Open Access
1 September 2021 Integration time adjusted completeness
Dean Keithly, Dmitry Savransky, Corey Spohn
Author Affiliations +
Abstract

Future, large-scale, exoplanet direct-imaging missions will be capable of discovering and characterizing Earth-like exoplanets. These mission designs can be evaluated using completeness, the fraction of planets from some population that are detectable by a telescope at an arbitrary observation time. However, the original formulation of completeness uses instrument visibility limits and ignores additional integration time and planetary motion constraints. Some of the sampled planets used to calculate completeness may transit in and out of an instrument’s geometric and photometric visibility limits while they are being observed, thereby causing the integration time agnostic calculation to overestimate completeness. We present a method for calculating completeness that accounts for the fraction of planets that leave the visibility limits of the telescope during the integration time period. We define completeness using the aggregate fraction of an orbital period during which planets are detectable, calculated using the specific times that planets enter and leave an instrument’s visibility limits and the integration time. To perform this calculation, we derive analytical methods for finding the planet-star projected separation extrema, times past periastron that these extrema occur, and times past periastron that the planet-star projected separation intersects a specific separation circle. We also provide efficient numerical methods for calculating the planet-star difference in magnitude extrema and times past periastron corresponding to specific values Δmag. Our integration time adjusted completeness shows that, for a planned star observation at 25 pc with 1-day and 5-day integration times, integration time adjusted completeness of Earth-like planets is reduced by 1% and 5% from the integration time agnostic completeness, respectively. Integration time adjusted completeness calculated in this manner also provides a computationally inexpensive method for finding dynamic completeness—the completeness change on subsequent observations.

1.

Introduction

Direct imaging blind search mission schedules can be optimized1 by maximizing completeness2—the fraction of exoplanets from an assumed planet population that are detectable by a particular instrument at an arbitrary observation time. Completeness is typically parameterized by a limiting planet-star brightness difference (Δmaglim), the inner working angle (IWA) of the instrument, and the outer working angle (OWA) of the instrument. The original Monte Carlo approach developed by Brown2 involved creating a cloud of synthetic planets by sampling the underlying Keplerian orbital elements (KOE) and physical parameters of a planet population and determining the fraction of those individually simulated planets within the visible limits of the instrument. Multiplying completeness by the exoplanet occurrence rate gives the expected exoplanet yield for observing a given star. Although completeness is a good metric for predicting instrument performance, the calculation described above only captures an instant in time and does not include whether the time that a planet is within the visible region of the telescope is sufficient to actually make a detection.

The original method of calculating completeness was developed by Brown.2 Our adaptation of this method in EXOSIMS3 is extensively outlined in Ref. 1 and involves randomly generating planet KOE, planet geometric albedo (p), and planetary radius (R) sampled from NASA’s Exoplanet Program Analysis Group Study Analysis Group 13 (SAG13)1,4,5 probability density functions, generating a finely binned 2D histogram of Δmag versus s (planet-star separation), and fitting a 2D spline to the histogram bins. In all current methods, completeness is calculated as the double integral of s and Δmag over the joint probability density function as in Eq. (17) of Ref. 1 and Eq. (7) in Ref. 6. The limits of integration define a detectable planet as one in which Δmag<Δmaglim and diIWA<s<diOWA, where di is the distance of the host star from the spacecraft. While the Δmaglim in this paper is used to describe a general upper limit of integration for calculating completeness, this Δmaglim can be formulated as a function of integration time as in Eq. (12) of Ref. 1 by making assumptions about a multitude of instrument parameters and external noise sources. This approach, as conventionally implemented, is a good estimator for completeness,1 but it does not take planet motion in time into account.

Searches like the Gemini Planet Imager and Nancy Grace Roman Space Telescope (Roman) Coronagraphic Instrument (CGI)7 are sensitive to larger planets with larger planet-star separations and longer orbital periods.1 They make use of Brown completeness to plan blind searches. However, future telescope concepts like the Habitable Exoplanet Observatory (HabEx)8 and the Large UV/Optical/IR Surveyor (LUVOIR)9 seek to find smaller Earth-like exoplanets, with shorter orbital periods around stars farther away. Their search criteria will result in Brown completeness based yield overestimations due to planet motion out of instrument visibility limits. Figure 1 shows a schematic of a direct imaging observation; it demonstrates how a planet can be within the photometric and geometric visibility limits of the instrument and still not be detected. The red regions of the planet’s orbit indicate where the planet is within the visible limits but not detectable because the planet will move out of the region in less time than it takes to detect the planet [case (b) in this figure]. Since there will be some Earth-like exoplanets that are counted toward the completeness score but are not actually within the visibility constraints of the instrument long enough to be directly imaged, we need a new method that only counts targets within the instrument visibility long enough to be observed or characterized.

Fig. 1

Schematic of a direct imaging observation. The line represents the projection of a planet’s orbit about its parent star (yellow) into the plane of the sky as seen by a distant observer. The arrow indicates the planet’s direction of motion. Blue portions of the orbit (a) indicate times when the planet is detectable. Red portions of the orbit (b) indicate times when the planet is within the visibility limits of the instrument, but it is not detectable due to integration time constraints. The dashed portions of the orbit (c) indicate times when the planet is too faint to be observed. The shaded gray regions represent the projected inner and OWAs of the instrument. The planet is unobservable on portions of the orbit intersecting these gray regions. Green dots indicate transitions in and out of instrument visibility limits.

JATIS_7_3_037002_f001.png

Figure 2 demonstrates this phenomenon in the separation versus Δmag phase space commonly used to define instrument contrast curves and completeness. Figure 2(b) shows a short-period, fast-moving, Mars-like planet that transitions into and out of the assumed instrument’s visible limits before it can be detected. Figure 2(c) shows the effect for a Uranus-like planet that crosses through and exits the visible region in less time than it takes for it to be detected. In these cases, Brown’s completeness calculation would include simulated planets in these regions, thus overestimating the overall detection yield. Additionally, Fig. 2(d) shows that the last possible moment a Neptune-like planet could be imaged before it leaves the visible limits of the telescope. This demonstrates a portion of the orbit where the planet is within the visible limits of the telescope but is not detectable.

Fig. 2

The Δmag versus s curve for Neptune (blue) and Mars (red). The instrument’s visible region (white) is bounded by 0.066 and 0.517 arc sec, and Δmag=31 at di=22.87 pc (we selected instrument parameters to make a demonstration on solar system planets about a Sun-like star, but any instrument will have planets with similar effects). Any planet in the grayed region is not visible by the instrument. Dots indicate calculated intersection points between the planet and the visible region bounds. The black dashed portions of planet orbits indicate portions of the orbit where the planet is not detectable. (b) A portion of Mars’s orbit where the planet enters and exits the visible limits of the instrument in less time than the required integration time.1 (c) A portion of Neptune’s orbit where the planet enters and exits the visible region of the instrument in less time than the required integration time. (d) Neptune’s detectable region adjusted for integration time. If the observation of Neptune in (d) is started when it is left of the blue × and in the visible region, then it will be detected; if it is started right of the blue ×, the planet will not be detected because the required integration time is larger than the time the planet will be in the visible region of the instrument.

JATIS_7_3_037002_f002.png

This paper presents a method, implemented in the EXOSIMS modeling software,3 for calculating integration time adjusted completeness that allows us to investigate exactly how much the original completeness definition overestimates planet yields. EXOSIMS is an exoplanet direct imaging mission modeling software used to simulate the Roman CGI7 and future mission concepts such as HabEx8 and LUVOIR.9 EXOSIMS simulates populations of planets, observatories, instruments, and underlying dynamics to create full, end-to-end mission simulations. We derive distributions of potential mission yield from ensembles of these mission simulations. EXOSIMS utilizes completeness as a heuristic for target selection and observation scheduling,1 implementing both Brown’s Monte Carlo approach to completeness2 as well as the analytical completeness formulation from Ref. 6. However, EXOSIMS determines whether actual detections occur by evaluating the effective signal-to-noise ratio (SNR) of individual simulated observations on simulated planets. Keithly et al.1 discussed the full optimization process for a single-visit blind search and showed that the ideal yield from an EXOSIMS simulation ensemble (when mission constraints are discounted) for a limited blind search mission of Roman is equivalent to the yield calculated via Brown Completeness.

Comparisons between Brown completeness-based exoplanet yield estimations and Monte Carlo of full mission simulations have been done in Ref. 10 and the Standard Definitions and Evaluation Team Final Report.11 The yields calculated by completeness are higher than the average yield from a Monte Carlo of full missions simulated in EXOSIMS as noted in Table 6 of Ref. 11. These works attribute the difference in yields to the additional mission constraints captured by the full mission simulation approach, inefficiencies in the optimization algorithms, and inefficiencies in the scheduling algorithms implemented in EXOSIMS.12 The unmentioned assumption is that Brown completeness-based yield estimates are accurate, but this paper demonstrates how the omission of planet motion and integration time reduces the resulting completeness yield estimates.

To know the amount of time that an individual planet spends within the observable limits of an instrument, we need to calculate the times in a planet’s orbit when a planet enters or exits these limits. This means that we need methods for calculating when a planet has a given planet-star separation and Δmag. Once we know all of the times when a planet enters or exits an instrument’s visibility limits, we know the fraction of time that the planet is able to be detected with that instrument. Averaging over these visibility fractions for a large number of samples evaluates completeness in a fundamentally different way from its original formulation. These same methods also allow for the evaluation of integration time adjusted completeness as well as a new method of calculating dynamic completeness.13

Section 2 presents the derivation of this chain of calculations and discusses practical aspects of their implementation. Section 3 provides detailed validation of the methods and presents results of the various calculations enabled by them. Finally, in Section 4, we discuss various aspects of the algorithms and results, and we lay out future applications for this methodology.

2.

Methods

In this section, we present our detailed process for calculating integration time adjusted completeness. We do this by finding the time windows in which a planet is within the separation and Δmag visibility limits of an instrument, discounting each time window by the integration time. The general overview of this process is as follows.

  • 1. Calculate locations of apparent intersections between the projection of the 3D orbit into the plane of the sky and the sWA circle about the star.

    • (a) Parameterize the 3D orbit as an ellipse in the plane of the sky.

    • (b) Formulate the planet-star separation equation and its derivative.

    • (c) Solve for the magnitude and locations of planet-star separation extrema.

    • (d) Identify and assign the subset of algebraic solutions that are separation extrema.

    • (e) Classify the expected number of intersections between the projected ellipse and the sWA circle.

    • (f) Solve for intersections between the projection of the orbital ellipse and the sWA circle.

    • (g) Identify and assign the subset of algebraic solutions that form intersections.

  • 2. Calculate locations and magnitudes of Δmag intersections.

    • (h) Formulate an expression isolating the Δmag and ν terms and its derivative.

    • (i) Express this formulation as a polynomial and solve for the algebraic solutions.

    • (j) Identify and assign the subset of algebraic solutions that are Δmag extrema.

    • (k) Classify the expected number of Δmag and Δmaglim intersections.

    • (l) Solve for Δmag and Δmaglim intersections.

    • (m) Identify and assign the subset of algebraic solutions that form the intersections.

  • 3. Calculate ν from X and Y for each intersection and extrema.

  • 4. Calculate t from ν of each intersection and extrema.

  • 5. Combine times of sWA and Δmaglim intersections to create time windows between intersections.

  • 6. Identify time windows in which the planet is visible or not visible.

  • 7. Calculate integration time adjusted completeness averaging the orbital fraction of time that a planet is visible discounted by the integration time.

The general equations for Δmag and s used in these derivations are

Eq. (1)

Δmag=2.5log10(p(R|r̲k/i|)2Φ(β)),
and

Eq. (2)

s=r_k/i(r_k/i·r^̲i/SC)r^̲i/SC.
Here R is the planet radius, p is the geometric albedo of the planet, and Φ(β) is the planet phase function. The other variables are defined in Fig. 3, where β is the star-planet-observer angle (also called the phase angle), r_k/i is the vector from star i to planet k, and r^̲i/SC is the unit vector from the spacecraft (SC) to the star (r^̲ is the unit vector of r_). The plane of the sky for a given observation lies in the x^̲ and y^̲ plane, where r^̲i/SC defines z^̲ of the target system. For our purposes, the direction of x^̲ is arbitrary, but is typically taken to be a well-defined, inertially fixed direction, such as the ICRS mean equinox or pole direction. Here and throughout this paper, i in a subscript refers to the i’th target star and i (not subscripted) refers to the orbit inclination.

Fig. 3

The orbital path of planet k (black) in a general XYZ Cartesian coordinate system. r_k/i (straight blue arrow) describes the location of the planet k (blue circle) relative to star i (yellow circle). The plane of the sky is noted by the translucent red parallelogram entirely in the x^- and y^- plane. The dashed ellipse (red) is the projection of the planet orbit (black) onto the plane of the sky. h^- is the orbit angular momentum vector. e^- is the orbit eccentricity vector. n^- is the line of nodes. The star * is generally referred to in subscripts as * and often referred to in subscripts as the i’th star. This is not to be confused with the variable i, which is the inclination of the planet orbit. ν is the true anomaly of the planet. ω is the argument of periapsis of the planet. Ω is the longitude of the ascending node of the planet.

JATIS_7_3_037002_f003.png

2.1.

Projected Orbit and Separation Intersection

In this section, we derive an analytical expression for the true anomaly (ν) of s-orbit intersection points between a circle in the plane of the sky and the projection of a 3D Keplerian orbit on the plane of the sky. A practical example of a circle in the plane of the sky is the projected inner or OWA of the instrument, equal to IWAdi (or OWAdi) for target star distance di. An orbit defined by KOE takes the shape of an ellipse in the orbital plane. The perpendicular projection of a 3D elliptical orbit onto the plane of the sky is given by the x^̲ and y^̲ components of the 3D orbit. However, the resulting expression is not easily solvable for the true anomalies at intersection points. Instead, we can express the orbit projection in the plane of the sky as another ellipse. This simplifies the difficult intersection problem into analytically solvable subproblems and yields a relatively simple solution to the intersection between a circle in the plane of the sky and the projected orbit in the plane of the sky. Here we include the brief outline of this process, with the full procedure detailed in Appendix B.

  • 1. Project the 3D orbital ellipse into a 2D projected ellipse and formulate the separation equation.

  • 2. Find the expected number of s intersection points.

  • 3. Calculate s intersection point coordinates.

2.1.1.

Projection of the elliptical orbit

Given the KOE of a 3D orbit and a plane to project it on, we calculate the semimajor and semiminor axes of the projected orbit (ap and bp, respectively) as well as the angle between the projected semimajor axis and x^̲, which we call (θ). We first convert the KOE into orbital radius components in the Cartesian XYZ coordinate system as in Fig. 3:

Eq. (3)

X=r(cos(Ω)cos(ω+ν)sin(Ω)sin(ω+ν)cos(i))Y=r(sin(Ω)cos(ω+ν)+cos(Ω)sin(ω+ν)cos(i))Z=rsin(i)sin(ω+ν),
where r is the orbital radius (magnitude of r_k/i) given by

Eq. (4)

r=a(1e2)1+ecos(ν).
Ω is the longitude of the ascending node, ω is the argument of periapsis, ν is the true anomaly, and e is the eccentricity of the planet’s orbit.

Figure 4 shows a schematic view of the projection of the orbit onto the plane of the sky. There are two particularly important points associated with the projected orbit ellipse. First, F is the filled focus (star location) of the orbit and retains the same coordinates in the plane of the sky. When observing a star, F is the center of all working angle circles and the origin of the XYZ coordinate system. Note that in Fig. 4 the projected ellipse is located well below the original 3D ellipse for clarity, but the points F indicated by the orange circle and F indicated by the orange × are in fact coincident.

