KEPLER ECLIPSING BINARY STARS. VII. THE CATALOG OF ECLIPSING BINARIES FOUND IN THE ENTIRE KEPLER DATA SET

, , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , and

Published 2016 February 23 © 2016. The American Astronomical Society. All rights reserved.
, , Citation Brian Kirk et al 2016 AJ 151 68 DOI 10.3847/0004-6256/151/3/68

1538-3881/151/3/68

ABSTRACT

The primary Kepler Mission provided nearly continuous monitoring of ∼200,000 objects with unprecedented photometric precision. We present the final catalog of eclipsing binary systems within the 105 deg2 Kepler field of view. This release incorporates the full extent of the data from the primary mission (Q0-Q17 Data Release). As a result, new systems have been added, additional false positives have been removed, ephemerides and principal parameters have been recomputed, classifications have been revised to rely on analytical models, and eclipse timing variations have been computed for each system. We identify several classes of systems including those that exhibit tertiary eclipse events, systems that show clear evidence of additional bodies, heartbeat systems, systems with changing eclipse depths, and systems exhibiting only one eclipse event over the duration of the mission. We have updated the period and galactic latitude distribution diagrams and included a catalog completeness evaluation. The total number of identified eclipsing and ellipsoidal binary systems in the Kepler field of view has increased to 2878, 1.3% of all observed Kepler targets. An online version of this catalog with downloadable content and visualization tools is maintained at http://keplerEBs.villanova.edu.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The contribution of binary stars and, in particular, eclipsing binaries (EBs) to astrophysics cannot be overstated. EBs can provide fundamental mass and radius measurements for the component stars (e.g., see the extensive review by Andersen 1991). These mass and radius measurements in turn allow for accurate tests of stellar evolution models (e.g., Pols et al. 1997; Schroder et al. 1997; Guinan et al. 2000; Torres & Ribas 2002). In cases where high-quality radial velocity (RV) measurements exist for both stars in an EB, the luminosities computed from the absolute radii and effective temperatures can lead to a distance determination. Indeed, EBs are becoming widely used to determine distances to the Magellanic Clouds, M31, and M33 (Guinan et al. 1998; Wyithe & Wilson 2001, 2002; Ribas et al. 2002; Bonanos et al. 2003, 2006; Hilditch et al. 2005; North et al. 2010).

Large samples are useful to determine statistical properties and for finding rare binaries which may hold physical significance (for example, binaries with very low mass stars, binaries with stars in short-lived stages of evolution, very eccentric binaries that show large apsidal motion, etc.). Catalogs of EBs from ground-based surveys suffer from various observational biases such as limited accuracy per individual measurement, complex window functions (e.g., observations from ground based surveys can only be done during nights with clear skies and during certain seasons).

The NASA Kepler Mission, launched in 2009 March, provided essentially uninterrupted, ultra-high precision photometric coverage of ∼200,000 objects within a 105 deg2 field of view in the constellations of Cygnus and Lyra for four consecutive years. The details and characteristics of the Kepler instrument and observing program can be found in Batalha et al. (2010), Borucki et al. (2011), Caldwell et al. (2010), Koch et al. (2010). The mission has revolutionized the exoplanetary and EB field. The previous catalogs can be found in Prša et al. (2011, hereafter Paper I) and Slawson et al. (2011, hereafter Paper II) at http://keplerEBs.villanova.edu/v1 and http://keplerEBs.villanova.edu/v2, respectively. The current catalog and information about its functionality is available at http://keplerEBs.villanova.edu.

2. THE CATALOG

The Kepler Eclipsing Binary Catalog lists the stellar parameters from the Kepler Input Catalog (KIC) augmented by: primary and secondary eclipse depth, eclipse width, separation of eclipse, ephemeris, morphological classification parameter, and principal parameters determined by geometric analysis of the phased light curve.

The online Catalog also provides the raw and detrended data for ∼30 minutes (long) cadence, and raw ∼1 minutes (short) cadence data (when available), an analytic approximation via a polynomial chain (polyfit; Prša et al. 2008), and eclipse timing variations (ETV; Conroy et al. 2014, hereafter Paper IV and J. Orosz et al. 2016, in preparation). The construction of the Catalog consists of the following steps (1) EB signature detection (Section 3); (2) data detrending: all intrinsic variability (such as chromospheric activity, etc.) and extrinsic variability (i.e., third light contamination and instrumental artifacts) are removed by the iterative fitting of the photometric baseline (Prša et al. 2011); (3) the determination of the ephemeris: the time-space data are phase-folded and the dispersion minimized; (4) Determination of ETVs (Section 8.6); (5) analytic approximation: every light curve is fit by a polyfit (Prša et al. 2008); (6) morphological classification via Locally Linear Embedding (LLE; Section 6), a nonlinear dimensionality reduction tool is used to estimate the "detachedness" of the system (Matijevič et al. 2012, hereafter Paper III); (7) EB characterization through geometric analysis and (8) diagnostic plot generation for false positive (FP) determination. Additional details on these steps can be found in Papers I, II, III, and IV. For inclusion in this Catalog we accept bonafide EBs and systems that clearly exhibit binarity through photometric analysis (heartbeats and ellipsoidals (Section 8.1). Throughout the Catalog and online database we use a system of subjective flagging to label and identify characteristics of a given system that would otherwise be difficult to validate quantitatively or statistically. Examples of these flags and their uses can be seen in Section 8. Although best efforts have been taken to provide accurate results, we caution that not all systems marked in the Catalog are guaranteed to be EB systems. There remains the possibility that some grazing EB signals may belong to small planet candidates or are contaminated by non-target EB signals. An in-depth discussion on Catalog completeness is presented in Section 10.

