Telesc. Instrum. Syst., 7 (2), 021210 (2021). https://doi.org/10.1117/1.JATIS.7.2.021210 Google Scholar

3. 

D. Mawet et al., “Deep exploration of Eridani with Keck Ms-band vortex coronagraphy and radial velocities: mass and orbital parameters of the giant exoplanet,” Astron. J., 157 33 (2019). https://doi.org/10.3847/1538-3881/aaef8a ANJOAA 0004-6256 Google Scholar

4. 

E. B. Ford, “The effects of multiple companions on the efficiency of space interferometry mission planet searches,” Publ. Astron. Soc. Pac., 118 364 –384 (2006). https://doi.org/10.1086/500813 PASPAU 0004-6280 Google Scholar

5. 

O. Guyon et al., “Simultaneous Exoplanet characterization and deep wide-field imaging with a diffractive pupil telescope,” Astrophys. J., 767 11 (2013). https://doi.org/10.1088/0004-637X/767/1/11 ASJOAB 0004-637X Google Scholar

6. 

S. Blunt et al., “Orbits for the impatient: a Bayesian rejection-sampling method for quickly fitting the orbits of long-period exoplanets,” Astron. J., 153 229 (2017). https://doi.org/10.3847/1538-3881/aa6930 ANJOAA 0004-6256 Google Scholar

7. 

A. Horning, R. Morgan and E. Nielson, “Minimum number of observations for exoplanet orbit determination,” Proc. SPIE, 11117 111171C (2019). https://doi.org/10.1117/12.2529741 PSISDG 0277-786X Google Scholar

8. 

C. M. Guimond and N. B. Cowan, “Three direct imaging epochs could constrain the orbit of Earth 2.0 inside the Habitable Zone,” Astron. J., 157 188 (2019). https://doi.org/10.3847/1538-3881/ab0f2e ANJOAA 0004-6256 Google Scholar

10. 

B. S. Gaudi et al., “The Habitable Exoplanet Observatory (HabEx) mission concept study final report,” (2020). Google Scholar

11. 

C. C. Stark et al., “Maximized exoEarth candidate yields for starshades,” J. Astron. Telesc. Instrum. Syst., 2 041204 (2016). https://doi.org/10.1117/1.JATIS.2.4.041204 Google Scholar

12. 

J. F. Kasting, D. P. Whitmire and R. T. Reynolds, “Habitable zones around main sequence stars,” Icarus, 101 108 –128 (1993). https://doi.org/10.1006/icar.1993.1010 ICRSA5 0019-1035 Google Scholar

13. 

L. A. Rogers, “Most 1.6 earth-radius planets are not rocky,” Astrophys. J., 801 41 (2015). https://doi.org/10.1088/0004-637X/801/1/41 ASJOAB 0004-637X Google Scholar

14. 

K. J. Zahnle and D. C. Catling, “The Cosmic Shoreline: the evidence that escape determines which planets have atmospheres, and what this may mean for Proxima Centauri B,” Astrophys. J., 843 122 (2017). https://doi.org/10.3847/1538-4357/aa7846 ASJOAB 0004-637X Google Scholar

15. 

T. D. Robinson and C. T. Reinhard, “Earth as an exoplanet,” (2018). Google Scholar

16. 

M. C. Turnbull et al., “The search for Habitable Worlds. 1. The viability of a Starshade mission,” Publ. Astron. Soc. Pac., 124 418 –447 (2012). https://doi.org/10.1086/666325 PASPAU 0004-6280 Google Scholar

17. 

S. Ertel et al., “The HOSTS survey for exozodiacal dust: observational results from the complete survey,” Astron. J., 159 177 (2020). https://doi.org/10.3847/1538-3881/ab7817 ANJOAA 0004-6256 Google Scholar

18. 

C. Leinert et al., “The 1997 reference of diffuse night sky brightness,” Astron. Astrophys. Suppl. Ser., 127 1 –99 (1998). https://doi.org/10.1051/aas:1998105 AAESB9 0365-0138 Google Scholar

19. 

S. B. Shaklan, L. Marchen and E. Cady, “Shape accuracy requirements on starshades for large and small apertures,” Proc. SPIE, 10400 104001T (2017). https://doi.org/10.1117/12.2273436 PSISDG 0277-786X Google Scholar

20. 

J. J. Condon, “Errors in elliptical Gaussian fits,” Publ. Astron. Soc. Pac., 109 166 –172 (1997). https://doi.org/10.1086/133871 PASPAU 0004-6280 Google Scholar

21. 

K. Mede and T. D. Brandt, “The Exoplanet simple orbit fitting toolbox (ExoSOFT): an open-source tool for efficient fitting of astrometric and radial velocity data,” Astron. J., 153 135 (2017). https://doi.org/10.3847/1538-3881/aa5e4a ANJOAA 0004-6256 Google Scholar

22. 

D. Foreman-Mackey et al., “emcee: The MCMC Hammer,” Publ. Astron. Soc. Pac., 125 306 –312 (2013). https://doi.org/10.1086/670067 PASPAU 0004-6280 Google Scholar

23. 