Fig. 4

The original 3D elliptical orbit of a planet (black line) containing points A, B, C, and D; the gray endpoints of the purple semimajor axis AB and semiminor axis CD. The planet orbits about the star, which is located at focus F (orange circle). The orange dot and orange × denoted as F are the same point in space, but the red projection is shown as offset from the original orbit for clarity. The projection of the original 3D elliptical orbit onto the XY plane is given by the red ellipse containing points A, B, C, and D (blue diamonds; perpendicular projections of A, B, C, and D). Point O is the geometric center of the 3D elliptical orbit and projects to O in the plane of the sky. Point P (green circle) is any arbitrary point along the original 3D ellipse and maps to the semimajor axis and semiminor axis components H and K, respectively (green ×). P (pink circle) is the perpendicular projection of P, and H and K (pink ×) are projections of H and K, respectively. The components of these perpendicular projections preserve the ratios of their values to the semimajor and semiminor axes, which, given the equation for an ellipse, can be used to prove the projection of an ellipse is itself an ellipse. The blue lines AB and CD form conjugate diameters of the red ellipse and are the projection of the semimajor axis and semiminor axis of the original 3D ellipse onto the plane of the sky. These conjugate diameters can then be used to find the semimajor axis and semiminor axis of the red projected ellipse IR and ST, respectively. x_dr, y_dr, z_dr are the components of the derotated reference frame (dr) as defined in the text.

JATIS_7_3_037002_f004.png

Second, O is the geometric center of the 3D orbit and is found most efficiently by averaging the XYZ locations of the planet at apoapsis and periapsis:

Eq. (5)

FO¯=12(r_k/i(ν=0)+r_k/i(ν=π)),
where FO is the line segment (equivalently Euclidean vector) from F to O and r_k/i(ν) is the evaluation of Eq. (3) for the given value of ν. As x^̲ and y^̲ define the plane of the sky, the XYZ coordinates of the geometric center of the projected ellipse (O) are given by

Eq. (6)

O=FO¯·x^̲,FO¯·y^̲,0.

Appendix J provides a proof that any generic 3D ellipse projects to a 2D ellipse, as shown graphically in Fig. 4. The projection of the semimajor axis and semiminor axis from the 3D orbit onto the 2D plane of the sky form conjugate diameters of the projected ellipse. Any two diameters of an ellipse are conjugate diameters if and only if the tangent line to the ellipse at an endpoint of one diameter is parallel to the other diameter.14 Each pair of conjugate diameters of an ellipse has a corresponding tangent parallelogram, sometimes called a bounding parallelogram. In Fig. 4, the pair of blue lines (AB and CD) are specific examples of conjugate diameters. The pair of purple lines (IR¯ and ST) are also examples of conjugate diameters, and they form the semimajor and semiminor axes of the projected ellipse.

The semimajor and semiminor axes of the projected ellipse may be found from any conjugate diameters. We take the conjugate diameters AB and DC and draw line QQ through B, perpendicular to DC as shown by the gray line in Fig. 5. Points Q and Q are chosen such that BQ¯=BQ¯=OD¯. The principal axes IR and ST lie on the bisectors of the angles formed by lines OQ and OQ. From this construction, we calculate the projected ellipse semimajor axis, semiminor axis, and angular offset of the semimajor axis from x^̲.16 IR and ST are given by

Eq. (7)

IR¯=OQ¯+OQ¯,

Eq. (8)

TS¯=OQ¯OQ¯.
Defining the angle between OB¯ and OD¯ as ϕ, the cosine rule applied to triangles OBQ and OBQ yielding

Eq. (9)

|OQ¯|2=|OB¯|2+|OD¯|22|OB¯||OD¯|sinϕ,

Eq. (10)

|OQ¯|2=|OB¯|2+|OD¯|2+2|OB¯||OD¯|sinϕ,
where |OB| denotes the length of OB¯. Inserting these into Eq. (7), we obtain

Eq. (11)

OR¯·OS¯=|OB¯||OD¯|sinϕ,

Eq. (12)

|OR¯|2+|OS¯|2=|OB¯|2+|OD¯|2.
|IR| must be twice the semimajor axis of the projected ellipse (ap), so

Eq. (13)

ap=|IR¯|2=|OR¯|,
and |TS| must be twice the semiminor axis of the projected ellipse, so

Eq. (14)

bp=|TS¯|2=|OS¯|.
The angle of the semimajor axis of the projected ellipse from x^̲ is calculated using the average of the angles between OQ and x^̲ and OQ and x^̲:

Eq. (15)

θ=12(tan1(OQ¯·y^̲OQ¯·x^̲)+tan1(OQ¯·y^̲OQ¯·x^̲)).
Appendix E provides the full expressions for ap, bp, and θ via full expansions of Eqs. (13)–(15), respectively.

Fig. 5

The projected ellipse (red) has semimajor axis IR and semiminor axis TS (purple). Both are calculable from the conjugate diameters that are the projections of the semimajor and semiminor axes of the original 3D ellipse (AB, CD, blue). The line QQ is drawn such that it is perpendicular to the smaller conjugate diameter (CD) and is bisected by B: |BQ|=|BQ|=|OC|. The semimajor axis of the projected ellipse is the angular bisector of OQ and OQ.15 Finally, the semiminor axis TS of the projected ellipse is perpendicular to the semimajor axis.

JATIS_7_3_037002_f005.png

Now that we know all of the parameters necessary to describe the projected ellipse, we can standardize this ellipse into a simpler form to simplify subsequent calculations. We define a new frame (dr as in Fig. 4) as the derotation and geometric centering of the projected orbit such that the semimajor axis of the projected ellipse (OI¯) is aligned with x^̲dr, the semiminor axis of the projected ellipse (OS) is aligned with y^̲dr, and O is the origin of the dr coordinate system. The dr coordinates of the star location F (x*, y*) are given by a simple rotation of the projection of OF onto the x̲,y̲ plane by angle θ:

Eq. (16)

[x*y*]dr=[cos(θ)sin(θ)sin(θ)cos(θ)][OF¯·x^̲OF¯·y^̲]XYZ.

2.1.2.

Global and local extrema of planet-star separation

Before solving for the true anomalies where the orbit’s projected separation s is equal to sWA (general working angle separation WAdi), we first need to know how many of these sWA-orbit intersections we are looking for. We find the expected number of solutions by finding the s extrema throughout the orbit. We do so by solving for the roots of the derivative of the projected planet-star separation.

We start with the general equation for an ellipse:

Eq. (17)

(xeap)2+(yebp)2=1,
where xe and ye are the coordinates of any point on the ellipse and ap, bp are the semimajor and semiminor axes, respectively. We rewrite this in terms of xe, giving

Eq. (18)

ye=bp1xe2ap2.
The projected separation is given by

Eq. (19)

s2=(x*+xe)2+(y*+ye)2.
Taking the derivative of Eq. (19) with respect to xe, substituting Eq. (18) (and its derivative), and setting it equal to zero, we have

Eq. (20)

0=δs2δxe=2x*+2xe+2bpxey*apap2xe22bp2xeap2.
Isolating the square root term to one side and squaring both sides of the equation gives

Eq. (21)

(2x*+2xe2bp2xeap2)2=(2bpxey*apap2xe2)2,
which is expanded with coefficients of xe, to get the polynomial expression

Eq. (22)

0=xe4+8ap2x*+8bp2x*(4ap48ap2bp2+4bp4)/ap2xe3+4ap4+8ap2bp2+4ap2x*24bp4+4b2y*2(4ap48ap2bp2+4bp4)/ap2xe2+8ap4x*8ap2bp2x*(4ap48ap2bp2+4bp4)/ap2xe+4ap4x*2(4ap48ap2bp2+4bp4)/ap2.
This expression is a fourth-order polynomial, which we write in standard quartic form as

Eq. (23)

0=xe4+A0xe3+B0xe2+C0xe+D0,
where A0, B0, C0, and D0 are the constants and functions of ap, bp, x*, and y*, defined in Appendix F. We now apply the analytical solutions of the general quartic expression in standard form given in Appendix K.

Solving this quartic gives us a set of four xe solutions (xe) corresponding to two global extrema and two local extrema (if the latter exist for a particular orbit). Figure 6 shows a schematic representation of an orbit with four extrema. We use the imaginary components of the solutions, geometry of the ellipse, and magnitude of the higher order terms in the quartic solutions to identify which solutions belong to which extrema. We first use the magnitude of the imaginary components of solutions to determine how many extrema there are and filter out solutions that are not extrema. The algebraic solutions of the quartic polynomial all give values in the first quadrant (quadrants 1 to 4 are numbered counter-clockwise such that quadrant 1 has strictly positive coordinates, see the four quadrants of Fig. 6), so we must use geometry of the problem to determine the proper sign of each extremum’s true coordinates xe,g and ye,g [g references solutions 0 to 3 Eq. (91)–Eq. (94)].

Fig. 6

Planet orbit (black) in the dr frame and planet-star separation extrema. The minimum separation (cyan) always occurs in the same quadrant as the star in the dr frame (orange ×). The maximum separation (red) always occurs in the quadrants opposite the star. The local minimum separation (magenta) and local maximum separation (gold) occur in the same half-plane about the y axis as the star, but in opposite half-planes about the x axis as the star (fourth quadrant). This configuration applies to the vast majority of orbits. A small number of edge cases exist, including circular orbits and some edge on orbits. The blue dots represent the foci of the projected ellipse in the dr frame, the orange dot is the ellipse center, and the purple dashed lines are the semimajor and semiminor axes.

JATIS_7_3_037002_f006.png

All numerical solutions to the quartic have some degree of imaginary component due to accumulation of numerical errors. Quartic solution sets with only two extrema will have two real solutions (solutions with small imaginary components only due to numerical error) and two solutions with large imaginary components (algebraic solutions that are artifacts to be thrown away). The majority of KOE have only two s extrema. The only case in which no extrema exist is in a circular face-on orbit. In this specific case, all solutions to the quartic will be nearly identical and have large imaginary components, and the resulting s extrema will be identical (smin=smax). We assume that all solutions are real if |I(xe,g)|<105    g{0,1,2,3}. We assume only two solutions are real if |I(xe)|<105 for only two solutions. We define a new ordered set containing either two or four elements depending on magnitude of the imaginary components as

Eq. (24)

xR={|R(xe,g)|:|I(xe,g)|<105    g0..3},
where the absolute value is due to the algebraic solutions of the quartic being only defined in the first quadrant.

We now know the number of expected solutions from the dimension of xR but need to leverage the quartic algebraic solution and geometry to determine which components belong to which quadrant. For orbits with xR containing only two solutions, the first two solutions to the quartic (xR,0 and xR,1) produce the largest and smallest xR,gx*, respectively. Due to the shape of an ellipse, these must necessarily produce smin and smax. We define the set yR as the application of Eq. (18) to each element of xR. We define four separation quantities from the possible sign combinations of the coordinate magnitudes in xR and yR as

Eq. (25)

s±0=(xR,0x*)2+(yR,0±y*)2,

Eq. (26)

s+±0=(xR,0+x*)2+(yR,0±y*)2,

Eq. (27)

s±1=(xR,1x*)2+(yR,1±y*)2,

Eq. (28)

s+±1=(xR,1+x*)2+(yR,1±y*)2.

Typically, s--1 is smaller than s--0, but in 0.01% of cases, s--0 can be the smallest separation due to numerical error. (In this case, we ensure |s--1s--0|<108 and swap values.) In addition to s--1<s--0, we also know that s+0<s+1, s+1<s+0, and s++1<s++0 for all cases. Using this knowledge, we reduce the number of comparisons that we need to find the minimum planet-star separation to

Eq. (29)

smin={s--1  where(s--1<s+0)  and  (s--1<s+1)  and  (s--1<s++1),s+0  where(s+0<s--1)  and  (s+0<s+1)  and  (s+0<s++1),s+1  where(s+1<s+0)  and  (s+1<s--1)  and  (s+1<s++1),s++1  where(s++1<s+0)  and  (s++1<s+1)  and  (s++1<s--1).
The maximum separation similarly is found as

Eq. (30)

smax={s--0,  where(s--0>s+0)  and  (s--0>s+1)  and  (s--0>s++0),s+0,  where(s+0>s--0)  and  (s+0>s+1)  and  (s+0>s++0),s+1,  where(s+1>s+0)  and  (s+1>s--0)  and  (s+1>s++0),s++0,  where(s++0>s+0)  and  (s++0>s+1)  and  (s++0>s--0).
We are able to find xe and ye of the minimum and maximum separation diamonds drawn in Fig. 6 using the same logic as for finding smin and smax.

For the orbits with four solutions in xR, the first two elements will always be the global extrema, and the last two elements will be the local extrema, which are given by

Eq. (31)

smin=s--1,

Eq. (32)

smax=s++0,

Eq. (33)

slmin={s+3,s+2>s+3s+2,else,

Eq. (34)

slmax={s+2,s+2>s+3s+3,else,
where s+2 and s+3 are calculated in the same manner as s+0 and s+0. We are able to find the coordinates of all four using the same logic as for finding smin, smax, slmin, and slmax. This procedure yields the coordinates of all existing extrema in the dr frame. To find their locations on the projection of orbit in the plane of the sky, we apply the inverse of Eq. (16).

2.1.3.

Intersections between a circle and an ellipse

We find sWA-orbit intersections by formulating the circle-ellipse intersections as another quartic, solving this, and assigning the algebraic solutions to intersections. We assign solutions using the number of s extrema, the size of the sWA intersecting circle relative to these s extrema, and the ellipse geometry. To formulate the sWA-orbit intersections as a quartic, we start with Eq. (19) and substitute in Eq. (18). This gives us a separation equation solely as a function of xe and star location. We expand and transform this into a general polynomial of xe with a general s

Eq. (35)

0=(ap42ap2bp2+bp4ap4)xe4+(4ap2x+4bp2x*ap2)xe3+(2ap2bp22ap2s2+6ap2x*2+2ap2y*22bp4+2bp2s22bp2x*2+2bp2y*2ap2)xe2+(4bp2x+4  s2x4x*34x*y*2)xe+(bp42bp2s2+2bp2x*22bp2y*2+s42  s2x*22  s2y*2+x*4+2x*2y*2+y*4).
As in Sec. 2.1.2, we divide by the leading coefficient to convert this to the general quartic form:

Eq. (36)

0=xe4+A1xe3+B1xe2+C1xe+D1,
with A1, B1, C1, and D1 given in Appendix G. We solve this using the general quartic solution as given in Appendix K. This results in a solution for the xe of intersections in the first quadrant of the dr frame.

We always have four algebraic xe solutions that may or may not correspond to actual sWA-orbit intersections. The number of s extrema (either two or four) and projected separation relative to these extrema determine how many intersections will occur. (If we expect two intersections, then two of the four algebraic solutions must be real solutions, and the other two are some combination of repeated roots or non-physical imaginary solutions.) In each of these cases, we must handle the assignment of xe and ye solutions to the correct quadrants. Figure 7 shows a schematic representation of two cases corresponding to four total intersections.

Fig. 7

Diagrams of orbits where the projected ellipse in the dr frame produces four intersections (green dots) with the sWA separation circle (green circle). The general orbit’s projected ellipse (black) is centered at the origin, and the projected ellipse axes (purple) define the x^¯dr and y^dr axes. (Note that the projected ellipse semimajor axis has been derotated such that the star is always located in the first quadrant.) The separation circle center (orange dot) is the star’s location relative to the orbit’s projected ellipse. The xh points are the quartic solutions, where subscripts are reordered in ascending distance from the star’s x position. This ordering means that x2 must always occur in the fourth quadrant, x0 may occur in either the first or second quadrants, and x1 may occur in either the third or fourth quadrants. For 99.992% of KOE sampled from the SAG13 population that produce four intersections, the x3 intersection occurs in the first quadrant (a). The other 0.008% of KOE result in the x3 intersection occurring in the fourth quadrant (b).