In this release, we have updated the Catalog in the following ways:

  • 1.  
    The light curves of Kepler Objects of Interest (KOIs) once withheld as possibly containing planetary transit events but since rejected have been included.
  • 2.  
    An increased baseline allowed for ETVs to be determined and therefore a greater precision of all ephemerides (Orosz 2012; Conroy et al. 2014). Systems previously having indeterminable periods were re-examined and included, if additional eclipses were observed.
  • 3.  
    Period and BJD0 error estimates are provided across the Catalog. The period error analysis is derived from error propagation theory and applied through an adaptation of the Period Error Calculator algorithm of Mighell & Plavchan (2013) and BJD0 errors are estimated by fitting a Gaussian to the bisected primary eclipse.
  • 4.  
    Quarter Amplitude Mismatch (QAM) systems have been rectified by scaling the affected season(s) to the season with the largest amplitude; a season being one of the four rotations per year to align the solar arrays. Once corrected, the system was reprocessed by the standard pipeline.
  • 5.  
    An additional 13 systems identified by the independent Eclipsing Binary Factory (EBF) pipeline (Section 3.2) have been processed and added to the Catalog.
  • 6.  
    Additional systems identified by Planet Hunters (Section 3.3), a citizen science project that makes use of the Zooniverse toolset to serve flux-corrected light curves from the public Kepler data set, have been added.
  • 7.  
    All systems were investigated to see if the signal was coming from the target source or a nearby contaminating signal. If a target was contaminated, the contaminated target was removed and the real source, if a Kepler target, was added to the Catalog (Section 4). A follow-up paper (M. Abdul-Masih et al. 2016, in preparation) will address the EBs that are not Kepler targets, but whose light curves can be recovered from target pixel files (TPFs) to a high precision and fidelity.
  • 8.  
    Long-cadence exposure causes smoothing of the light curves due to ∼30-minutes integration times (Section 5). Deconvolution of the phased long-cadence data polyfit allowed removal of integration smoothing, resulting in a better representation of the actual EB signal.
  • 9.  
    Classifications were done via LLE (Section 6), a general nonlinear dimensionality reduction tool, to give a value between 0 and 1, to represent detachedness as in Matijevič et al. (2012).
  • 10.  
    Visualization of the data set through tSNE (Maaten & Hinton 2008) has been provided to reveal global and local structure of similarity between systems (Section 7).
  • 11.  
    Threshold Crossing Events have been manually vetted for additional EB signals resulting in additions to the Catalog.
  • 12.  
    Follow-up data are provided where applicable. This may consist of spectroscopic data (Section 9) or additional follow-up photometric data.

3. CATALOG ADDITIONS

The previous release of the Catalog (Paper II) contained 2165 objects with EB and/or ellipsoidal variable signatures, through the second Kepler data release (Q0-Q2). In this release, 2878 objects are identified and analyzed from the entire data set of the primary Kepler mission (Q0-Q17). All transit events were identified by the main Kepler pipeline (Jenkins 2002; Jenkins et al. 2010) and through the following sources explained here.

3.1. Rejected KOI Planet Candidates

The catalog of KOIs (Mullally et al. 2015) provides a list of detected planets and planet candidates. There is an inevitable overlap in attributing transit events to planets, severely diluted binaries, low mass stellar companions, or grazing EBs. As part of the Kepler working group efforts, these targets are vetted for any EB-like signature, such as depth changes between successive eclipses (the so-called even–odd culling), detection of a secondary eclipse that is deeper than what would be expected for an R < 2RJup planet transit (occultation culling), hot white dwarf transits (Rowe et al. 2010; white dwarf culling), spectroscopic follow-up where large amplitudes or double-lined spectra are detected (follow-up culling), and automated vetting programs known as "robovetters" (J. Coughlin et al. 2016, in preparation). High-resolution direct imaging (AO and speckle) and photo-center centroid shifts also indicate the presence of background EBs. These systems can be found in the online Kepler EB catalog by searching the "KOI" flag.

3.2. EBF

The EBF (Parvizi et al. 2014) is a fully automated, adaptive, end-to-end computational pipeline used to classify EB light curves. It validates EBs through an independent neural network classification process. The EBF uses a modular approach to process large volumes of data into patterns for recognition by the artificial neural network. This is designed to allow archival data from time-series photometric surveys to be direct input, where each module's parameters are tunable to the characteristics of the input data (e.g., photometric precision, data collection cadence, flux measurement uncertainty) and define the output options to produce the probability that each individual system is an EB. A complete description can be found in Stassun et al. (2013). The neural network described here was trained on previous releases of this Catalog. The EBF identified 68 systems from Quarter 3 data. Out of the 68 systems submitted to our pipeline only 13 were validated and added to the Catalog.

3.3. Planet Hunters

Planet Hunters is a citizen science project (Fischer et al. 2012) that makes use of the Zooniverse toolset (Lintott et al. 2008) to serve flux-corrected light curves from the Kepler public release data. This process is done manually by visual inspection of each light curve for transit events. For a complete description of the process see Fischer et al. (2012). Identified transit events not planetary in nature are submitted to our pipeline for further vetting and addition to the Catalog.

3.4. Increased Baseline Revisions

An increased timespan allows the ephemerides for all EB candidates to be determined to a greater precision. All ephemerides have been manually vetted, but for certain systems it was impossible to uniquely determine the periods, i.e., for systems with equal depth eclipses (versus a single eclipse at half-period). To aid in this process we computed periodograms using three methods: Lomb–Scargle (Lomb 1976; Scargle 1982), Analysis of Variance (Schwarzenberg-Czerny 1989), and Box-fitting Least Squares (BLS) (Kovács et al. 2002), as implemented in the vartools package (Hartman 2012). Systems with equal depth eclipses may be revised in the future with additional follow-up data (Section 9).

3.5. Period and Ephemeris Error Estimates

We now provide error estimates on both the period and time of eclipse (${\mathrm{BJD}}_{0}$) for every EB in the catalog. The period error is determined through an adaptation of the Period Error Calculator algorithm of Mighell & Plavchan (2013). Using error propagation theory, the period error is calculated from the following parameters: timing uncertainty for a measured flux value, the total length of the time series, the period of the variable, and the maximum number of periods that can occur in the time series.