S. H. Rafels et al., “SISTER: Starshade imaging simulation toolkit for exoplanet reconnaissance (conference presentation),” Proc. SPIE, 11117 111170O (2019). https://doi.org/10.1117/12.2528332 PSISDG 0277-786X Google Scholar

24. 

R. Hu et al., “Starshade exoplanet data challenge,” J. Astron. Telesc. Instrum. Syst., 7 (2), 021216 (2021). https://doi.org/10.1117/1.JATIS.7.2.021216 Google Scholar

25. 

E. Peretz et al., “Exoplanet imaging performance envelopes for starshade-based missions,” J. Astron. Telesc. Instrum. Syst., 7 (2), 021215 (2021). https://doi.org/10.1117/1.JATIS.7.2.021215 Google Scholar

Biography

Andrew Romero-Wolf received his BA degree in mathematics and the BA degree in physics (with honors) from the University of Chicago in 2002, a master’s in physical science from the University of Chicago in 2005, and a PhD from the University of Hawaii at Manoa in 2010. He is currently a member of the technical staff at the Jet Propulsion Laboratory, California Institute of Technology in the Deep Space Tracking group.

Geoffrey Bryden is an astronomer at the Jet Propulsion Laboratory (JPL). His research focuses on the detection and characterization of planetary systems beyond the Solar System, in particular the debris left over after planets form. He received his PhD in astrophysics from the University of California in 1999.

Greg Agnes received his BS from Rensselaer, his MS from University of MD and his PhD from Virginia Tech. He currently is a senior mechanical systems engineer in the Advanced Deployable Structures group at JPL, where he architects large precision and deployable space instruments.

Jonathan W. Arenberg is chief mission architect for Science and Robotic Missions at Northrop Grumman Space Systems. His degrees are in physics and engineering, from the University of California, Los Angeles. Arenberg has contributed to the Chandra X-ray Observatory, Starshade and the James Webb Space Telescope and the Astro 2020 strategic missions. He has authored over 200 conference presentations, papers, a recent book, awarded 15 European and U.S. Patents and is an SPIE fellow.

Samuel Case Bradford earned his PhD at Caltech, and his BS degree from UC Berkeley. His PhD research focused on developing signal processing techniques to assess the stability of systems with changing structural properties. He is currently the supervisor of the Advanced Deployable Structures Group at NASA’s Jet Propulsion Laboratory and is a visitor in Aerospace at Caltech. His recent research has been focused on the design, fabrication, and testing of deployable space structures.

John Debes received his BA degree in physics in 1999 from Johns Hopkins University and his PhD in astronomy and astrophysics from Pennsylvania State University in 2005. He is an ESA/AURA astronomer and deputy head of the Science Mission Office at the Space Telescope Science Institute. He has authored or co-authored more than 70 refereed papers and one book chapter. His interests include high-contrast imaging of exoplanets and debris disks, as well as the study of planetary systems that survive postmain sequence evolution.

Matthew Greenhouse has been a civil servant astrophysicist at the Goddard Space Flight Center Observational Cosmology Laboratory since 1996. He is a senior member of the James Webb Space Telescope project science team and leads a NASA Strategic Astrophysics Technology project for development of micro-shutter arrays for space-based multi-object spectroscopy.

Renyu Hu is a scientist at the Jet Propulsion Laboratory (JPL) and his research strives to identify and characterize habitable environments in the Solar System and beyond. He is the Starshade scientist of the NASA Exoplanet Exploration Program, providing science leadership to the S5 Starshade Technology Development Activity and managing the Starshade Science and Industry Partnership program. He received his PhD in planetary science from Massachusetts Institute of Technology in 2013.

Jason Rhodes is a senior research scientist at NASA’s JPL. He specializes in the development of space missions for cosmology and exoplanets.

John Ziemer received his BSE degree in aerospace engineering from the University of Michigan and his MA and PhD degrees in mechanical and aerospace engineering from Princeton University with a focus on electric propulsion. He has been at JPL for more than 20 years and is currently a principal systems engineer working on a variety of different concept and technology development efforts.

Biographies of the other authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Andrew Romero-Wolf, Geoffrey Bryden, Greg Agnes, Jonathan W. Arenberg, Samuel Case Bradford, Simone D'Amico, John Debes, Matt Greenhouse, Renyu Hu, Steve Matousek, Jason Rhodes, and John Ziemer "Starshade Rendezvous: exoplanet orbit constraints from multi-epoch direct imaging," Journal of Astronomical Telescopes, Instruments, and Systems 7(2), 021219 (4 June 2021). https://doi.org/10.1117/1.JATIS.7.2.021219
Received: 25 January 2021; Accepted: 20 May 2021; Published: 4 June 2021
Lens.org Logo
CITATIONS
Cited by 1 scholarly publication.
Advertisement
Advertisement
KEYWORDS
Planets

Stars

Exoplanets

Imaging systems

Signal to noise ratio

Observatories

Telescopes

Back to Top