JATIS_7_3_037002_f007.png

For KOE with four extrema and smin<sWA<slmin, we know that there will be two intersections on the same y side of the ellipse as the star (quadrants 1 or 2 in the dr frame). Of the four quartic solutions to Eq. (36) that we have to choose from, we know that x0 is one of them. The other solution could either be x1 if I(x1)<109 or x3 if I(x1)>109.

For KOE with four extrema and slmax<sWA<smax, we know that there will be two intersections on the opposite x side of the ellipse as the star (quadrants 2 and 3 in the dr frame). In all cases, x0 and x1 are the intersection solutions. This is because the first two solutions have the largest term in the quartic solution [Eqs. (91) and (92)]. x1 is slightly smaller than x0 because it subtracts the second largest term. The relative magnitudes of x0 and x1 determine that y0 occurs in the same side of the ellipse as the star and y1 must occur on the opposite side of the ellipse as the star. Therefore, (x0, y0) occurs in quadrant 2, and (x1, y1) occurs in quadrant 3.

For KOE with four extrema and slmin<sWA<slmax, we know that there will be four intersections. Unlike in Sec. 2.1.2 and the rest of this paper where x0 through x3 are ordered as in Appendix K, we order xh based off Δxh=|xhx*|. We order the quartic solutions from x0 to x3 such that x0 is where min({Δxh    h{0,1,2,3}}) and x3 is where max({Δxh    h{0,1,2,3}}). These newly ordered Δxh correspond to those shown in Fig. 7. The (x2,y2) intersection occurs in either quadrant 1 or 2, but it is always above and to the left of the star in the dr frame and has the second largest Δxh component. The (x3, y3) intersection occurs in either quadrant 1 or 4, but it always has the largest Δxh component. We resolve the sign of y3 by testing it in both quadrants. In >99.992% of cases, we assign y3 to quadrant 1 as in Fig. 7(a), but for a minority of cases, its correct assignment is quadrant 4 as in Fig. 7(b). The (x1,y1) intersection always occurs in quadrant 4. The (x0,y0) intersection occurs in either quadrant 3 or quadrant 4, but it is always below and to the left of the star in the dr frame and has the smallest Δxh component.

For KOE with two extrema and smin<sWA<smax, we know that there will be two circle-ellipse intersections. The KOE determine where the star is located in the dr frame. The location of the star in the dr frame relative to the vertices of the projected ellipse [(0,bp), (0,bp), (ap,0), and (ap,0)] determines the star-vertex separation ordering and subsequently to which quadrants the two intersection solutions belong. Instead of calculating the star-vertex separation for each orbit, we divide the first quadrant into four regions that specify the star location type [types 0 to 3 as indicated in Fig. 8(a)]. This means that any KOE with the star in location type 2 has the associated star-vertex separation ordering.

Fig. 8

The regions identifying the star type and which vertices are closest to the host star. (a) The first quadrant of a projected ellipse (black curve) with the semimajor and semiminor axes (purple) and three dashed lines dividing the quadrant into four regions defining the separation ordering. The pink dashed line represents the line of points equidistant from (0,bp) and (ap,0). The gray dashed line represents the line of points equidistant from (ap,0) and (0,bp). The turquoise dashed line represents the line of points equidistant from (0,bp) and (ap,0). The yellow dashed line represents the line of points equidistant from (ap,0) and (0,bp). We say that the star in (a) is a type 2 star and has the associated separation ordering. (b)–(e) Color each region of the ellipse, identifying which vertex is closest, second closest, third closest, and fourth closest. The star in (a) is type 2 and has the top vertex as the closest as seen in (b). These plots are based off the projected ellipse of a planet with a=0.40  AU, e=0.23, i=0.69  rad, Ω=3.49  rad, and ω=5.64  rad.

JATIS_7_3_037002_f008.png

Using the equidistant lines between ellipse vertices, we divide the first quadrant into four regions (types 0 to 3) as shown in Fig. 8(a). Regions 0 and 2 are divided by the line defined by

Eq. (37)

yapx,bpy=apbpx+ap22bpbp2,
where a projected ellipse and star is of type 0 if y*>yapx,bpy(x*) and one of types 1, 2, or 3 otherwise depending on the other equidistant lines. Similarly, regions 1 and 2 are divided by the line defined by

Eq. (38)

yap+x,bpy=apbpx+ap22bpbp2,
regions 2 and 3 are divided by the line defined by

Eq. (39)

yap+x,bp+y=apbpxap22bp+bp2,
and the star-ellipse classification type is calculated similar to type 0.

An example star is shown in Fig. 8(a). This star location is type 2 because it is in the region bounded by the three dashed lines (gray, pink, and teal). This type 2 star has the separation ordering indicated in Fig. 8(a) and Table 1. Let us consider a sWA separation circle such that sapx*,y*<sWA<sx*,bp+y*. We, therefore, know that the two intersections must occur in quadrant 2 and quadrant 4 of the dr frame.

Table 1

Separation order from smallest to largest by star location type.

TypeConditionSeparation order from smallest to largest
FirstSecondThirdFourth
0sap+x*,y*<sx*,bp+y*sx*,bpy*sapx*,y*sap+x*,bpsx*,bp+y*
1sx*,bp+y*<sapx*,y*sx*,bpy*sx*,bp+y*sapx*,y*sap+x*,y*
2sapx*,y*<sx*,bp+y*
sx*,bp+y*<sap+x*,y*
sx*,bpy*<sapx*,y*
sx*,bpy*sapx*,y*sx*,bp+y*sap+x*,y*
3sapx*,y*<sx*,bpy*sapx*,y*sx*,bpy*sx*,bp+y*sap+x*,y*

The distances sx*,bp+y*, sx*,bpy*, sap+x*,y*, and sapx*,y* in Fig. 8(a) are the distances of the star to each of the ellipse vertices and are calculated by

Eq. (40)

sx*,bp±y*=x*2+(bp±y*)2,

Eq. (41)

sap±x*,y*=(ap±x*)2+y*2.
Figure 8(a) and Table 1 define orbit types sorted by vertex distances from smallest to largest based on star location type in the first quadrant of the dr frame.

Finally, after determining the correct number of intersection solutions and the proper quadrants to which these solutions belong, we rerotate and translate the intersection solution locations back into the 2D projection of the 3D orbit.

2.2.

Δmag Intersections

In this section, we present our method for calculating the values of ν on a planet’s orbit where the planet has a specific value of Δmaglim, called Δmag intersections. As in Sec. 2.1, to compute these solutions, we first need to calculate all Δmag extrema (Δmagmin, Δmagmax, Δmaglmin, and Δmaglmax). The process for calculating Δmag extrema is detailed in Appendix C, but briefly included as follows:

  • 1. express Δmag as a polynomial in cos(ν);

  • 2. find the values of ν and Δmag for all polynomial roots; and

  • 3. remove invalid and duplicate solutions.

The general process for calculating the true anomalies where a planet has Δmaglim is similar to the extrema-finding process but with some minor modifications. After calculating the Δmag extrema, we determine how many Δmag intersections a given orbit should have with a particular Δmaglim value. If Δmaglim<Δmagmin or Δmaglim>Δmagmax, there are no intersections. If Δmagmin<Δmaglim<Δmaglmin or Δmaglmax<Δmaglim<Δmagmax, then there are exactly two intersections. If Δmaglmin<Δmaglim<Δmaglmax, then there are exactly four intersections. When the orbit does not contain local Δmag extrema, there are just two Δmag intersections if Δmagmin<Δmaglim<Δmagmax.

Knowing the number of solutions to expect, we follow the same steps as above: finding a governing polynomial equation, solving for its roots (represented by the ordered set x), and filtering out the relevant solutions. We additionally throw out any solutions with large errors from the input Δmaglim. The full process outline is included in Appendix D and discussed in depth below.

We start with the definition of Δmag given in Eq. (1). Although this expression contains multiple terms that are fully defined by an orbit’s KOE, it is also a function of other planet properties, including the planet’s radius, geometric albedo, and phase function. To make the mathematical development presented below tractable, we assume that the quasi-Lambert phase function17 is a sufficient approximation of any planet’s phase function. In general, the quasi-Lambert phase function is a better representation of the Earth’s phase function than the Lambert phase function.

This quasi-Lambert phase function is given by

Eq. (42)

ΦL(β)=cos4(β2).
We take advantage of this function’s form by substituting the half-angle formula:

Eq. (43)

cos(β2)=1+cos(β)2.
At the same time, from the orbital geometry defined in Fig. 3 and Eq. (3), we write

Eq. (44)

β=cos1(sin(i)sin(ν+ω)),
where the final term is expanded by the angle addition formula as

Eq. (45)

sin(ν+ω)=sin(ν)cos(ω)+cos(ν)sin(ω).
Substituting in β from Eq. (44) expanded with Eq. (45) into Eq. (43) allows us to reduce the order of the fully substituted Eq. (42). Note that this expression for β depends on making the approximation that the observer-star and observer-planet vectors are parallel, which introduces minor error given the large distances to even the nearest stars.18

After making these substitutions, as well as taking Eq. (4) for the planet-star distance term |r̲k/i|, simplifying, and collecting all of the non-orbital planet parameters on the left side, we find

Eq. (46)

a2(1e2)2pR2102.5Δmag=14(ecos(ν)+1)2(1+sin(i)cos(ω)sin(ν)+sin(i)sin(ω)cos(ν))2.
We now have an expression that isolates terms containing ν and can be decomposed into a numerically solvable polynomial. To improve solving efficiency and determine which subset of planets should have ν solutions for a given Δmag, we need to first calculate the Δmag extrema over the full orbit. The outline of the process for calculating these Δmag extrema is included in Appendix C.

To calculate these Δmag extrema, we first find the derivative of Eq. (46) and multiply by two (for simplification purposes) to get

Eq. (47)

0=e2sin2(i)sin3(ν)cos(ν)cos2(ω)3e2sin2(i)sin2(ν)sin(ω)cos2(ν)cos(ω)2e2sin2(i)sin(ν)sin2(ω)cos3(ν)+e2sin2(i)sin(ν)cos3(ν)cos2(ω)+e2sin2(i)sin(ω)cos4(ν)cos(ω)2e2sin(i)sin2(ν)cos(ν)cos(ω)3e2sin(i)sin(ν)sin(ω)cos2(ν)+e2sin(i)cos3(ν)cos(ω)e2sin(ν)cos(ν)esin2(i)sin3(ν)cos2(ω)4esin2(i)sin2(ν)sin(ω)cos(ν)cos(ω)3esin2(i)sin(ν)sin2(ω)cos2(ν)+2esin2(i)sin(ν)cos2(ν)cos2(ω)+2esin2(i)sin(ω)cos3(ν)cos(ω)2esin(i)sin2(ν)cos(ω)4esin(i)sin(ν)sin(ω)cos(ν)+2esin(i)cos2(ν)cos(ω)esin(ν)sin2(i)sin2(ν)sin(ω)cos(ω)sin2(i)sin(ν)sin2(ω)cos(ν)+sin2(i)sin(ν)cos(ν)cos2(ω)+sin2(i)sin(ω)cos2(ν)cos(ω)sin(i)sin(ν)sin(ω)+sin(i)cos(ν)cos(ω),
which is a function of the sin(ν) and cos(ν) terms. We now fully expand this expression and substitute in

Eq. (48)

sin(ν)=1cos2(ν)
to get an expression in cos(ν) only. We define x=cos(ν), isolate the 1x2 term, square both sides, and expand to get an eighth degree polynomial in x of the form

Eq. (49)

0=A2x8+B2x7+C2x6+D2x5+E2x4+F2x3+G2x2+H2x+I2.
The coefficients of this expression are included in Eq. (75) in Appendix H.

Although eighth degree polynomials do not have analytical solutions, a numerical root solver can determine the eight roots of this function denoted as the ordered set x. We discard any xp>1    p{08}, xp<1    p{08}, or solutions with large imaginary components. We then calculate the remaining true anomaly solutions by ν0=cos1(x) and ν1=2πν0. Of the remaining valid solutions, we identify whether each solution is an extremum by evaluating whether Δmag(ν0±δν) are both larger or both smaller than Δmag(ν0). If identified as a potential extremum, the smallest and largest extrema are assigned to Δmagmin and Δmagmax. The remaining extrema are checked for duplicates, which are identified by solutions with (ν, Δmag) values close to existing extrema. If any solutions are remaining, an additional check and assignment is then made for the local extrema. Through this process, a solution identified in ν0 has the associated solution in ν1 removed.

To calculate the Δmag intersections, we apply the same process for turning Eq. (46) into a polynomial as in the Δmag extrema calculation. The full outline for calculating the Δmag intersections is given in Appendix D. By following this process, we arrive at

Eq. (50)

ξ=A3x8+B3x7+C3x6+D3x5+E3x4+F3x3+G3x2+H3x+I3,
where ξ is the collection of constants on the left side of Eq. (46). The coefficients of this polynomial are given in Eq. (76) of Appendix I.

We again use a numerical root solver to find the roots of this function x and again discard all xp>1   p{08}, xp<1    p{08}, and solutions with large imaginary components. We then calculate the true anomalies of all remaining solutions by ν0=cos1(x) and ν1=2πν0. We then evaluate Δmag0=Δmag(ν0) and Δmag1=Δmag(ν1) and remove solutions where |Δmag0Δmaglim|>0.01 and |Δmag1Δmaglim|>0.01. (Note that 0.01 is <±0.08% error on Δmag.) We do an iterative process of selecting (ν, Δmag), which are unique (not duplicate solutions) and are closest to the expected Δmaglim, until we have the expected number of solutions. In the majority of cases in which solutions exist, there are generally only two viable solutions, the assignment of which is simple. In some cases, the solution selection is more ambiguous as double roots are possible.

2.3.

ν from X and Y

Sections 2.1.2, 2.1.3, and 2.2 give locations of extrema at intersections in the plane of the sky of the form (x,y), but we need to know the true anomalies of the orbit where these intersections occur.

We start with X and Y in Eq. (3) and solve for (1+ecos(ν))/(a(1e2)), resulting in

Eq. (51)

1+ecos(ν)a(1e2)=1Y[sinΩcos(ω+ν)+cosΩsin(ω+ν)cosi],
and

Eq. (52)

1+ecos(ν)a(1e2)=1X[cos(Ω)cos(ω+ν)sin(Ω)sin(ω+ν)cos(i)].
In both of these expressions, we substitute angle addition formulas of sin(ω+ν) and cos(ω+ν) and subsequently set the two equations equal to each other.

We set Eqs. (51) and (52) equal to each other, expand, isolate a cos(ν) term and sin(ν) term, solve for ν, and rearrange to get

Eq. (53)

ν=tan1[XYsinΩcosωXYcosΩcosisinω+cosΩcosωsinΩcosicosωXYsinΩsinω+XYcosΩcosicosω+cosΩsinω+sinΩcosicosω].
We now have an analytical expression for ν solely as a function of the KOE, X, and Y of a particular point on the orbital ellipse.

By combining the original X and Y equations to solve for ν analytically, we have created two potential solutions at ν and ν±π. One of these is the correct ν value and the other is not. To calculate the correct intersection point, we calculate the separations at both ν and ν±π to find absolute error in s and use the smaller error of the two.

2.4.

Calculate t from ν

We have ν for the locations where s extrema, Δmag extrema, sWA-orbit intersections, and Δmag intersections occur, but we need them in terms of time. We calculate eccentric anomaly E of these events directly from ν as

Eq. (54)

E=tan1(1e2sinνe+cosν),
which gives the corresponding time

Eq. (55)

t=Eesin(E)T2π,
where T is the orbital period of a general planet.

2.5.