To revise a precise value and estimate the uncertainty on ${\mathrm{BJD}}_{0}$, we use the eclipse bisectors. The bisectors work by shifting the ${\mathrm{BJD}}_{0}$ until the left and the right eclipse sides of the phased data overlap as much as possible. The overlap function is fitted by a Gaussian, where the mean is the ${\mathrm{BJD}}_{0}$ estimate, and the width is the corresponding error. For those in which this estimates an error larger than the measured width of the eclipse, generally due to extremely low signal to noise, we assume the width of the eclipse as the error instead. Since there is not a well defined ${\mathrm{BJD}}_{0}$ for binary heartbeat stars that do not exhibit an actual eclipse in the light curve, we do not estimate or provide uncertainties for these objects.

4. CATALOG DELETIONS

In order to provide the EB community with the most accurate information, we checked each of the 2878 systems against extrinsic variability (i.e., third light contamination, cross talk, and other instrumental artifacts). We generated diagnostic plots for each available quarter of data for every object in the Kepler EB catalog to confirm whether the target was the true source of the EB signal; if the target was not the source of the signal it was deemed a FP.

To do this, we design and generate two diagnostic plots per quarter that—when considered together with all other data—give insight into which object in the TPF map is responsible for the binary signal observed in the light curve. Figures 1 and 2 show a heat map (left panels) that depicts the correlation of the eclipse depth of each individual pixel within the Kepler aperture (outlined in black) compared to the eclipse depth of the summed light curves of all the pixels in the aperture. The color scale of the heat map is normalized to the summed Kepler light curve divided by the number of pixels in the aperture; the average of pixels in the mask is assigned to white. The heat map shows where the signal contribution is located. Red colors represent pixels with a higher signal contribution while blue colors represent pixels with a lower signal contribution than the average-value pixels. A green circle represents the target KIC that is being processed while any other KIC objects in the frame are plotted with red circles. The coordinates of these circles come from KIC catalog positions. The radius of the circles correspond to the Kepler magnitude. The collection of white dots, typically located near the green circle, represent the location and shape of the centroid. The centroid is the center of light in the pixel window at a given time during the quarter.

Figure 1.

Figure 1. Diagnostic plots for one quarter. In this example the target in question (KIC 3342425) is responsible for the binary signal. In the left-hand plot we see that the pixel contributing the most flux is under the target. On the right-hand side we observe the centroid movement in time. The dispersion in the y-direction (for this particular system) is due to the target eclipsing, causing the flux to decrease and the centroid to migrate toward the brighter star, which can be seen in the upper part of the left-hand plot.

Standard image High-resolution image

The second diagnostic plot (Figures 1 and 2, right panels) illustrates an expanded view of the centroid movement throughout a given quarter. The varying colors of each circle edge represent the transition in time from the beginning of the quarter (blue) to the end of the quarter (red). The gray-scale color of each circle represents the detrended flux of the light curve at that moment, with darker shades representing lower flux. As the binary eclipses, the flux decreases and the centroid migrates toward areas of higher flux (typically away from the binary source). This, in conjunction with the heat map, gives insight into which object is responsible for the binary activity.

Figure 2.

Figure 2. Diagnostic plots for one quarter. In this example the target in question (KIC 3338674) is not responsible for the binary signal seen. On the left-hand side we see that the pixels contributing the most signal are not associated with our target. On the right-hand side we see that the centroid is constantly pulled elsewhere while only returning to our target when the off-target star is in eclipse. The binary light curve is generated by KIC 3338660, making KIC 3338674 a false positive.

Standard image High-resolution image

These diagnostic plots, in conjunction with other data, were analyzed and EB Working Group members voted on which object in the window was responsible for the binary signal. This was a blind vote with the results tabulated continuously. If the decision was not unanimous the targets were discussed in open group sessions using additional resources from data validation (DV) disposition reports (Coughlin et al. 2014) when available at NExScI.37

In addition to checking extrinsic variability, ephemeris cross-matching was performed across the Catalog to identify more FPs. This cross-matching solved ambiguity between double and half-period systems (for systems whose primary and secondary shape are identical). Any matches found this way were manually inspected to identify the true source and remove the FP.

5. LIGHT CURVE DECONVOLUTION

Due to Kepler's long-cadence 30 minute exposure, the phased light curves suffer from a convolution effect. For short-period binaries this phase-dependent smoothing can have a significant impact on the overall shape of the polyfit representation, which would then propagate through the remaining steps of the pipeline. To mitigate this effect, we deconvolve the original polyfit that was determined from the phased long-cadence data. Since there are an infinite number of functions that would convolve to the original polyfit, we impose that the deconvolved representation must also be described by a polyfit. This does not necessarily guarantee a unique solution, but does add the constraint that the deconvolved curve resembles the signal from a binary. We start with the original polyfit and use a downhill simplex algorithm to adjust the various coefficients and knots, minimizing the residuals between the original polyfit and the convolved candidate-polyfit. This process results in another polyfit that, when convolved with a 30 minute boxcar, most closely resembles the original polyfit and, therefore, the phased long-cadence data (Figure 3). This results in a better representation of the actual light curve of the binary which can then be used to estimate geometrical properties of the binary.

Figure 3.

Figure 3. Original polyfit (dashed line) and deconvolved polyfit (solid line) plotted on top of short-cadence (light dots) and long-cadence (darker x) Q2 data of KIC 11560447 and 6947064. In both cases, the deconvolution was successful in finding a polyfit which when convolved best fits the long-cadence data, but this does not necessarily fit the short-cadence data or make physical sense.

Standard image High-resolution image

Kepler's short-cadence photometry (1 minute exposures) is short enough that the convolution effect is negligible. However, short cadence data are not available for all systems. Thus, to stay internally consistent, we use only long-cadence data to determine physical parameters and ETVs. Nevertheless, short-cadence data allow us to confirm the effect of convolution and test the performance of our deconvolution process. Comparing the deconvolved polyfit with the short-cadence data from the same EB shows that deconvolution is essential for more accurate approximations but can also result in a representation that does not make physical sense (see the right panel in 3). For this reason, if the deconvolved polyfit for a short-period EB looks suspicious, the deconvolution process likely introduced undesired artifacts into the light curve representation.