Converting from Star to Star

The planet-star intersection points and visibility ranges can be saved as either true anomalies or as specific times for a reference star. It makes the most sense to store the values as times and use 1M as the reference time. The times are scaled to any star mass M by

Eq. (56)

T=TMM.

2.6.

Calculating Completeness

We calculate integration time adjusted completeness using the aggregated fraction of time that planets are detectable. Using the methods described in Sec. 2 on the instrument’s photometric visibility limit (Δmaglim) and astrometric visibility limits (IWA and OWA), we calculate the specific times that the planet enters or exists the instruments visibility limits. By collecting these times and inspecting intermediate test points, we identify the time windows in which any planet is detectable by the instrument. Given an integration time (tmax) required to reach some SNR consistent with a clear detection (typically a value > 5), we throw out all visibility time windows less than tmax and discount all other visibility windows by tmax. Dividing the sum of all integration time discounted visibility windows for a planet by its orbital period gives us the fraction of time that a planet is detectable by the instrument. Aggregating the fractions of time that planets are detectable by the instrument gives us integration time adjusted completeness. In mathematical form, this is given as

Eq. (57)

Ctmax=1N(  k  j(δtj,ktmaxUk)Tk).
Here δtj,k is the j’th time window larger than tmax in which the k’th planet is visible, Uk is a boolean that indicates that the planet is always visible 0 or at least sometimes visible 1, N is the total number of planets, and Tk is the orbital period for the k’th planet.

We also calculate the completeness for an Earth-like exoplanet of the Earth-like exoplanets subpopulation:

Eq. (58)

C,tmax=1N(  k  j(δtj,ktmaxUk)Tk),
where N is the total number of Earth-like planets simulated.

Similarly, this capability can be extended to the entirety of the Kopparapu classification scheme.4 Our method uses orders of magnitude fewer exoplanets and an equivalent memory but fills in the Δmag space using an order of magnitude better accounting with strategically calculated true anomalies.

2.7.

Calculating Dynamic Completeness

The methods presented in this paper can also be used to calculate dynamic completeness.13 Dynamic completeness extends the original concept of completeness by considering the fraction of planets observable on subsequent observations of the same target. For the original formulation of completeness, which relies on simulating a cloud of planets, this requires propagating every simulated planet along its orbits. Dynamic completeness is frequently used to compute the fraction of planets in a population that are initially undetectable but become detectable upon a second observation some time later.

We define P as the unordered set of all planets, Pdetected,1 as the set of planets detected at observation time one, and Pundetected,1 as the set of planets not detected at observation time one. To find the sets of detected planets at the first observation, we start by generating a random observation time for the k’th planet (tstart,k) between 0 and Tk. We then determine which planets are within the instrument’s visibility limits at tstart,k to create Pdetected,1 and its complement Pundetected,1.

For a subsequent observation some time later (twait), we find the time past periastron at which the observation occurs by [mod(tstart+twait,Tk)] and determine which planets are within the instrument’s visibility limits at that time. This defines the unordered set of planets detected at observation time two as Pdetected,2. Its complement is Pundetected,2, the set of planets not detected at observation time two.

Dynamic completeness for the second observation of a target star is given by the fraction of planets undetected in the initial observation but detected on the second one:

Eq. (59)

C2=1Nn(Pundetected,1Pdetected,2),
where n(P) is a function giving the number of elements in set P and N is the total number of planets.

Dynamic completeness of the m’th visit is similarly given by the fraction of previously undetected planets that are detected on that visit:

Eq. (60)

Cm=1Nn((j=0m1Pundetected,j)Pdetected,m).

3.

Results

3.1.

ν from s

We apply the methods developed Sec. 2 to calculate the planet-star separation extrema, the true anomalies at which they occur, the times past periastron at which the extrema occur, the true anomalies where the separation circle intersects with the projected orbital ellipse, and the times at which these intersections occur for a planet orbit. Figure 9 shows the projected separation of a sample orbit with global and local extrema calculated via Sec. 2 methods. This figure also demonstrates the ability of our methods to find all true anomalies (and times) when the projected separation takes a specific value (in this case sWA=1  AU).

We now determine the accuracy of the methods implemented. To do this, we calculate the true anomalies of sWA-orbit intersections for 105 orbits using the method in Sec. 2.1.3 and sWA=1  AU. We then calculate the planet-star separation by substituting these true anomalies back into Eq. (2) and plot the error between these separations and the input sWA in Fig. 10. Of the 105 planets orbits simulated, 25,000 orbits produced two or more intersections. The largest intersection error observed is <106. Machine precision error is 1016, and the square root of this is 108, which indicates that we have approached machine precision in our method.

3.2.

ν from Δmag

We apply the methods developed Sec. 2 to calculate the Δmag extrema, the true anomalies at which they occur, the times past periastron at which the extrema occur, the true anomalies where the Δmag intersections occur, and the times at which these Δmag intersections occur for a planet orbit. Figure 11 shows the Δmag of a sample orbit with global and local extrema calculated via Sec. 2 methods. This figure also demonstrates the ability to find all true anomalies (and times) when the Δmag takes a specific value (in this case Δmaglim=25.0).

To check the error in true anomaly produced by the Δmag intersection method in this paper, we compare the ν of Δmag intersections calculated with alternative numerical solving methods. The method presented in this paper finds Δmag intersections of 106 planets within 6.3  s. The first error checking method uses a cubic spline fit to 300 (ν, Δmag) points along each planet’s orbit. We then subtract Δmaglim=29 (our test point) from the spline and find the roots. This univariate spline root solving method is capable of executing in 419 s on the 510,120 planets in the population that produced two intersections, a rate of 8.1×104  s per planet. The univariate spline root solving method is 100× slower than the Δmag intersection method in Sec. 2.2. The separation error between the univariate spline method and the Δmag intersection method is <6×105 for >99.98% of planets and <104 for the other 0.02%. These high-error targets are planets with low variation in Δmag(ν) orbits. The second-error checking method solves for the (ν, Δmag) intersections using a numerical minimization method on a random subset of 104 planets. This method is far more inefficient, taking 1646 s on 104 planets with two intersections at a rate of 0.1646 s per planet. Using a minimization function allows us to determine the error in Δmag of an intersection point to within 108. This numerical minimization method independently confirms the accuracy of solutions to the univariate spline roots method on the limited number of targets tested. We plotted the normalized frequency of true anomaly error of both the numerical minimization method and univariate spline root solving method compared against the Δmag intersection method in Fig. 12. Note that the total number of incidences of a given error is normalized by the bin width and total number of targets, so the non-normalized frequency of the 3×1011 bin is incredibly large compared with the 5×104 bin. Both methods indicate that the resulting true anomalies of the Δmag intersections are within 104  rad of each other.

Although the Δmag intersection method featured in this paper is orders of magnitude faster than the univariate-spline-roots method and just as accurate, the quasi-Lambert phase function required for its use is not the best phase function for all planets; however, it does fit Earth-like planet phase function quite well. At a substantial time cost, the univariate-spline-roots method can be used for planets with any phase function.

3.3.

Convergence and Validation

We test convergence of our method for calculating completeness by repeatedly calculating completeness on 105 planets. We test convergence of Brown’s method by repeatedly calculating completeness over a set of a logarithmically increasing number of planets. We use the same SAG13 planet population parameters from Ref. 1 but make the substitution of a quasi-Lambert phase function to define the underlying planet population. Since this technique will be most relevant to future highly capable telescopes, we are using the HabEx IWA of 0.045 arc sec, OWA of 6 arc sec, and upper Δmag limit of 25 on a star that is 10 pc away.8 This resulting instrument visibility limits and planet population cover a well populated region of the Δmag versus s joint probability density function [see Fig. 1 in Ref. 1 for a comparable example of the joint probability density function]. Convergence of the two completeness methods is demonstrated in Fig. 13.

Both methods presented in Fig. 13 demonstrate similar percent convergence to their own converged means. Completeness repeatedly calculated using the method in Sec. 2.6 with 105 planets has a mean of 0.25785 with a standard deviation of 0.0010 for 4.28×108 planets. In comparison, the Brown completeness method converges to 0.25783. This error in the converged mean of the two methods is consistent with the completeness error identified in Fig. 14. We expect that completeness calculated using the method in this paper will have comparatively better convergence in sparsely populated regions of the Δmag versus s joint probability density function (see Fig. 1 in Ref. 1).

Since completeness for an individual planet is the fraction of time that a planet is visible to the instrument, we validate individual planet visibility windows using test points. To calculate a ground truth fraction of time that a planet is visible, we create 105 test points, evenly spaced in time, and calculate both the planet-star separation and planet-star Δmag at each of theose test points. We then determine whether each point is within the visibility limits of the instrument. The fraction of visible points is the fraction of time that a planet is visible. We repeated this for 25,000 planets due to computational cost of the test point method. The histogram of error in completeness for individual planets calculated using the test point method and method in this paper is shown in Fig. 14. The maximum error in completeness is observed to be near 5×105, which is consistent with the converged error between Brown completeness and integration time adjusted completeness.

The test point method that we used to validate integration time adjusted completeness calculations took over 52 h on only 25,000 planets (less than a quarter of the planets needed to compute completeness using integration time adjusted completeness). The computation time for the integration time adjusted completeness calculation on 105 planets is 21.34 s with a standard deviation of 0.4 s over 1000 calculations. Brown completeness, including the joint probability density function generation, has a mean execution time of 3.19 s and standard deviation of 0.13 s over 1000 calculations. This execution time of all of these methods scales linearly with the number of planets. Some room for optimization exists in integration time adjusted completeness.

3.4.

Completeness Versus Integration Time

To evaluate the effects of integration time on completeness, we calculate integration time adjusted completeness for various integration times, star distances, and planet populations (SAG13 and Earth-like planet population). Figure 15 shows the integration time adjusted completeness for the SAG13 planet population1 and the Earth-like planet subpopulation defined in Appendix L. The decrease in completeness for longer integration times is most prominent for nearby stars for both populations. As expected, longer integration times decrease completeness.

For the assumed observatory parameters, the Brown completeness of an Earth-like population for a target 5 pc away is 0.583. As we increase integration time to 1, 2, and 5 days, the associated integration time adjusted completeness decreases by 0.63%, 1.27%, and 3.15%, respectively. Recalculating Brown completeness for a target at 25 pc in Fig. 15(b) gives 0.583. As we increase integration time to 1, 2, and 5 days on this 25 pc target, the associated integration time adjusted completeness decreases by 0.989%, 1.97%, and 4.92%, respectively. Our integration time adjustment of completeness shows Brown completeness overestimates exoplanet yields for any observation. Brown completeness applied to stars farther away results in a substantial overestimation of exoplanet yield.

Technically, the Δmaglim used as the upper limit for calculating completeness is a function of integration time and will approach a theoretical upper limit as the noise floor is reached. By calculating Δmaglim using Eq. (12) of Ref. 1, we can calculate completeness at each integration time to get the completeness versus integration time curve. We created a 4-m telescope as in Appendix M and evaluated completeness at each integration time to create an example demonstrating how Brown completeness and integration time adjusted completeness vary with integration time in Fig. 16. We can see that the completeness of both methods tracks very closely until they diverge. Brown completeness approaches its asymptotic limit as expected, whereas integration time adjusted completeness approaches a maximum and begins decreasing past 0.2  day.

3.5.

Dynamic Completeness

We compute dynamic completeness for the example test case in Fig. 1 of Ref. 13 using the method described in this paper and replicating the approach of the original work. This computation is only for a second epoch. We replicated the original work by sampling a large number of planets, finding the planets initially visible, propagating these planets to some time past initial observation, and determining what fraction of these planets were not detected the first time but were detected the second time. We replicated Brown’s work using both the Lambert phase function (orange line in Fig. 17) to match the original work as well as the quasi-Lambert phase function (red line in Fig. 17).

We compared the dynamic completeness computation time for 1000 dynamic completeness calculations testing 1000 individual points in time past the initial observation. Brown’s dynamic completeness had an average execution time of 34.62 s with a standard deviation of 0.99 s. Calculating dynamic completeness using the method presented in this work resulted in an average execution time of 4.757 s with a standard deviation of 0.036 s. The dynamic completeness calculation presented in this paper is 7× faster than the traditional Brown completeness method, enabling its use in optimization.

4.

Discussion

4.1.

Convergence

Both methods of calculating completeness suffer from sparse sampling of the exoplanet population parameter space. Brown completeness is most affected as it needs planets sampled over Ω, ω, a, e, i, ν, p, and R. The current implementation of our method needs the same parameters except for ν. We are therefore surprised that Brown’s method and our method share such similar completeness convergence. We hypothesize that the similar convergence results are because the integration bounds used for calculating completeness are over a dense region of the joint probability density function. If we examined a more sparse region of the joint probability density function such as high Δmag or larger separation (where the larger, lower occurrence rate planets occur), we expect our completeness methods to have marginally better convergence for low numbers of planets. Regardless, integration time adjusted completeness easily reaches completeness errors below 103 (translates to 0.8% where C=0.12, the largest completeness of an observation in Ref. 1), which is a sufficient error for use in optimization.

4.2.

Reducing Parameter Spaces

The integration time adjusted completeness method implemented in EXOSIMS is not cached for computational efficiency like the Brown completeness is. Although we are able to store the joint probability density function of Δmag versus s with Brown’s method, integration time completeness requires the KOE of each simulated planet to be stored. The curse of dimensionality prohibits us from finely sampling the entire subspace of exoplanets, making the caching of integration time adjusted completeness prohibitive, but not all our methods require all 6 KOE, p, and R. So, in the future, it may be possible to create a representative subsampling of planets weighted by occurrence rate based on the parameters needed for different calculations. For example, s extrema, sWA-orbit intersections, and Δmag extrema calculations only require ω, a, e, and i (Note that the ν locations of the Δmag extrema are independent of p and R, but the magnitude of Δmag extrema are dependent upon them). The photometric property p×R could then be sampled and independently combined with individual (ω, a, e, i) combinations. The primary benefit is that, given some instrument parameters and star properties, we could determine the subset of parameters that can physically be observed prior to calculating completeness, thus making the per-star completeness calculations more efficient.

4.3.

Limitations

The approach we implemented in this paper did not vary the Δmaglim as we varied the integration time. If properly executed, Δmaglim should increase with increased integration time. However, calculating the planet true anomaly intersections with a given Δmaglim is the most expensive calculation, so repeatedly calculating this to convergence is not desirable.

The Δmaglim of coronagraph designs are working angle dependent and therefore separation dependent, but we assume that this limit is a fixed quantity. If there is substantial variation in the limiting Δmag, it is better to use the conservative value with this method. If a higher level of predictive accuracy is desired, these methods could be used to ascertain if any intersection with the Δmaglim exists and subsequently another method could be used on the subset of planets with known intersections.

4.4.

Impact of the Integration Time Adjusted Completeness

As we showed in this paper, integration time adjusted completeness and Brown completeness converge to the same value to within 0.00002 when evaluated at tmax=0 day. The integration time adjusted completeness for an observation of an Earth-like planet population on a 1-M star 5 pc away and integration time of 1 day is 0.64% lower than the comparable Brown completeness. For reference, the Roman target list in Table 9 of Ref. 1 has a maximal integration time of 1.71 days but most are <0.6 day. This means that we could expect an average reduction in yield below 0.64% when observing a population of Earth-like planets around Sun-like stars. This integration time completeness adjustment is within the 3.19%, 3σ, margin of error from 1000 Monte Carlo simulations of the cycle 6 Roman in Ref. 1 and cannot be considered statistically significant. However, a 4-day integration time on a Sun-like star 25 pc away observing an Earth-like population has a decrease in completeness above the Ref. 1 3σ margin of error. However, the threshold of statistical significance should not deter the widespread use of integration time adjusted completeness as the adoption of this method can shore up the differences between completeness-based yields and simulation-based yields. We can also say that integration time has a muted effect when observing stars that are farther away. This can most likely be attributed to the increase in smin and resulting decrease in the total time-fraction that smaller period planets spend within the visible limits of the telescope. Big planets with the orbital radius of Jupiter will move more slowly and be less affected by integration times.