6. CLASSIFICATION OF LIGHT CURVES

EB light curves come in a variety of different shapes that are governed by a number of parameters. In order to overcome the drawbacks of manual classification, we performed an automated classification of all identified EB light curves with a general dimensionality reduction numerical tool called LLE (Roweis & Saul 2000). This method is able to project a high-dimensional data set onto a much lower-dimensional manifold in such a way that it retains the local properties of the original data set, so all light curves that are placed close together in the original space are also nearby each other in the projected space (Paper III). This way the relations between the light curves are much easier to investigate. The input space is represented by a collection of polyfit models corresponding to each of the EB light curves in the Catalog. All polyfits were calculated in 1000 equidistant phase points (representing the original space) and then projected to a two-dimensional space, resulting in an arc-shaped manifold. Inspection of the underlying light curves along the arc revealed that one end of the arc was populated by well-detached binaries while the other hosted the overcontact binaries and systems with ellipsoidal variations. Semi-detached systems were in the middle. We assigned a single number to each of the light curves based on where on the arc their projection is located to provide an easy-to-use single-number quantitative representation of the light curves. Values of the parameter range from 0 to 0.1 for well-detached systems, values below 0.5 predominantly belong to detached systems and between 0.5 and 0.7 to semi-detached systems. Overcontact systems usually have values between 0.7 and 0.8, while even higher values up to 1 usually belong to ellipsoidal variables.

In addition to the classification number, we also provide the depths and widths of the primary and secondary eclipses, as well as the separations between eclipses from the polyfit. Depths are determined from the amplitude of the polynomial fitted to the eclipse and are in units of normalized flux. A depth measurement is provided so long as it is larger than three times the estimated scatter of the light curve baseline. For this reason, values are not provided for a secondary eclipse if the signal to noise is not sufficient. Widths are determined by the "knots" connecting the individual polynomial sections and are in units of phase. Separations are determined as the distance between the primary and secondary minimum, also in units of phase, and are defined to always be less than or equal to 0.5. These values are only approximate measures and are only as accurate as the polyfits themselves. Nevertheless, they do provide some measure of the signal-to-noise and also allow for an estimate of eccentricities (Prša et al. 2015).

7. VISUALIZING KEPLER DATA

Visualization is an important aspect of data mining—it allows a more intuitive and interactive approach to analyzing and interpreting the data set. The employed algorithm, t-Distributed Stochastic Neighbor Embedding (t-SNE; Maaten & Hinton 2008) is particularly useful for high-dimensional data that lie on several different, but related, lower-dimensional manifolds. It allows the user to simultaneously view EB properties from multiple viewpoints.

7.1. The t-SNE Algorithm

An extension to the EB classification with LLE has been performed using a new technique of visualizing high-dimensional data, first proposed and developed by Maaten & Hinton (2008). This technique, called t-SNE, is a modified version of the Stochastic Neighbor Embedding technique and has a specific appeal for visualizing data, since it is capable of revealing both global and local structure in terms of clustering data with respect to similarity. We provide a brief overview of the basic principles of the t-SNE technique, the results of its application to Kepler data and its implementation in the interactive visualization of the Catalog.

Stochastic Neighbor Embedding defines data similarities in terms of conditional probabilities in the high-dimensional data space and their low-dimensional projection. Neighbors of a data-point in the high-dimensional data space are picked in proportion to their probability density under a Gaussian. Therefore, the similarity of two data-points is equivalent to a conditional probability. In t-SNE, the conditional probability is replaced by a joint probability that depends upon the number of data-points. This ensures that all data-points contribute to the cost function by a significant amount, including the outliers. The conditional probability of the corresponding low-dimensional counterparts in SNE is also defined in terms of a Gaussian probability distribution, but t-SNE has introduced a symmetrized Student t-distribution, which leads to joint probabilities of the map. This allows for a higher dispersion of data-points in the low-dimensional map and avoids unwanted attractive forces, since a moderate distance in the high-dimensional map can be represented well by larger distances in its low-dimensional counterpart.

An input parameter that defines the configuration of the output map is the so-called perplexity. Therefore, instead of providing the desired number of nearest neighbors, the user provides a desired value of the perplexity and leaves it up to the method to determine the number of nearest neighbors, based on the data density. This in turn means that the data itself affects the number of nearest neighbors, which might vary from point to point.

7.2. t-SNE and Kepler Data

We have applied t-SNE to various samples of Kepler data and compared the results to other methods (LLE, manual flags). We have also attempted two different approaches to the two-dimensional projection: one with direct high-dimensional to two-dimensional mapping and one equivalent to the LLE classification method already implemented in the Catalog (Paper III), where the high-dimensional data space is first mapped to a 3D projection, and subsequently mapped to a 2D projection. The latter approach performs better in terms of clustering data-points and reveals a lumpier data structure, while the first approach results in a rather continuous structure, which might render a more elegant visualization of certain parameter distributions over the projection.

Both the two-step and the direct projection seem to be in agreement with the results obtained with LLE. One advantage of LLE over t-SNE is its simple structure, which allowed for a quantitative classification in terms of the morphology parameter (Paper III). The complex two-dimensional structure of the t-SNE projection makes that task substantially more difficult, so we use t-SNE only as a visualization tool, and retain LLE for classification.

The distribution of the LLE morphology parameter on the t-SNE projection is depicted in Figure 4. The concordance between the two methods is evident from the continuous variation of the morphology parameter along the map. It also serves to further illustrate the performance of t-SNE in terms of large-scale and small-scale structure. For example, the two islands in the right corner of the two-step projection (Figure 4, panel (a)), which incorporate mainly noisy and unique light curve data, still manage to retain the large-scale gradient of the morphology parameter directed from the bottom to the top of the projection.

Figure 4.

Figure 4. Distribution of the LLE morphology parameter over the two-step (panel (a)) and direct 2D t-SNE projection (panel (b)).

Standard image High-resolution image

Further comparison also shows the two-step t-SNE ability to distinguish the main types of EBs: objects with a morphology parameter c (Paper III) $\lesssim 0.4$, corresponding to detached binaries, are a separate cluster from those with $0.4\lesssim c\lesssim 0.8$ and those with $0.7\lesssim c\lesssim 0.8$. Those with $c\gtrsim 0.8$, corresponding to ellipsoidal variations and objects with uncertain classification, are grouped into the topmost cluster, separate from all the previous ones. Based on this, it is safe to conclude that t-SNE also performs in line with the manual classification and it might be useful in speeding up or even replacing the subjective manual classification process.

To further illustrate the performance of the technique, we have generated plots of all cataloged parameters over the projection. In all cases the distribution seems fairly smooth, which attests to the broad range of technique applicability. Plots of the distribution of primary eclipse width and secondary eclipse depth obtained with polyfit, and the ratio of temperatures ($T2/T1$) and $\mathrm{sin}i$ obtained through geometric analysis, are provided in Figure 5 (panels (a)–(d) respectively).

Figure 5.

Figure 5. Distribution of primary eclipse width (panel (a)), secondary eclipse depth (panel (b)), $T2/T1$ (panel (c)) and $\mathrm{sin}i$ (panel (d)) parameters over the direct 2D t-SNE projection. The data points that do not have a value for the desired parameter are marked in gray.

Standard image High-resolution image

An example of the power of tSNE to reveal substructure and distinguish between many different types of similarities which may arise from parameters such as inclination, primary eclipse widths, secondary eclipse depth etc., within a given classification can be seen in the "branches" of Figure 4 (panel (b)) and revealed in Figure 5. The two groups of light curves corresponding to ellipsoidal variations ($c\gtrsim 0.8$) at the bottom of Figure 4 (panel (b)) are especially indicative of this, since only the left branch corresponds to manually classified ellipsoidal light curves, while the right branch, although having a morph parameter that corresponds to ellipsoidal curves, is in fact composed of noisy or unique light curves that can not be classified into any of the primary morphological types. Most of the heartbeat stars can be found at the bottom of this branch.

To fully benefit from the capabilities of the technique, we implemented an interactive version in the Catalog, which combines the t-SNE visualization of Kepler data with a broad range of parameter distributions. This extends the data possibilities to a more intuitive and interactive approach, enabling the user to directly view the properties of certain light curves with respect to the whole Catalog. This can be found at http://keplerEBs.villanova.edu/tsne.

8. INTERESTING CLASSES OF OBJECTS IN THE CATALOG

8.1. Heartbeat Stars

Heartbeat stars are a subclass of eccentric ellipsoidal variables introduced by Thompson et al. (2012). The most prominent feature in the heartbeat star light curve is the increased brightness at periastron (where periastron is defined as the closest approach of the two binary star components) caused by stellar deformation, which is a consequence of gravitational interactions; and heating (Figure 6). The morphology of the heartbeat-star light curve defines the heartbeat star—a periastron variation preceded and succeeded by a flat region (ignoring pulsations and spots etc.). As shown by Kumar et al. (1995), the shape of the brightening primarily depends on three orbital properties: argument of periastron, eccentricity, and inclination. This is true for the majority of objects where irradiation effects cause only minor modifications (Burkart et al. 2012). However, this effect is temperature dependent and so for objects with hotter components that both contribute significantly to the flux, as in the case of KOI-54 (Welsh et al. 2011), which contains two A stars, the irradiation contributes notably to the light curve (approximatly half of the amplitude of the periastron variation of KOI-54 is a consequence of irradiation). Overall, the amplitude of the periastron brightening depends on the temperature and structure of the stellar components, and the periastron distance. This enables heartbeat stars to be modeled without the presence of eclipses and thus at any inclination. Table 1 contains the Kepler catalog identifiers (KIC) and corresponding periods for 173 currently known heartbeart stars in the Kepler sample. These systems are flagged with the "HB" flag in the Catalog.

Figure 6.

Figure 6. Time series of a selection of Kepler heartbeat star light curves (left panel) and phase folded data (right panel). The phase folded data clearly depict the oscillations that are integer multiples of the orbital frequency: tidally induced pulsations. It is worth noting that KIC 8112039 is KOI 54 (Welsh et al. 2011).