In Appendix M, we optimized an exoplanet direct imaging mission maximizing single-visit Brown completeness yield and included the resulting Design Reference Mission (DRM) in Table 2. The resulting DRM has a Brown completeness yield of 387.16 exoplanet detections in a single-visit detection survey on average, but the integration time adjusted completeness yield expects only 354.67 exoplanets on average. This means that Brown completeness for the particular instrument parameters optimized over the SAG13 planet population with the particular mission parameters overpredicts the actual exoplanet yield by 32.49 exoplanets on average, 9.16% more than the integration time adjusted completeness yield. When calculating completeness, we are careful to scale the completeness limits of integration by the star’s luminosity. For the integration time adjusted completeness method, we additionally scale the planet’s orbital periods based off the mass of the host star. The percentage difference above and beyond that expected from Fig. 15 can be found by looking at these two adjustments applied to each star and the target list in Table 2. The average star in the target list has a larger mass and a brighter luminosity than that of the Sun. The brighter luminosity results in a smaller smin and smax, which serves to incorporate more smaller semimajor axis planets into the completeness calculations, and the larger star masses result in shorter periods, meaning that the planet visibility windows all decrease in duration.

Integration time adjusted completeness is crucial for determining the ability of an Earth-like planet to be spectrally characterized. A spectral characterization with a coronagraph could take between a few days and 60 days. Because of the long integration time, the planners of a HabEx use the “characterization completeness” of 10% and maximum integration time of 60 days as a filter on stars to consider observing.8 The calculation of this critical filter could be substantially improved by considering integration time adjusted completeness if the spectrum of the planet must be taken all at once (i.e., observation spanning multiple weeks) and not spread across multiple epochs. Just because a planet can be detected does not mean that it can be spectrally characterized.

4.5.

Dynamic Completeness and Computation Cost

The greatest benefit of using the methods in this paper to calculate completeness is the marginal additional cost of calculating dynamic completeness. Generally, dynamic completeness requires the computation of true anomalies from time, which is prohibitive for >105 orbits at >100 different times in the future. With our methods, calculating visibility windows allows us to use boolean operations to compute dynamic completeness about 7× faster than Brown’s dynamic completeness method. The computational cost of calculating completeness using the method described in this paper is an order of magnitude larger than Ref. 2, but it can use orders of magnitude fewer planets to do so. Unlike Brown’s method, within the computation time of our method, we also get additional desirable information about the detectable planets such as their s extrema and Δmag extrema.

4.6.

Revisiting the Same Exoplanet

Another limitation present in the planning of exoplanet direct-imaging missions is the telescope keep-out angles. Figure 3 of Ref. 1 contains a keep-out map for a subset of the DRM created in that paper. The smallest percent of time that a target star is visible for the Roman is nominally 28%. Due to symmetry of the keep-out region, this translates into two separate time windows of visibility of 51 days. Instead of considering the integration time an integration time input, it can also be considered a revisit time input for determining the probability of being able to observe a planet twice in the same target-star visibility window.

4.7.

Exoplanet Classification

The underlying methods in this paper are used to find locations along a planet’s orbit where s and Δmag intersections occur. The methods in this paper, with some modification, can also be used to probabilistically classify an exoplanet subtype.4 If an exoplanet is detected with a particular (s, Δmag) and an uncertainty region of s±σs and Δmag±σΔmag, then the methods in this paper can be applied to each of these four bounding lines. By finding the average orbital time-fraction that exoplanets of each type spend in the bounding uncertainty box, we can find the probability that the exoplanet belongs to a specific exoplanet subtype. This requires additional work beyond the scope of this paper.

5.

Conclusion

We have demonstrated an accurate method for calculating integration time adjusted completeness and its adaptation to calculating dynamic completeness. In the process, we also created fast and accurate methods for calculating the true anomalies where a planet’s orbit has specific values of projected separation Δmag and their extrema. We demonstrated how to use these methods to calculate a more accurate integration time adjusted completeness using the fractions of time that a planet is detectable by an instrument. We demonstrate that traditional methods of calculating completeness overestimate the number of observable planets because they do not subtract the integration time used in observing the target. For a Sun-like star at 25 pc with 1-day and 5-day integration times, integration time adjusted completeness of Earth-like planets is reduced by 1% and 5%, respectively. We applied integration time adjusted completeness to a target list optimized using the Brown completeness method and found that Brown completeness overestimated yields by 9.61%. We also demonstrated that our methods can be used to quickly calculate dynamic completeness for determining the optimal time to revisit a target star.

6.

Appendix A: Common Notation

There are many common variable forms used in this paper that are simply summarizable. Any variable x_ refers to a 3D vector in X, Y, and Z coordinates of something. We use x to indicate an array of variables, specifically used when referring to multiple roots of a polynomial. |AB| refers to the length of line segment AB. Line segment AB is treated as a vector. C is a projection of point C.

There are multiple different subscripts with different meanings used in this paper. A subscript with xi refers to the index of a host star (out of some whole, non-descript, star catalog). A subscript xk refers to an individual planet out of the whole large set of planets. Subscripts of xmin, xmax, xlmin, and xlmax are descriptors on the individual variable x that indicate that the variable is associated with the minimum, maximum, local minimum, or local maximum, respectively. When solving for the distance between points inside an ellipse and the vertices or co-vertices of that ellipse with semimajor axis a, we use the shorthand notation of the form sa+x,y. [This is specifically the distance between point (x, y) and (a, 0).]

We denote coefficients of the polynomials of the four methods that solve the quartic in this paper as A#.

In Fig. 4, we reference many points on the 3D elliptical orbit and 2D projection of this orbit into the plane of the sky. Points on the original 3D ellipse are labeled B, H, P, C, K, O, D, A, B, H, P, C, K, O, D, and A.

In Appendix K, we use p0 to p11 as intermediate constants for simplifying the full quartic expression. P, D, and Δ (by itself and only in this section) are intermediate constants derived from quartic coefficients for determining the sign of the quartic solutions.

K0 to K9 are intermediate simplifying coefficients used in representing Eqs. (13)–(15) in Sec. 2.1.1. The full expansion of these equations is included in Appendix E.

Fig. 9

Planet-star separation of the planet is plotted in black. The separation extrema are indicated by diamonds, where the maximum is red, local maximum is yellow, local minimum is magenta, and minimum is teal. Lines are drawn at the minimum and maximum separations. For the input separation of s=1  AU, the green dots are the analytically calculated orbit intersections. (a) The separation versus true anomaly and (b) the separation versus time past periastron for a planet orbit with a=1.82  AU, e=0.09, Ω=3.37  rad, ω=4.86  rad, and i=1.25  rad.

JATIS_7_3_037002_f009.png

Fig. 10

sWA-orbit intersection error histogram calculated for a separation of sWA=1  AU for 105 planet orbits generated from the SAG13 planet population. Of these 105 planet orbits, 6201 orbits have two intersections with the sWA circle, and 17,952 have four intersections with the sWA circle. This results in a total of 84,210 planet-star intersections. After calculating and identifying true anomalies of intersections using the methods described in this paper, we evaluated the planet-star separation of each orbit at the true anomalies and calculated error from the input sWA.

JATIS_7_3_037002_f010.png

Fig. 11

(a) The Δmag of a planet (black line) plotted against ν and (b) time past periastron. Δmag extrema are indicated by diamonds, where the maximum is red, local maximum is yellow, local minimum is magenta, and minimum is teal. Separation minimum and maximum are indicated by horizontal lines. For an input Δmaglim=25 (green line), we calculated the specific true anomalies (green dots) where the planet’s Δmag intersects this line. The specific planet KOE are a=5.36  AU, e=0.56, ω=5.06  rad, Ω=0.69  rad, i=0.81  rad), p=0.3, and R=4R.

JATIS_7_3_037002_f011.png

Fig. 12

A histogram of ν(Δmag) error between the root solving method described in this paper and a root solver of a cubic spline fit method (purple) for 510, 120 planet orbits that produce two intersections (totaling 1,020,240 datapoints). We also compare the error between the root solving method described in this paper and an optimization method (blue) for 104 planet subset of the planet orbits that produce intersections. The optimization method investigates fewer planets because it is orders of magnitude more computationally expensive, but it is a fundamentally different approach to validating our method than a root solver.

JATIS_7_3_037002_f012.png

Fig. 13

Convergence of % error in completeness for varying numbers of planets used in its computation for Brown’s method (green) and the method presented in this work with tmax=0 day (purple). Completeness at the maximum number of planets is assumed to be the converged value of completeness of the respective methods. The converged mean of Brown’s method is 0.25783 compared with the converged mean of the method in this paper of 0.25785. The standard deviation of the method in this paper is 0.0010.

JATIS_7_3_037002_f013.png

Fig. 14

Per planet completeness error histogram between the integration time adjusted completeness method with tmax=0 day and the test point method for 25,000 planets.

JATIS_7_3_037002_f014.png

Fig. 15

Integration time adjusted completeness values of (a) SAG13 planet population and (b) Earth-like planet population with IWA = 0.045 arc sec, OWA = 6 arc sec, and Δmag=25 for varying integration times and star distances.

JATIS_7_3_037002_f015.png

Fig. 16

Brown completeness and integration time adjusted completeness versus integration time for HIP 32,279 assuming a mass of 1.564 M, luminosity of 7.12 L, and distance of di=3.51 pc. The assumed telescope is a 4-m monolith with IWA = 0.045 arc sec and OWA = 6 arc sec and Δmaglim computed using the instrument noise model as in Ref. 1.

JATIS_7_3_037002_f016.png

Fig. 17

Dynamic completeness assuming the same parameters as in Ref. 13 (Δmag=26, MHIP29271=1.103M, di=10.215 pc, OWA = 600 arc sec, IWA = 0.075 arc sec, 0.7La1.5L, 0e0.35, p=0.26, R=1R, and L=0.83L). In this work (blue), we assume a quasi-Lambert phase function, but in Ref. 13, a Lambert phase function is assumed. Work was done to replicate Brown’s original work (Brown Lambert, orange line) and replicate Brown’s work using the quasi-Lambert phase function (Brown quasi-Lambert, red line).

JATIS_7_3_037002_f017.png

7.

Appendix B: True Anomalies of s Intersection Process

The process outlined here is used in Sec. 2.1.2.

  • 1. Define KOE.

  • 2. r̲k/i from KOE.

  • 3. Orbiting foci F from KOE.

  • 4. Calculate 3D orbit ellipse center.

  • 5. Project 3D ellipse to 2D ellipse.

    • (a) Prove that 3D ellipses project to 2D ellipses.

    • (b) Project 3D ellipse center to 2D ellipse center.

    • (c) Project ellipse orbiting foci into plane of the sky.

    • (d) Project semiminor axis line of 3D ellipse to 2D ellipse.

    • (e) Project semimajor axis line of 3D ellipse to 2D ellipse.

    • (f) Calculate QQ construction line from semiminor axis and semiminor axis.

    • (g) Calculate projected ellipse semimajor axis and semiminor axis from OQ and OQ.

    • (h) Calculate projected semimajor axis angular offset from X axis of plane of the sky.

  • 6. Derotate the projected ellipse.

  • 7. Center the projected ellipse.

  • 8. Find minimum, maximum, local minimum, and local maximum of projected planet-star separation.

    • (i) Formulate projected planet-star separation equation.

    • (j) Reformat and set δs2/δxe equal to 0.

    • (k) Combine coefficients into standard quartic form.

    • (l) Use standard general quartic solutions.

    • (m) Take the absolute value and only real component of xe solutions.

    • (n) Calculate ye associated with each solution.

    • (o) Assign solutions to minimum, maximum, local minimum, and local maximum planet-star separations.

      • i. For all solutions (all real and only 2 real):

        • A. Calculate smin from x1 and assign (xmin,ymin) of quadrant 1.

        • B. Calculate smax from x0 and assign (xmax,ymax) of quadrant 1.

      • ii. All real solutions (where all solutions have I(y)<105, additionally assign).

        • C. The larger of s(x2) and s(x3) becomes slmax and assign (xlmax,ylmax) of quadrant 1.

        • D. The smaller of s(x2) and s(x3) becomes slmin and assign (xlmin,ylmin) of quadrant 1.

      • iii. Assign solution signs to proper quadrants.

  • 9. Find circle and projected derotated centered ellipse intersection points.

    • (p) Formulate circle-ellipse intersection equation.

    • (q) Reformat into standard quartic form.

    • (r) Use standard general quartic solutions.

    • (s) Classify intersection solutions.

      • iv. Inside outer separation bounds, smin<s<smax.

      • v. Inside local min/max separation bounds, slmin<s<slmax.

      • vi. Outside outer separation bounds (no intersections).

    • (t) Calculate intersection solutions for planets with smin, smax, slmin, and slmax.

      • vii. Two intersections on same y side of ellipse as the star, smin<s<slmin x={x3,I(x1)>1010x1,I(x1)<1010x0.

      • viii. Four intersections where, slmin<s<slmax

        • Calculate Δx=xk|x|, (k{0,1,2,3}).

        • Sort Δxk from min to max and rearrange the associated xk to match that order.

        • Same XY points is x3.

        • Same X opposite Y is x2.

        • Opposite X Same Y is x0.

        • Opposite X opposite Y is x1.

      • ix. Two intersections on opposite x side of ellipse as the star, slmax<s<smax.

        • x0 and x1.

        • y1 is opposite sign of |y|.

    • (u) Calculate intersection solutions for planets with only smin and smax.

      • x. Calculate projected ellipse quadrant separation bounds sx,b+y, sx,by, sa+x,y, and sax,y.

      • xi. Identify star location type and quadrant order.

        • Type 0 occurs where sx+a,y<sx,y+b. Smallest to largest order: sx,by, sax,y, sx+a,y, sx,y+b.

        • Type 1 occurs where sx,y+b<sax,y. Smallest to largest order: sx,by, sx,y+b, sax,y, sx+a,y.

        • Type 2 occurs where sax,y<sx,y+b and sx,y+b<sx+a,y and sx,by<sax,y. Smallest to largest order: sx,by, sax,y, sx,y+b, sx+a,y.

        • Type 3 occurs where sax,y<sx,by. Smallest to largest order: sax,y, sx,by, sx,y+b, sx+a,y.

8.

Appendix C: ν from Δmag Extrema process

Here we present the full outline of our process for calculating true anomalies where the planet’s orbit has a Δmag extrema. These steps are discussed in Sec. 2.2.

  • 1. Substitute into Eq. (1) components and expand until all terms are functions of KOE.

  • 2. Manipulate the Δmag equation until all ν terms are isolated on the right side.

  • 3. Take the derivative.

  • 4. Replace sin(ν) with 1cos(ν)2.

  • 5. Replace all cos(ν) with x.

  • 6. Isolate all square root terms to general form A21x2=B2.

  • 7. Square both sides.

  • 8. Find polynomial coefficients of x.

  • 9. Solve with root solver.

  • 10. Filter out x solutions with imaginary components (>107) (108 is approximately 1016, machine precision).

  • 11. Filter out invalid solutions, 1<x or 1>x.

  • 12. Compute ν0=cos1(x) and ν1=2πν0.

  • 13. Compute associated Δmag0 and Δmag1.

  • 14. Remove solutions that are not extrema.

  • 15. Assign solutions Δmagmax=min(Δmag0,Δmag1).

  • 16. Assign solutions Δmagmin=max(Δmag0,Δmag1)

  • 17. Remove duplicate solutions.

  • 18. Remove assigned solutions.