Standard image High-resolution image

Table 1.  The Heartbeat Stars in the Kepler Sample

KIC Period R.A.(°) Decl.(°)
  (days) (J2000) (J2000)
1573836 3.557093 291.5025 37.1775
2010607 18.632296 290.5056 37.4590
2444348 103.206989 291.6687 37.7041
2697935 21.513359 287.4679 37.9666
2720096 26.674680 292.9791 37.9109
3230227 7.047106 290.1126 38.3999
3240976 15.238869 292.9291 38.3280
3547874 19.692172 292.2829 38.6657
3729724 16.418755 285.6555 38.8505
3734660 19.942137 287.6238 38.8554
3749404 20.306385 292.0795 38.8371
3764714 6.633276 295.7400 38.8623
3766353 2.666965 296.0538 38.8943
3850086 19.114247 291.3975 38.9647
3862171 6.996461 294.4970 38.9801
3869825 4.800656 296.1339 38.9990
3965556 6.556770 294.3909 39.0767
4142768 27.991603 287.2629 39.2600
4150136 9.478402 289.6193 39.2939
4247092 21.056416 286.8797 39.3784
4248941 8.644598 287.6032 39.3977
4253860 155.061112 289.2337 39.3865
4359851 13.542328 290.0829 39.4008
4372379 4.535171 293.5203 39.4478
4377638 2.821875 294.7438 39.4994
4450976 12.044869 287.3733 39.5261
4459068 24.955995 290.0244 39.5634
4470124 11.438984 293.1966 39.5072
4545729 18.383520 286.4008 39.6984
4649305 22.651138 290.1033 39.7816
4659476 58.996374 293.1102 39.7564
4669402 8.496468 295.5339 39.7623
4761060 3.361391 295.5471 39.8618
4847343 11.416917 294.9517 39.9294
4847369 12.350014 294.9584 39.9046
4936180 4.640922 295.1137 40.0712
4949187 11.977392 297.9250 40.0884
4949194 41.263202 297.9279 40.0549
5006817 94.811969 290.4560 40.1457
5017127 20.006404 293.5486 40.1117
5034333 6.932280 297.4004 40.1495
5039392 236.727941 298.4009 40.1722
5090937 8.800693 289.2657 40.2555
5129777 26.158530 299.0478 40.2189
5175668 21.882115 288.3171 40.3280
5213466 2.819311 298.1299 40.3999
5284262 17.963312 294.4163 40.4802
5286221 15.295983 294.9404 40.4966
5398002 14.153175 300.0150 40.5270
5511076 6.513199 282.9722 40.7270
5596440 10.474857 282.7207 40.8992
5707897 8.416091 292.5603 40.9969
5733154 62.519903 298.6118 40.9491
5736537 1.761529 299.2878 40.9129
5771961 26.066437 284.7599 41.0434
5790807 79.996246 291.7718 41.0981
5818706 14.959941 298.6959 41.0417
5877364 89.648538 292.1599 41.1982
5944240 2.553222 285.9139 41.2916
5960989 50.721534 292.0520 41.2661
6042191 43.390923 291.7434 41.3100
6105491 13.299638 284.9915 41.4333
6117415 19.741625 289.8611 41.4082
6137885 12.790099 295.8830 41.4747
6141791 13.659035 296.7550 41.4846
6290740 15.151827 293.6252 41.6615
6292925 13.612220 294.2413 41.6198
6370558 60.316584 293.8644 41.7169
6693555 10.875075 292.4559 42.1372
6775034 10.028547 291.0820 42.2686
6806632 9.469157 298.8907 42.2099
6850665 214.716056 287.7663 42.3227
6881709 6.741116 296.6588 42.3698
6963171 23.308219 295.8099 42.4592
7039026 9.943929 293.4486 42.5866
7041856 4.000669 294.2317 42.5901
7050060 22.044000 296.2551 42.5302
7259722 9.633226 283.1575 42.8904
7293054 671.800000 295.0398 42.8795
7350038 13.829942 287.5149 42.9118
7373255 13.661106 294.8559 42.9372
7431665 281.400000 287.3766 43.0095
7511416 5.590855 286.0600 43.1662
7591456 5.835751 285.6098 43.2055
7622059 10.403262 295.9200 43.2818
7660607 2.763401 281.8825 43.3002
7672068 16.836177 287.8638 43.3042
7799540 60.000000 281.1918 43.5250
7833144 2.247734 295.3246 43.5054
7881722 0.953289 288.2734 43.6977
7887124 32.486427 290.4833 43.6224
7897952 66.991639 294.2019 43.6553
7907688 4.344837 297.0531 43.6438
7914906 8.752907 298.8165 43.6733
7918217 63.929799 299.6020 43.6960
7973970 9.479933 296.3820 43.7489
8027591 24.274432 291.3655 43.8680
8095275 23.007350 290.9708 43.9708
8112039 41.808235 296.5647 43.9476
8123430 11.169990 299.3394 43.9996
8144355 80.514104 281.8770 44.0299
8151107 18.001308 285.5886 44.0477
8164262 87.457170 291.2468 44.0004
8197368 9.087917 300.9348 44.0943
8210370 153.700000 280.7721 44.1887
8242350 6.993556 294.9346 44.1298
8264510 5.686759 300.9686 44.1920
8322564 22.258846 298.7603 44.2662
8328376 4.345967 300.3823 44.2801
8386982 72.259590 298.4551 44.3134
8456774 2.886340 299.7266 44.4091
8456998 7.531511 299.7784 44.4611
8459354 53.557318 300.4068 44.4142
8508485 12.595796 296.4474 44.5586
8688110 374.546000 291.3925 44.8952
8696442 12.360553 294.5535 44.8503
8702921 19.384383 296.6650 44.8531
8703887 14.170980 296.9350 44.8503
8707639 7.785196 297.8877 44.8683
8719324 10.232698 301.1143 44.8258
8803882 89.630216 284.7969 45.0991
8838070 43.362724 298.1911 45.0936
8908102 5.414582 298.8360 45.1165
8912308 20.174443 300.0684 45.1016
9016693 26.368027 289.8839 45.3041
9151763 438.051939 290.6850 45.5684
9163796 121.006844 295.3375 45.5048
9408183 49.683544 293.4932 45.9876
9535080 49.645296 295.2635 46.1480
9540226 175.458827 297.0340 46.1985
9596037 33.355613 294.9424 46.2501
9701423 8.607397 287.6039 46.4128
9711769 12.935909 292.4648 46.4918
9717958 67.995354 294.9158 46.4853
9790355 14.565548 298.7419 46.5773
9835416 4.036605 293.6965 46.6205
9899216 10.915849 295.1617 46.7506
9965691 15.683195 297.3268 46.8452
9972385 58.422113 299.2823 46.8988
10004546 19.356914 288.7470 46.9307
10092506 31.041675 297.8457 47.0723
10096019 6.867469 298.8533 47.0828
10159014 8.777397 297.7946 47.1565
10162999 3.429215 298.8757 47.1436
10221886 8.316718 296.8986 47.2816
10334122 37.952857 289.6644 47.4004
10611450 11.652748 296.1244 47.8307
10614012 132.167312 296.9287 47.8830
10664416 25.322174 291.4008 47.9246
10679505 5.675186 296.9342 47.9812
10863286 3.723867 292.7082 48.2070
10873904 9.885633 296.6001 48.2960
11044668 139.450000 297.9449 48.5578
11071278 55.885225 284.4661 48.6553
11122789 3.238154 283.1246 48.7349
11133313 27.400893 289.6666 48.7296
11240948 3.401937 290.1068 48.9034
11288684 22.210063 287.1591 49.0800
11403032 7.631634 291.9279 49.2970
11409673 12.317869 295.1364 49.2733
11494130 18.955414 284.0268 49.4154
11506938 22.574780 291.6465 49.4934
11568428 1.710629 296.1853 49.5568
11568657 13.476046 296.2798 49.5178
11572363 19.027753 297.7151 49.5738
11649962 10.562737 284.5680 49.7795
11700133 6.754017 284.2339 49.8532
11769801 29.708220 295.1429 49.9931
11774013 3.756248 296.8993 49.9163
11859811 22.314148 289.3813 50.1894
11923629 17.973284 296.3860 50.2050
11970288 20.702319 295.2126 50.3169
12255108 9.131526 289.5991 50.9386