  • 19. Check if solution is extrema.

  • 20. Assign Δmaglmin and Δmaglmax.

9.

Appendix D: ν from Δmaglim Intersection Process

Here we present the full outline of our process for calculating true anomalies where the planet’s orbit has a specified Δmag, which we call an intersection. These steps are discussed in Sec. 2.2.

  • 1. Find Δmag extrema.

  • 2. Find planets where Δmag<Δmagmax, Δmag>Δmagmin, and Δmag>Δmaglmax or Δmag<Δmaglmin (these planets have only two intersections).

  • 3. Substitute into Δmag equation components and expand until all terms are functions of KOE.

  • 4. Manipulate the equation until all ν terms are isolated on the right side.

  • 5. Expand.

  • 6. Replace sin(ν) with 1cos(ν)2.

  • 7. Replace all cos(ν) with x.

  • 8. Isolate all square root terms to general form A21x2=B2.

  • 9. Square both sides.

  • 10. Expand.

  • 11. Find polynomial coefficients of x.

  • 12. Solve with root solver.

  • 13. Filter out solutions with imaginary components (>107).

  • 14. Filter out solutions 1<xk or 1>xk.

  • 15. Take arrays ν0=cos1(x) and ν1=2πν0.

  • 16. Solve for associated arrays Δmag0 and Δmag1.

  • 17. Remove solutions where |Δmag0Δmag|>0.01 and |Δmag1Δmag|>0.01 (Note that 0.01 is <±0.08% error on dmag).

  • 18. Remove duplicate solutions.

  • 19. Verify that there are only two viable solutions.

  • 20. Assign solutions to ν.

10.

Appendix E: Projected Ellipse Semimajor Axis and Semiminor Axis, and θ

We write the analytical expressions for ap, bp, and θ using full expansions of Eqs. (13)–(15). These expressions are far too long to be practically conveyed, so we use the following intermediary parameters here and only here to simplify these expressions:

Eq. (61)

K0=e(1e2)K1=sin(Ω)cos(ω)+sin(ω)cos(Ω)cos(i)K2=e+11eK3=a(1e2)K4=sin(Ω)sin(ω)cos(i)+cos(Ω)cos(ω)K5=K12K32+K42K32+K32sin2(i)sin2(ω)K6=K3(sin(Ω)cos(ω)sin(ω)cos(Ω)cos(i))/(1e)K7=K3(sin(Ω)sin(ω)cos(i)cos(Ω)cos(ω))/(1e)K8=K3(sin(Ω)cos(ω+2tan1(K2))+sin(ω+2tan1(K2))cos(Ω)cos(i))/(ecos(2tan1(K2))+1)K9=K3(sin(Ω)sin(ω+2tan1(K2))cos(i)+cos(Ω)cos(ω+2tan1(K2)))/(ecos(2tan1(K2))+1).

The simplified expression for the semimajor axis of the projected ellipse is

Eq. (62)

ap=12(K0a2K1K5K0a2K4K5+K6K9)2+(K0a2K1K5+K0a2K4K5+K7+K8)2+12(K0a2K1K5K0a2K4K5K7+K8)2+(K0a2K1K5+K0a2K4K5+K6+K9)2.
The simplified expression for the projected ellipse semiminor axis is

Eq. (63)

bp=12|K0a2K1K5K0a2K4K5+K6K9|2+|K0a2K1K5+K0a2K4K5+K7+K8|2+12|K0a2K1K5K0a2K4K5K7+K8|2+|K0a2K1K5+K0a2K4K5+K6+K9|2.
The simplified expression for the angle between the projected ellipse semimajor axis and X axis is

Eq. (64)

θ=12[tan1(K0a2K1K5K0a2K4K5+K6K9K0a2K1K5+K0a2K4K5+K7+K8)+tan1(K0a2K1K5+K0a2K4K5+K6+K9K0a2K1K5+K0a2K4K5+K7K8)].

11.

Appendix F: s Extrema Polynomial

We reduced Eq. (22) from Sec. 2.1.2 into the standard reduced form of a quartic expression:

Eq. (65)

0=xe4+2ap2x*+2bp2x*(ap42ap2bp2+bp4)/ap2xe3+ap4+2ap2bp2+ap2x*2bp4+bp2y*2(ap42ap2bp2+bp4)/ap2xe2+2ap4x*2ap2bp2x*(ap42ap2bp2+bp4)/ap2xe+ap4x*2(ap42ap2bp2+bp4)/ap2.
The coefficients of the quartic expression are

Eq. (66)

A0=2ap2x*+2bp2x*(ap42ap2bp2+bp4)/ap2,

Eq. (67)

B0=ap4+2ap2bp2+ap2x*2bp4+bp2y*2(ap42ap2bp2+bp4)/ap2,

Eq. (68)

C0=2ap4x*2ap2bp2x*(ap42ap2bp2+bp4)/ap2,

Eq. (69)

D0=ap4x*2(ap42ap2bp2+bp4)/ap2.

12.

Appendix G: s Intersection Polynomial

After dividing by the coefficient of xe4 in Eq. (35) from Sec. 2.1.3, we get

Eq. (70)

0=xe44ap2x*ap2bp2xe3+2ap2(ap2bp2ap2s2+3ap2x*2+ap2y*2bp4+bp2s2bp2x*2+bp2y*2)ap42ap2bp2+bp4xe2+4ap4x*(bp2+s2x*2y*2)ap42ap2bp2+bp4xe+ap4(bp42bp2s2+2bp2x*22bp2y*2+s42  s2x*22  s2y*2+x*4+2x*2y*2+y*4)ap42ap2bp2+bp4.
The coefficients of the general quartic expression in Eq. (36) are

Eq. (71)

A1=4ap2x*ap2bp2,

Eq. (72)

B1=2ap2(ap2bp2ap2s2+3ap2x*2+ap2y*2bp4+bp2s2bp2x*2+bp2y*2)ap42ap2bp2+bp4,

Eq. (73)

C1=4ap4x*(bp2+s2x*2y*2)ap42ap2bp2+bp4,
and

Eq. (74)

D1=ap4(bp42bp2s2+2bp2x*22bp2y*2+s42  s2x*22  s2y*2+x*4+2x*2y*2+y*4)ap42ap2bp2+bp4.

13.

Appendix H: Δmag Extrema Polynomial

Here we include the coefficients of Eq. (49) from Sec. 2.2 as

Eq. (75)

A2=e4sin4(i),B2=3e3(esin(ω)+sin(i))sin3(i),C2=14e2(4e2sin2(i)sin2(ω)8e2sin2(i)+8e2sin2(ω)+5e2+34esin(i)sin(ω)+13sin2(i))sin2(i),D2=12e[2e3sin2(i)sin3(ω)8e3sin2(i)sin(ω)+3e3sin(ω)+5e2sin3(i)sin2(ω)11e2sin3(i)+10e2sin(i)sin2(ω)+7e2sin(i)+17esin2(i)sin(ω)+3sin3(i))sin(i),E2=5e4sin4(i)sin2(ω)4+5e4sin4(i)47e4sin2(i)sin2(ω)43e4sin2(i)2+e44+3e3sin3(i)sin3(ω)210e3sin3(i)sin(ω)+7e3sin(i)sin(ω)2+2e2sin4(i)sin2(ω)21e2sin4(i)4+4e2sin2(i)sin2(ω)+7e2sin2(i)2+7esin3(i)sin(ω)2+sin4(i)4,F2=3e4sin3(i)sin3(ω)2+3e4sin3(i)sin(ω)23e4sin(i)sin(ω)2e3sin4(i)sin4(ω)25e3sin4(i)sin2(ω)2+3e3sin4(i)5e3sin2(i)sin2(ω)7e3sin2(i)2+e32e2sin3(i)sin3(ω)28e2sin3(i)sin(ω)+5e2sin(i)sin(ω)2+esin4(i)sin2(ω)22esin4(i)+esin2(i)sin2(ω)+3esin2(i)2+sin3(i)sin(ω)2,G2=e4sin4(i)sin4(ω)4+e4sin4(i)sin2(ω)2e4sin4(i)4e4sin2(i)sin2(ω)2+e4sin2(i)2e447e3sin3(i)sin3(ω)2+7e3sin3(i)sin(ω)27e3sin(i)sin(ω)25e2sin4(i)sin4(ω)4e2sin4(i)sin2(ω)+9e2sin4(i)45e2sin2(i)sin2(ω)5e2sin2(i)2+e243esin3(i)sin3(ω)22esin3(i)sin(ω)+esin(i)sin(ω)2sin4(i)4+sin2(i)4,H2=e3sin4(i)cos4(ω)2+e3sin2(i)cos2(ω)e32+5e2sin3(i)sin(ω)cos2(ω)25e2sin(i)sin(ω)23e(cos2(i)1)2(cos(4w)1)64+esin4(i)cos4(ω)25esin2(i)sin2(ω)2esin2(i)cos2(ω)2sin3(i)sin3(ω)2,andI2=e2sin4(i)cos4(ω)4+e2sin2(i)cos2(ω)2e24+esin3(i)sin(ω)cos2(ω)2esin(i)sin(ω)2(cos(2i)1)2(cos(4w)1)128sin2(i)sin2(ω)4.

14.

Appendix I: Δmag Intersection Polynomial

Here we include the coefficients of Eq. (50) from Sec. 2.2 as

Eq. (76)

A3=e4sin4(i)16,B3=14e3(esin(ω)+sin(i))sin3(i),C3=18e2(e2cos2(i)cos2(ω)3e2cos2(ω)+3e2+8esin(i)sin(ω)3cos2(i)+3)sin2(i),D3=14e[e3sin3(ω)cos2(i)+e3sin3(ω)+e3sin(ω)cos2(i)2e2sin3(i)cos2(ω)4e2sin(i)cos2(ω)+6e2sin(i)6esin(ω)cos2(i)+6esin(ω)+sin3(i))sin(i),E3=e4sin4(i)sin4(ω)16e4sin4(i)sin2(ω)8+e4sin4(i)16+e4sin2(i)sin2(ω)8e4sin2(i)8+e416+e3sin3(i)sin3(ω)e3sin3(i)sin(ω)+e3sin(i)sin(ω)e2ξsin2(i)sin2(ω)+e2ξsin2(i)2+3e2sin4(i)sin2(ω)43e2sin4(i)4+3e2sin2(i)sin2(ω)2+3e2sin2(i)4+esin3(i)sin(ω)+sin4(i)16,F3=e3sin4(i)sin4(ω)4e3sin4(i)sin2(ω)2+e3sin4(i)4+e3sin2(i)sin2(ω)2e3sin2(i)2+e34e2ξsin(i)sin(ω)+3e2sin3(i)sin3(ω)23e2sin3(i)sin(ω)2+3e2sin(i)sin(ω)22eξsin2(i)sin2(ω)+eξsin2(i)+esin4(i)sin2(ω)2esin4(i)2+esin2(i)sin2(ω)+esin2(i)2+sin3(i)sin(ω)4,G3=e2ξsin2(i)cos2(ω)2e2ξ2+3e2sin4(i)cos4(ω)83e2sin2(i)cos2(ω)4+3e282eξsin(i)sin(ω)e  sin3(i)sin(ω)cos2(ω)+e  sin(i)sin(ω)ξ  sin2(i)sin2(ω)2+ξ  sin2(i)cos2(ω)2+(cos(2i)1)2(cos(4w)1)256sin4(i)cos4(ω)8+3  sin2(i)sin2(ω)8+sin2(i)cos2(ω)8,H3=eξsin2(i)cos2(ω)eξ+esin4(i)cos4(ω)4esin2(i)cos2(ω)2+e4ξsin(i)sin(ω)sin3(i)sin(ω)cos2(ω)4+sin(i)sin(ω)4,andI3=ξ2ξsin2(i)cos2(ω)2ξ2+sin4(i)cos4(ω)16sin2(i)cos2(ω)8+116.

15.

Appendix J: Proof that Ellipses Project to Ellipses

Let us suppose that we have some ellipse on an arbitrary plane indicated by the black ellipse in Fig. 4. We will say AB and CD are the principal axes of some this 3D ellipse. AB and CD intersect at point O, the geometric center of this 3D ellipse. We will say point P is any point on the ellipse with H being the projection of point P onto axis AB and K being the projection of point P onto axis CD. By the definition of an ellipse, we have

Eq. (77)

OH¯2OB¯2+OK¯2OD¯2=1.

Now we let A, B, C, D, O, K, H, and P be the perpendicular projections of points A, B, C, D, O, K, H, and P onto any given plane. Since perpendicular projections preserve the ratios of segments on a line, we can now say

Eq. (78)

OH¯2OB¯2+OK¯2OD¯2=1.
This equation means that point P belongs to the ellipse having AB and CD as conjugate diameters. For an ellipse, two diameters are conjugate if and only if the tangent line to the ellipse at an endpoint of one diameter is parallel to the other diameter.

16.

Appendix K: Quartic Solution

The quartic function has a known analytical solution.19 We start with a set of useful simplifying terms

Eq. (79)

p0=(3A28+B)3,

Eq. (80)

p1=(A(A28B2)+C)2,

Eq. (81)

p2=A(A(3A2256B16)+C4)+D,

Eq. (82)

p3=3A28+B,

Eq. (83)

p4=2A(A28B2),

Eq. (84)

p5=p0108p18+p2p33,

Eq. (85)

p6=p0216+p116p2p36+p524+(p2p3212)3273,

Eq. (86)

p7=A242B3,

Eq. (87)

p8=2p2+p3263p6,

Eq. (88)

p9=2p53+p7,

Eq. (89)

p10=2p6+p7+p8,
and

Eq. (90)

p11=A224B3.

The solutions are four piecewise functions:

Eq. (91)

x0={A4+p92p11+2p53+2Cp4p92for  p2+p3212=0A4p102p112p6+p9+2C+p4p102otherwise,

Eq. (92)

x1={A4+p92+p11+2p532C+p4p92for  p2+p3212=0A4p102+p112p6+p9+2C+p4p102otherwise,

Eq. (93)

x2={A4p92p11+2p532Cp4p92for  p2+p3212=0A4+p102p112p6+p9+2Cp4p102otherwise,
and

Eq. (94)

x3={A4p92+p11+2p532Cp4p92for  p2+p3212=0A4+p102+p112p6+p9+2Cp4p102otherwise.

Theoretically, we always have four solutions. Theoretically, we can use equations of coefficients to determine how many roots are real, complex, and double roots. The general form of the quartic defines the expressions Δ, P, D2, R, and Δ0, which are

Eq. (95)

Δ=256D3192ACD2128B2D2+144BC2D27C4+144A2BD26A2C2D80AB2CD+18ABC3+16B4D4B3C227A4D2+18A3BCD4A3C34A2B3D+A2B2C2,

Eq. (96)

P=8B3A2,

Eq. (97)

D2=64D16B2+16A2B16AC3A4,

Eq. (98)

R=A3+8C4AB,

Eq. (99)

Δ0=B23AC+12D.

After evaluation of these constants, we can determine

Eq. (100)

solution types={2distinctRroots&2complex conjugate rootsΔ<04Rdistinct rootsΔ>0&P<0&D<02pair complex conjugate rootsΔ>0&(P>0|D>0).
However, the use of this theoretical classification does not work when numerical errors are introduced. The accumulation of numerical errors causes the solutions to be improperly classified for a sufficiently large number of cases to prohibit the use of these classifications. The Δ expression rarely evaluates to Δ=0 when it should do so quite frequently. Numerical rounding errors can also frequently result in the evaluation Δ<0 when in reality Δ>0 and vice versa.

17.

Appendix L: Earth-Like Subpopulation

We use a definition of Earth-like exoplanet similar to that used in Ref. 8. We use a more conservative planetary radius range of 0.9RR1.4R. We use the flux-at-planet range of 0.3586Lplan1.1080. To classify a planet, we use the time-averaged incident flux on the planet calculated by

Eq. (101)

Lplan=L*(a+ae22)2.