Download table as:  ASCIITypeset images: 1 2 3

Table 2.  The Systems With Tidally Induced Pulsations in the Kepler Sample

KIC Period R.A.(°) Decl.(°)
  (days) (J2000) (J2000)
3230227 7.047106 290.1126 38.3999
3547874 19.692172 292.2829 38.6657
3749404 20.306385 292.0795 38.8371
3766353 2.666965 296.0538 38.8943
3869825 4.800656 296.1339 38.9990
4142768 27.991603 287.2629 39.2600
4248941 8.644598 287.6032 39.3977
4949194 41.263202 297.9279 40.0549
5034333 6.932280 297.4004 40.1495
5090937 8.800693 289.2657 40.2555
8095275 23.007350 290.9708 43.9708
8112039 41.808235 296.5647 43.9476
8164262 87.457170 291.2468 44.0004
8264510 5.686759 300.9686 44.1920
8456774 2.886340 299.7266 44.4091
8703887 14.170980 296.9350 44.8503
8719324 10.232698 301.1143 44.8258
9016693 26.368027 289.8839 45.3041
9835416 4.036605 293.6965 46.6205
9899216 10.915849 295.1617 46.7506
11122789 3.238154 283.1246 48.7349
11403032 7.631634 291.9279 49.2970
11409673 12.317869 295.1364 49.2733
11494130 18.955414 284.0268 49.4154

Download table as:  ASCIITypeset image

As the stellar components are in close proximity at periastron (relative to their radii), they are subject to strongly varying gravitational forces. Consequently, the stars experience large torques during periastron flybys, making them likely candidates for apsidal motion: the precession of the line of apsides about the center of mass (Claret & Gimenez 1993). Since apsidal motion changes the argument of periastron, the advance can be determined by the change in the shape of the heartbeat light curve, assuming that the data set is long enough for the change to be detected. Figure 7 shows an example of how the light curve of a heartbeat star changes as a function of the argument of periastron for the case of KIC 3749404, which exhibits one of the largest periastron advance rates of ∼2° yr−1. The central density of the stellar components can be empirically inferred through the rate of the apsidal advance (Claret & Giménez 2010), enabling us to test our current understanding of stellar structure and evolution. Furthermore, as approximately half of all the known Kepler heartbeat stars have periods $3\lt P\lt 15\;\mathrm{days}$ and high eccentricities ($e\gt 0.3$), heartbeat stars are ideal for testing theories on relativistic apsidal motion (Gimenez 1985).

Figure 7.

Figure 7. Effect of the periastron value, ω, on the shape of the periastron brightening for one full cycle. The peaks change both their position in phase and their shape drastically with varying ω. KIC 3749404 is a heartbeat star that exhibits one of the largest periastron advance rates (∼2° yr−1; K. Hambleton et al. 2016, in preparation). In the subplot the effect on the light curve over the four year Kepler mission is depicted.

Standard image High-resolution image

8.1.1. Tidally Induced Pulsations

Approximately 15% of these interesting objects also demonstrate tidally induced pulsations, e.g., HD174884 (Maceroni et al. 2009), HD177863 (Willems & Aerts 2002), KOI-54 (Welsh et al. 2011; Burkart et al. 2012; Fuller & Lai 2012), and KIC 4544587 (Hambleton et al. 2013). The signature of a tidally induced pulsation is a mode (or modes) that is an integer multiple of the orbital frequency. From our sample, the record holder, KIC 8164262, has a tidally induced pulsation amplitude of 1 mmag. The majority of objects in our sample have tidally induced pulsations with maximum amplitudes of ∼0.1 mmag. Tidally induced modes occur when the orbital frequency is close to an eigenfrequency of the stellar component, causing the star to act like a forced oscillator, which significantly increases the amplitude of the mode. Zahn (1975) hypothesized the existence of tidally induced pulsations as a mechanism for the circularization of binary orbits, attributed to the exchange of orbital angular momentum and loss of orbital energy through mode damping. The high occurrence rate of tidally induced pulsations in heartbeat stars is due to their eccentric nature and small periastron distances leading to interaction times that are comparable to the eigenmodes of the stellar components. The right panel of Figure 6 shows a selection of heartbeat star light curves folded on their orbital periods. The stellar pulsations are clearly visible as they are exact integer multiples of the orbital frequency: the signature of tidally induced pulsations. These 24 systems are flagged with the "TP" flag.

8.2. Reflection Effect Binaries

The reflection effect is the mutual irradiation of the facing hemispheres of two stars in the binary system. This irradiation alters the temperature structure in the atmosphere of the star, resulting in an increased intensity and flux. This effect reveals itself with an increased flux level on the ingress and egress of the eclipse in the light curve. There are currently 36 of these systems (Table 3). These systems are flagged with the "REF" flag.