18.

Appendix M: 4-m Monolith Design Reference Mission

We calculated the Brown completeness and integration time adjusted completeness for a 4-m monolith telescope with a contrast of 1010 on the SAG13 planet population discussed in Ref. 1. The Δmag and separations of integration are scaled by the luminosity of each candidate target star. The orbital periods are similarly scaled by the mass of each target star in the case of integration time adjusted completeness. We optimized the mission schedule using the sequential least squares quadratic programming method discussed in Ref. 1 with the Brown completeness method. Our approach included the adoption of the use of the local zodiacal light minimum and filtering of target stars.

We assume a total telescope time of 213.67 days (7 months) and that each observation requires 0.1 day of overhead time and 0.042 day settling time. The coronagraph parameters used are from a vortex charge six coronagraph at a wavelength of 500 nm with an IWA of 0.045 arc sec and OWA of 2.127 arc sec and a working angle of integration consistent with that used in Ref. 1.

The resulting Brown completeness optimized integration times for a DRM are included in Table 2. The summed Brown completeness for this DRM is 68.89, whereas the same summed integration time adjusted completeness is 63.11, a difference of 5.78. We can multiply either of these numbers by the SAG13 exoplanet occurrence rate of 5.62 to get the expected number of exoplanets detected in a single-visit blind search. These are 387.16 and 354.67 on average, respectively. This means that 32.49 planets are lost in a mission by not taking into account integration times when optimizing the mission schedule.

Table 2

4-m monolith DRM.

NameDist (pc)Int. time (day)ΔmagCbrownCIAC
HIP 4394.340.39919.5570.0490.041
HIP 52225.710.18123.4760.0380.033
HIP 54413.670.623.4030.160.141
HIP 74616.780.06627.1420.2470.24
HIP 91018.750.38324.9160.160.146
HIP 95021.280.34124.5040.1060.095
HIP 129217.50.33822.1810.0540.047
HIP 1392 B15.190.29321.3620.0390.033
HIP 14753.570.62520.2370.1050.089
HIP 149923.250.13921.7360.0180.016
HIP 15998.590.37425.7930.470.438
HIP 180320.860.24122.2090.0360.031
HIP 20217.460.17926.9420.550.525
HIP 207223.810.12925.7730.1030.097
HIP 208125.970.05326.4320.0820.078
HIP 271125.480.19323.7520.0450.039
HIP 309311.060.74523.7310.2420.213
HIP 317024.960.19223.2570.0380.033
HIP 341929.530.03526.5450.0480.046
HIP 349722.060.18421.9280.0250.022
HIP 350526.750.16924.1710.0440.039
HIP 358315.160.56723.9190.170.151
HIP 37657.450.86123.9740.380.339
HIP 381023.450.2724.4460.0780.07
HIP 390915.750.51524.6850.2070.187
HIP 414814.170.29321.1920.0390.033
HIP 415118.740.35325.070.1660.152
HIP 53367.550.6624.8590.4590.418
HIP 586215.110.4625.0010.2380.217
HIP 637916.810.21721.1880.0280.024
HIP 670625.210.20723.990.0510.046
HIP 681328.620.1324.5350.0370.033
HIP 723519.050.16421.170.0210.018
HIP 733920.660.23622.1060.0350.03
HIP 751313.490.29625.7650.3090.288
HIP 760127.380.12923.040.0230.02
HIP 773421.370.1721.6430.0220.019
HIP 791812.740.52124.9780.2950.269
HIP 797817.430.48324.3020.1530.137
HIP 79817.530.73324.6110.440.398
HIP 81023.650.32326.3640.7450.704
HIP 836210.070.75324.2540.3180.285
HIP 843327.860.11923.0240.0210.019
HIP 849723.190.24125.0780.0980.09
HIP 879619.420.14126.1160.180.17
HIP 890317.990.0826.8140.2170.209
HIP 900717.850.19925.890.2070.194
HIP 923622.010.06926.5650.1380.132
HIP 988420.180.06926.7530.1720.165
HIP 1013810.780.74623.5440.2310.204
HIP 1030628.870.11523.7790.0260.023
HIP 1064410.780.52825.0920.3590.329
HIP 1072324.350.21823.5130.0470.041
HIP 1079812.670.60522.9750.1460.128
HIP 1102929.660.09723.5510.020.018
HIP 1178326.70.17924.7680.0550.049
HIP 121147.180.8823.8660.3790.337
HIP 1218625.780.17323.3070.0340.03
HIP 1244421.760.29923.5260.0670.059
HIP 1262324.190.24924.6610.0760.068
HIP 1265317.170.48824.4730.1670.15
HIP 1277711.130.32425.8290.3810.356
HIP 1284314.220.35225.4710.280.259
HIP 1340210.350.77123.5720.2450.216
HIP 1366526.710.15923.6160.0350.031
HIP 1414627.170.11225.5390.0620.058
HIP 1415020.640.16221.4350.0210.018
HIP 1463210.540.32825.8550.4020.375
HIP 1495422.580.29924.5110.090.08
HIP 1533012.010.66924.4370.2780.25
HIP 1537112.030.59224.7790.3030.275
HIP 154579.140.56725.0730.4130.378
HIP 155106.040.42625.7510.5810.541
HIP 1624521.680.26425.160.1210.111
HIP 165373.210.37726.1210.7680.722
HIP 1685213.960.33925.5420.2890.268
HIP 173789.040.28626.1620.4660.438
HIP 1742013.950.32621.3240.0440.038
HIP 1743916.030.28621.4750.0390.033
HIP 1765117.630.2525.640.2060.192
HIP 1885918.830.42924.3110.1320.118
HIP 1907616.940.46623.4330.1140.101
HIP 1920527.60.14724.0210.0360.032
HIP 1933521.00.34123.9730.090.08
HIP 198494.980.50425.5160.6250.579
HIP 1985921.330.24422.410.0390.034
HIP 1988413.040.1920.3970.0240.02
HIP 1989320.460.20625.6080.1520.141
HIP 1992118.240.32225.2610.1830.168
HIP 1999028.940.12224.3830.0330.03
HIP 2142120.430.03927.3510.170.166
HIP 2154729.430.10723.9290.0250.022
HIP 2177020.170.2425.4490.1530.141
HIP 2186128.670.12724.3180.0340.03
HIP 2226313.280.63524.4020.2430.218
HIP 2233626.420.15423.1830.0290.026
HIP 224498.070.21826.6170.5160.491
HIP 233118.710.83223.260.2580.226
HIP 2348226.080.18524.0740.0470.042
HIP 2369311.650.43425.3510.3460.319
HIP 2378326.290.18924.5870.0550.049
HIP 2387527.40.04426.480.0670.064
HIP 2394125.460.20824.3640.0570.051
HIP 241863.910.37719.2980.0440.037
HIP 2478625.030.18323.0450.0340.029
HIP 2481312.630.45825.2070.3110.285
HIP 2511020.890.34324.7490.120.108
HIP 2527814.390.51724.8470.2460.224
HIP 2554419.20.18321.3470.0240.02
HIP 2562313.020.16320.1480.020.017
HIP 258785.680.48320.340.0680.058
HIP 2639418.320.44624.0740.1280.114
HIP 2677912.280.64523.2150.1710.15
HIP 270728.930.25626.3260.4740.448
HIP 2728821.610.10826.1650.1410.134
HIP 2732119.440.15526.0120.1780.167
HIP 2743515.180.54423.5390.1450.128
HIP 2762826.730.08125.9360.0710.066
HIP 2788713.00.37621.4130.0530.045
HIP 2789026.250.1924.7540.0580.053
HIP 2810314.880.226.1430.2820.266
HIP 2836024.870.03127.2170.0980.096
HIP 2890825.620.14322.4890.0220.019
HIP 2895415.270.32721.5750.0460.039
HIP 2927110.20.58224.950.3690.337
HIP 292955.750.40120.0080.0510.043
HIP 2952517.950.33722.30.0550.048
HIP 2956816.720.42322.7520.0830.073
HIP 2965020.820.35824.4250.1090.097
HIP 2980019.250.39424.7510.1450.131
HIP 2986019.250.39623.7680.1020.09
HIP 3050321.880.22822.4010.0360.031
HIP 3159219.750.22525.5680.1630.151
HIP 323492.630.02530.1770.8720.861
HIP 3236218.00.14526.2210.210.199
HIP 3236624.380.20723.2380.0410.036
HIP 3243917.870.46524.4140.1520.136
HIP 3248016.720.49824.5650.180.163
HIP 3260729.60.05126.0980.0460.043
HIP 3276525.260.21524.4540.0610.055
HIP 329848.710.77722.7530.2050.179
HIP 3309425.890.15422.8240.0260.023
HIP 3327717.240.47123.7210.1250.111
HIP 3369018.330.24521.6380.0330.028
HIP 3381714.650.42522.0860.0710.061
HIP 3401719.130.37223.2590.0820.072
HIP 3483421.430.23125.3680.130.12
HIP 3513616.890.50324.1780.1560.139
HIP 35296 A14.590.44422.2140.0770.067
HIP 3636618.050.24525.6260.1970.183
HIP 3639927.240.13223.0320.0240.021
HIP 3643920.240.3824.3540.1140.102
HIP 3679525.30.18425.1720.0760.069
HIP 372793.510.06328.8160.7960.778
HIP 3734914.210.28121.130.0370.031
HIP 3760624.670.23324.4760.0670.06
HIP 3782610.360.07327.7560.4420.43
HIP 3878417.190.35822.2620.0590.051
HIP 3890816.20.53224.2790.1740.155
HIP 3915716.770.21621.1730.0280.024
HIP 3934217.310.19521.1210.0250.022
HIP 3975719.480.08326.6210.1840.177
HIP 3978023.290.26624.0170.0680.06
HIP 3990319.980.30525.1550.1480.135
HIP 4003522.380.29223.9550.0750.066
HIP 4069312.490.67423.6890.2050.181
HIP 4070219.560.19625.750.1710.159
HIP 4070628.630.12425.1770.0450.041
HIP 4084318.270.43824.6340.1560.141
HIP 4121126.620.15923.4930.0330.029
HIP 4148422.250.21622.3680.0330.029
HIP 4192612.210.62122.9330.1510.132
HIP 4243814.360.60724.1950.2040.182
HIP 4280811.140.63122.5740.1410.123
HIP 4358712.340.67323.5520.1970.173
HIP 4372617.390.44223.350.1050.092
HIP 4379724.150.22923.680.0520.046
HIP 4407521.030.32323.6090.0770.068
HIP 4412714.510.12926.6770.30.288
HIP 4414326.420.18224.3460.0490.044
HIP 4489719.190.37723.3720.0860.076
HIP 4517020.360.23221.9790.0330.028
HIP 4533319.570.39724.6020.1340.12
HIP 453436.110.56620.8080.090.076
HIP 4561717.270.13120.5370.0160.014
HIP 4650917.330.34225.2610.2010.184
HIP 4658012.910.28120.9190.0360.031
HIP 4673323.820.11225.9130.1050.099
HIP 4684317.790.16120.930.0210.018
HIP 4685313.480.15826.5430.3270.312
HIP 4708011.370.68924.4240.2950.265
HIP 4759215.010.45725.0250.2410.22
HIP 4811318.370.42524.7270.1590.143
HIP 4833111.160.23720.3990.030.025
HIP 4883328.150.13624.1380.0340.031
HIP 4908115.050.57124.4310.2060.185
HIP 4959328.240.13725.0740.0470.043
HIP 4966924.310.02927.690.1060.105
HIP 4980927.740.14123.9240.0340.03
HIP 499084.870.98922.8920.3560.312
HIP 5007522.850.19822.3310.030.026
HIP 5038422.810.25723.3480.0540.047
HIP 5050520.240.19821.6510.0260.022
HIP 5056421.370.30924.9210.1190.108
HIP 5095416.220.21925.9220.2440.229
HIP 5145912.780.46425.1770.3050.28
HIP 5150221.50.33324.570.1050.094
HIP 5152321.810.29824.9140.1120.102
HIP 5181426.530.17924.3620.0490.044
HIP 5193325.080.19523.4130.040.035
HIP 5322929.090.10825.2970.0430.04
HIP 5325329.130.10125.3910.0440.041
HIP 5372114.060.50924.9050.2570.234
HIP 5391024.450.03926.9510.1030.1
HIP 540352.541.021.6040.3080.263
HIP 5418228.990.12524.6960.0370.033
HIP 542114.860.26319.0930.0270.023
HIP 5474521.930.21322.220.0320.027
HIP 5487217.910.07426.9030.2190.212
HIP 55691 A13.710.17420.3770.0220.018
HIP 5577927.220.15824.1890.0410.037
HIP 5644527.230.13323.0750.0240.021
HIP 564529.560.79924.0010.310.276
HIP 5680226.730.15823.5960.0340.03
HIP 569979.610.68824.6180.3650.329
HIP 5699812.40.17820.1790.0220.018
HIP 574439.220.5525.1150.4130.378
HIP 5750717.470.35822.3570.0610.053
HIP 5763211.00.08827.4430.4170.404
HIP 5775710.930.25926.1480.3980.374
HIP 579399.090.78222.9920.220.192
HIP 5800125.50.03726.8840.0890.087
HIP 5834510.160.54921.7070.0930.08
HIP 5857612.760.67624.1440.2360.21
HIP 5880325.320.21224.3710.0580.052
HIP 589105.520.7821.5630.1670.143
HIP 5907219.760.20425.6890.1660.154
HIP 5919914.940.25325.8550.2730.256
HIP 5975022.350.24422.8160.0430.038
HIP 5977424.690.07226.260.0970.092
HIP 6096526.630.05126.4070.0750.071
HIP 6105321.780.26122.8340.0470.041
HIP 6108427.150.03726.7650.070.068
HIP 6117418.280.27125.4750.1880.174
HIP 6129116.180.24821.2740.0330.028
HIP 613178.440.38625.7550.4750.442
HIP 6214514.880.31821.4420.0430.037
HIP 6220717.380.45623.5670.1160.102
HIP 6252316.930.41822.7890.0830.073
HIP 629517.530.37820.3570.0490.041
HIP 6295625.310.0327.460.0930.091
HIP 6307629.290.11224.0850.0280.025
HIP 6361327.870.10725.5020.0550.051
HIP 637214.620.53820.2550.0810.068
HIP 643949.130.38125.7320.4470.416
HIP 6440820.670.33724.8570.1270.115
HIP 6458318.20.3825.0150.1740.159
HIP 6479217.560.46724.5860.1660.15
HIP 649248.560.55325.1470.4390.403
HIP 6510918.020.07626.8590.2160.209
HIP 6535215.450.2521.1640.0330.028
HIP 6535516.140.16220.6610.0210.017
HIP 6553021.170.22822.1630.0340.029
HIP 6572117.990.42624.8020.1690.153
HIP 6624922.710.09526.180.1240.117
HIP 6676515.650.29321.4390.0390.034
HIP 6715319.40.23625.540.1690.157
HIP 671555.410.31919.5350.0360.03
HIP 6727515.620.34525.3860.2420.224
HIP 67422 A13.410.36221.4130.050.043
HIP 6792711.40.13926.8840.3970.38
HIP 6803024.860.16522.5870.0260.023
HIP 6818410.060.72922.9590.1940.169
HIP 6868216.980.4222.8330.0850.074
HIP 6893318.030.07226.9120.2160.209
HIP 6967121.220.25822.5410.0430.037
HIP 6970122.240.19425.4910.1210.112
HIP 6971329.070.12424.7730.0370.034
HIP 6996518.030.4323.5290.1060.093
HIP 6997211.80.59222.4820.1240.108
HIP 6998926.10.18123.9030.0430.038
HIP 7031917.190.41122.8150.0820.072
HIP 7049714.530.26125.850.2840.266
HIP 7085719.890.17821.4260.0230.02
HIP 7087323.740.17222.2670.0250.022
HIP 7107526.610.05626.3110.0740.071
HIP 7118113.220.33721.2540.0460.039
HIP 7128415.830.31425.4950.2410.224
HIP 7185520.00.20321.6490.0270.023
HIP 7195718.270.20425.8230.1970.185
HIP 7256718.170.42923.6440.1090.097
HIP 7260322.980.28524.3980.0820.073
HIP 7262223.240.06226.5370.1190.114
HIP 7284811.510.71523.6020.2190.193
HIP 7310025.110.20223.6790.0460.04
HIP 7316526.90.16525.0430.0580.052
HIP 7399619.550.34725.0010.150.136
HIP 7427324.20.16722.3660.0250.022
HIP 7453717.670.31121.9820.0460.04
HIP 7460525.340.21124.3860.0590.053
HIP 7470215.850.27421.3670.0360.031
HIP 7497525.380.21124.3520.0580.052
HIP 7518114.810.58924.0920.1880.167
HIP 7571820.580.13921.2140.0180.015
HIP 7580921.850.17621.8030.0240.02
HIP 7621928.930.12324.3810.0330.03
HIP 7626723.010.03827.0870.1250.121
HIP 7682917.440.3425.260.1980.182
HIP 7705214.660.57823.7220.1650.146
HIP 7707022.680.08426.3060.1250.119
HIP 7725712.120.38525.50.3380.313
HIP 7735815.350.52523.3720.1320.117
HIP 7776015.890.35225.3360.2340.216
HIP 7780117.350.43723.2780.1020.089
HIP 7795212.380.12626.9080.3650.351
HIP 7807211.250.27226.0580.3850.361
HIP 7845917.240.48824.4290.1640.147
HIP 7852721.030.17325.7480.1450.135
HIP 7877514.520.45122.2440.0790.069
HIP 7919014.670.30421.3310.0410.035
HIP 7924817.570.32822.1130.0510.044
HIP 7953713.890.22620.7870.0290.025
HIP 7957821.550.16821.6620.0220.019
HIP 7967213.90.62524.2420.2170.194
HIP 7982229.740.10724.3460.0280.026
HIP 8017927.270.16524.6980.0490.044
HIP 8033128.230.05326.2020.0580.055
HIP 8033712.780.64124.4720.2610.235
HIP 8068612.120.49225.1160.320.293
HIP 813009.750.79423.9770.3020.268
HIP 8193514.260.16620.3860.0210.017
HIP 820039.810.17919.7320.0210.017
HIP 8202026.110.19224.8160.0610.055
HIP 8239619.540.08526.6010.1830.175
HIP 8258729.190.11123.8850.0260.023
HIP 8258817.250.30321.8270.0430.037
HIP 8262126.940.1322.7850.0210.019
HIP 8286015.260.41925.1380.2410.22
HIP 8300028.040.08425.7430.0560.052
HIP 8338918.590.2521.730.0340.029
HIP 8354117.840.28421.8150.040.034
HIP 8359110.710.21820.1970.0270.023
HIP 8360120.670.31723.2140.0670.058
HIP 836095.620.31119.5610.0350.029
HIP 8399013.620.2620.9290.0340.029
HIP 8414322.530.10426.1170.1260.119
HIP 844785.950.9523.0710.3320.29
HIP 8486214.330.57824.570.2310.208
HIP 8504219.520.31622.6280.0560.049
HIP 8523512.80.58822.8530.1360.119
HIP 852957.70.4520.7040.0630.053
HIP 8620123.160.25125.0070.0970.088
HIP 8640011.00.64422.6240.1470.128
HIP 8648621.450.2925.0370.1210.11
HIP 8661422.840.22625.2350.1070.098
HIP 8662023.060.25723.5380.0570.05
HIP 8674225.090.07926.1320.0910.086
HIP 8679615.510.51224.7440.2150.195
HIP 869748.310.25326.3930.5010.474
HIP 8817523.550.23325.0730.0940.085
HIP 8863529.70.06525.8330.0440.041
HIP 8869417.550.43923.3820.1040.092
HIP 8877126.630.08825.850.0710.066
HIP 8897211.020.68623.010.1780.156
HIP 8904217.610.47624.30.150.135
HIP 8934822.920.2824.790.0940.085
HIP 8947422.820.21322.540.0340.03
HIP 8996218.540.15426.1070.1970.186
HIP 9049623.970.08826.1290.1050.099
HIP 9079013.250.39921.5870.0590.05
HIP 912627.680.0329.1350.5590.549
HIP 917683.520.4119.2890.050.041
HIP 9202428.550.13624.8040.0410.037
HIP 9204319.210.22925.6030.1740.162
HIP 9216128.890.11525.2590.0440.041
HIP 9254925.670.18723.6970.0420.037
HIP 9374725.460.05526.4490.0880.084
HIP 9385816.950.43723.0180.0940.082
HIP 9396621.430.28523.0490.0560.049
HIP 9408327.30.15924.4030.0440.039
HIP 9437629.870.0625.8890.0430.04
HIP 9514918.830.29322.1470.0450.039
HIP 9531915.760.46522.8220.0970.085
HIP 9544715.180.53724.6620.2170.196
HIP 9550115.530.15826.3570.2690.256
HIP 95995 A16.960.36722.2780.0610.053
HIP 961005.750.53225.3740.5760.531
HIP 9625825.110.19823.5680.0430.038
HIP 9644118.340.28125.4240.1860.172
HIP 9689521.080.31523.4310.070.062
HIP 9690121.210.28422.9470.0540.047
HIP 976495.120.06128.6690.6890.673
HIP 9765027.870.13523.7120.030.027
HIP 9767519.190.40324.6790.1430.129
HIP 9794414.050.55223.0110.1280.113
HIP 9803613.70.25725.9420.3080.289
HIP 9806625.990.19624.5610.0570.051
HIP 9847021.250.32223.7210.0780.069
HIP 9869812.860.22520.6060.0290.024
HIP 9876715.860.53923.9470.1590.142
HIP 9879215.770.22221.0510.0290.025
HIP 9881917.770.45323.7910.1220.108
HIP 9892118.790.37123.0360.0760.067
HIP 9895917.730.42523.2980.0980.086
HIP 9903123.950.24223.9660.0610.054
HIP 9913723.430.19722.5330.0310.027
HIP 992406.110.29926.3120.5980.564
HIP 994616.010.78224.5370.4990.452
HIP 9957227.70.11822.8920.020.018
HIP 997016.20.40220.1370.0520.044
HIP 998258.910.82523.9230.3230.287
HIP 10001717.570.45323.6530.1170.104
HIP 10051126.250.13122.5090.020.018
HIP 10092519.520.21621.6570.0290.025
HIP 10134524.40.21223.3930.0440.039
HIP 10161227.790.15324.7970.0460.042
HIP 10198324.660.23224.4250.0660.059
HIP 10199714.380.49922.6220.1010.088
HIP 10204020.950.24222.2540.0370.032
HIP 10233324.170.19525.2580.090.083
HIP 10242214.270.20126.1910.2990.282
HIP 10248514.680.29725.6720.2750.256
HIP 10248822.290.06826.5550.1330.128
HIP 10338921.970.29123.5790.0670.059
HIP 10345822.140.17621.8630.0240.02
HIP 10367328.010.1223.1630.0220.02
HIP 1042143.50.77824.7040.6570.599
HIP 1042173.50.96623.8520.5620.503
HIP 104239 A17.570.1320.5840.0160.014
HIP 1050903.951.01922.7920.390.341
HIP 10519915.040.07727.1640.2910.282
HIP 1058589.260.36125.8040.4450.415
HIP 1064404.950.28819.2450.0310.026
HIP 10655927.10.1423.2080.0270.023
HIP 10669614.620.28321.2120.0370.032
HIP 10708921.20.17425.7280.1420.132
HIP 10735017.890.4323.4660.1040.092
HIP 10755611.870.14226.8180.380.364
HIP 10764915.990.53924.2020.1730.154
HIP 10797527.450.14323.6630.0310.028
HIP 10803626.620.17324.2390.0460.041
HIP 1088703.620.62525.1830.6880.633
HIP 10917611.730.25226.120.3710.349
HIP 10937821.560.16421.6280.0220.018
HIP 10942218.280.39924.9050.1680.153
HIP 10942728.30.07125.8940.0550.052
HIP 10982122.050.2422.6240.040.035
HIP 11064920.560.36924.3630.110.098
HIP 11144922.680.29424.3140.0830.074
HIP 11211723.640.2223.070.0420.037
HIP 11244716.30.28625.5860.2330.217
HIP 11293527.280.15524.0590.0380.034
HIP 11313726.750.13122.7210.0210.019
HIP 1132837.610.84922.8210.2420.211
HIP 11335715.610.55424.3060.1860.167
HIP 1133687.70.06528.2250.5540.54
HIP 11342119.860.3122.6880.0550.048
HIP 1135768.220.27320.0180.0330.028
HIP 11386029.420.10924.090.0270.024
HIP 1140463.281.021.9370.3080.265
HIP 11443027.650.12323.0110.0220.019
HIP 11457024.590.19225.2210.0850.077
HIP 1146226.540.81124.3770.460.415
HIP 11492420.50.36224.0420.0980.087
HIP 11494820.540.35623.940.0940.083
HIP 11499623.060.15425.6760.1120.104
HIP 11608516.930.29621.7110.0410.036
HIP 11658426.410.13325.4330.0680.063
HIP 11672714.10.18626.30.3050.29
HIP 11674511.420.4421.4270.0650.056
HIP 11677113.710.31725.660.30.279
HIP 1200056.110.54320.7040.0830.07

Acknowledgments

This work was funded by the Science Investigation Team of the Nancy Grace Roman Space Telescope under NASA (Grant No. NNX15AB40G). This work was an expansion of the work done in Ref. 20.

References

1. 

D. R. Keithly et al., “Optimal scheduling of exoplanet direct imaging single-visit observations of a blind search survey,” J. Astron. Telesc. Instrum. Syst., 6 027001 (2020). https://doi.org/10.1117/1.JATIS.6.2.027001 Google Scholar

2. 

R. A. Brown, “Single-visit photometric and obscurational completeness,” Astrophys. J., 624 1010 –1024 (2005). https://doi.org/10.1086/429124 ASJOAB 0004-637X Google Scholar

3. 

D. Savransky, C. Delacroix and D. Garrett, “EXOSIMS: exoplanet open-source imaging mission simulator,” (2017). Google Scholar

4. 

R. K. Kopparapu et al., “Exoplanet classification and yield estimates for direct imaging missions,” Astrophys. J., 856 122 (2018). https://doi.org/10.3847/1538-4357/aab205 ASJOAB 0004-637X Google Scholar

5. 

D. Garrett, “Exoplanet direct imaging detection metrics and exoplanet popu- lations,” Cornell University, (2018). Google Scholar

6. 

D. Garrett and D. Savransky, “Analytical formulation of the single-visit completeness joint probability density function,” Astrophys. J., 828 20 (2016). https://doi.org/10.3847/0004-637X/828/1/20 ASJOAB 0004-637X Google Scholar

7. 

N. J. Kasdin et al., “The Nancy Grace Roman Space Telescope Coronagraph Instrument (CGI) technology demonstration,” Proc. SPIE, 11443 114431U (2020). https://doi.org/10.1117/12.2562997 Google Scholar

8. 

Science and Technology Definition Team, “HabEx final report,” (2019). Google Scholar

9. 

Science and Technology Definition Team, “LUVOIR final report,” (2019). Google Scholar

10. 

R. Morgan et al., “Exoplanet yield estimation for decadal study concepts using EX,” in 227th Meeting Am. Astron. Soc., (2020). Google Scholar

11. 

R. Morgan et al., “The standard definitions and evaluation team final report a common comparison of exoplanet yield,” (2019). Google Scholar

12. 

Jet Propulsion Laboratory, “Habitable exoplanet observatory final report,” (2019). Google Scholar

13. 

R. A. Brown and R. Soummer, “New completeness methods for estimating exoplanet discoveries by direct detection,” Astrophys. J., 715 122 –131 (2010). https://doi.org/10.1088/0004-637X/715/1/122 ASJOAB 0004-637X Google Scholar

14. 

B. Spain, Analytical Conics, Dover, London (2007). Google Scholar

15. 

W. H. Besant, Conic Sections, Treated Geometrically, 9th ed.London(1895). Google Scholar

16. 

T. A. A. Broadbent, L. Kuipers and R. Timman, Handbook of Mathematics, 5th ed.Springer, Berlin (1970). Google Scholar

17. 

E. Agol, “Rounding up the wanderers: optimizing coronagraphic searches for extrasolar planets,” Mon. Not. R. Astron. Soc., 374 (4), 1271 –1289 (2007). https://doi.org/10.1111/j.1365-2966.2006.11232.x MNRAA4 0035-8711 Google Scholar

18. 

D. Savransky, E. Cady and N. J. Kasdin, “Parameter distributions of Keplerian orbiTS,” Astrophys. J., 728 (1), 66 (2011). https://doi.org/10.1088/0004-637X/728/1/66 ASJOAB 0004-637X Google Scholar

19. 

M. Abramowitz and I. A. Stegun, Handbook of Mathemaical Functions with Formulas, Graphsm and Mathematical Tables, 9th ed.Dover(1972). Google Scholar

20. 

D. Keithly and D. Savransky, “Integration time adjusted completeness,” Proc. SPIE, 11443 1144324 (2020). https://doi.org/10.1117/12.2562636 Google Scholar

Biography

Dean Keithly is a PhD candidate in the Sibley School of Mechanical and Aerospace Engineering at Cornell University. He received his bachelor’s degree in mechanical engineering from Michigan Technological University, where he built the Oculus-ASR satellite thermal control subsystem. He received his master’s degree in systems engineering from Cornell University, working on projects funded by the NASA Institute of Advanced Concepts. His primary research focus is on spacecraft modeling, mission modeling, optimization, and scheduling.

Dmitry Savransky is a professor in the Sibley School of Mechanical and Aerospace Engineering at Cornell University and PI of the Space Imaging and Optical Systems Laboratory. He received his PhD from Princeton University in 2011 and was a postdoctoral researcher at Lawrence Livermore National Laboratory, where he worked on integration, testing and commissioning of the Gemini Planet Imager. His research focuses on the optimization of ground and space-based exoplanet imaging surveys, control of autonomous optical systems, and advanced image processing.

Corey Spohn received a Bachelor of Science in both physics and engineering science and mechanics from Virginia Tech in May 2018. During his time at Virginia Tech he focused heavily on astrophysics while conducting research on anti-frosting surfaces, inter-droplet ice bridging, and flying snakes. For his senior design project, he investigated human-robot dynamics in a crowded panic situation by having humans play the children’s game sharks and minnows with robots in a motion capture studio. He began working towards a PhD in the fall of 2018 and joined the Space Imaging and Optical Systems Lab in December 2018 where he is currently researching exoplanet mission scheduling.

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.
Dean Keithly, Dmitry Savransky, and Corey Spohn "Integration time adjusted completeness," Journal of Astronomical Telescopes, Instruments, and Systems 7(3), 037002 (1 September 2021). https://doi.org/10.1117/1.JATIS.7.3.037002
Received: 30 March 2021; Accepted: 16 August 2021; Published: 1 September 2021
Lens.org Logo
CITATIONS
Cited by 1 scholarly publication.
Advertisement
Advertisement
KEYWORDS
Planets

Stars

Exoplanets

Visibility

Telescopes

Monte Carlo methods

Picosecond phenomena

Back to Top