Table 3.  The Reflection Effect Systems in the Kepler Sample

KIC Period Period Error BJD0 BJD0 Error R.A.(°) Decl.(°)
  (days) (days) (−2400000) (days) (J2000) (J2000)
2708156 1.891272 0.000003 54954.336 0.046 290.2871 37.9365
3339563 0.841232 0.000001 54965.378 0.051 290.7460 38.4431
3431321 1.015049 0.000001 54954.418 0.066 287.7745 38.5040
3547091 3.30558 0.00001 55096.3221 0.0177 292.0785 38.6315
4350454 0.965658 0.000001 54964.943 0.224 287.0525 39.4879
4458989 0.529854 0.000000 54954.1586 0.0337 290.0015 39.5339
5034333 6.93228 0.00002 54954.028 0.054 297.4004 40.1495
5098444 26.9490 0.0001 54984.023 0.082 291.5633 40.2678
5213466 2.81931 0.00001 55165.534 298.1299 40.3999
5736537 1.76153 0.00001 54965.955 299.2878 40.9129
5792093 0.600588 0.000000 54964.6434 0.0331 292.2002 41.0137
6262882 0.996501 0.000001 54965.112 0.064 282.6457 41.6087
6387887 0.216900 0.000001 54999.9686 0.0056 298.1854 41.7340
6791604 0.528806 0.000000 54964.6825 0.0284 295.6755 42.2406
7660607 2.76340 0.00001 54954.589 281.8825 43.3002
7748113 1.734663 0.000002 54954.144 0.102 290.1135 43.4365
7770471 1.15780 0.00000 55000.1826 0.0263 297.1805 43.4771
7833144 2.24773 0.00001 55001.232 295.3246 43.5054
7881722 0.953289 0.000002 54954.118 288.2734 43.6977
7884842 1.314548 0.000002 54954.876 0.082 289.6946 43.6260
8455359 2.9637 0.0002 55002.027 0.189 299.3618 44.4112
8758161 0.998218 0.000001 54953.8326 0.0243 293.6205 44.9673
9016693 26.3680 0.0001 55002.583 289.8839 45.3041
9071373 0.4217690 0.0000003 54953.991 0.109 283.1591 45.4156
9101279 1.811461 0.000003 54965.932 0.047 296.0016 45.4481
9108058 2.1749 0.0001 54999.729 0.049 297.7811 45.4243
9108579 1.169628 0.000001 54954.998 0.069 297.8999 45.4666
9159301 3.04477 0.00001 54956.303 0.063 293.6933 45.5170
9472174 0.1257653 0.0000001 54953.6432 0.0183 294.6359 46.0664
9602595 3.55651 0.00001 54955.853 0.074 297.1435 46.2285
10000490 1.400991 0.000002 54954.2437 0.0286 286.5560 46.9573
10149845 4.05636 0.00001 55001.071 0.237 295.0444 47.1964
10857342 2.41593 0.00003 55005.265 0.056 290.0115 48.2442
11408810 0.749287 0.000001 54953.869 0.047 294.7304 49.2944
12109845 0.865959 0.000003 55000.5582 0.0129 290.9192 50.6964
12216706 1.47106 0.00001 55003.5216 0.0252 295.7438 50.8309

Note. Those reported without ${\mathrm{BJD}}_{0}$ errors are also heartbeat stars.

Download table as:  ASCIITypeset image

8.3. Occultation Pairs

The majority of the stars that make up the Milky Way galaxy are M-type stars and, since they are faint because of their low mass, luminosity, and temperature, the only direct way to measure their masses and radii is by analyzing their light curves in EBs. An issue in the theory of these low-mass stars is the discrepancy between predicted and observed radii of M-type stars (Feiden 2015). When these stars are found in binaries with an earlier type component, the ratio in radii is quite large, and eclipses are total and occulting. Other extreme radius ratio pairs produce similar light curves, such as white dwarfs and main sequence stars, or main sequence stars and giants. Irrespective of the underlying morphology, these occultation pairs are critical gauges because eclipses are total and the models are thus additionally constrained.

There are currently 32 of these systems (Table 4). These systems are flagged with the "OCC" flag.

Table 4.  The Occultation Pairs in the Kepler Sample

KIC Period Period Error BJD0 BJD0 Error R.A.(°) Decl.(°)
  (days) (days) (−2400000) (days) (J2000) (J2000)
2445134 8.4120089 0.0000228 54972.648 0.035 291.8497 37.7386
3970233 8.254914 0.000055 54966.160 0.052 295.4819 39.0173
4049124 4.8044707 0.0000102 54969.0039 0.0229 289.3142 39.1314
4386047 2.9006900 0.0000181 55001.2005 0.0253 296.4303 39.4435
4740676 3.4542411 0.0000063 54954.3478 0.0325 289.9644 39.8114
4851464 5.5482571 0.0000159 55005.061 0.046 295.8887 39.9113
5370302 3.9043269 0.0000076 54967.3507 0.0286 294.0568 40.5497
5372966 9.2863571 0.0000262 54967.6753 0.0264 294.7440 40.5339
5728283 6.1982793 0.0000153 55003.981 0.057 297.6203 40.9431
6182019 3.6649654 0.0000074 55003.8010 0.0264 282.6849 41.5880
6362386 4.5924016 0.0000094 54956.954 0.036 291.2642 41.7488
6387450 3.6613261 0.0000069 54968.2691 0.0182 298.0917 41.7941
6694186 5.5542237 0.0000124 54957.0531 0.0325 292.6311 42.1808
6762829 18.795266 0.000072 54971.668 0.079 286.8304 42.2792
7037540 14.405911 0.000063 55294.236 0.044 293.0204 42.5088
7972785 7.3007334 0.0000187 54966.566 0.048 296.0477 43.7621
8230809 4.0783467 0.0000081 54973.317 0.049 290.8917 44.1689
8458207 3.5301622 0.0000065 54967.7749 0.0309 300.0916 44.4148
8460600 6.3520872 0.0000153 54967.2065 0.0286 300.7748 44.4878
8580438 6.4960325 0.0000158 54968.468 0.037 298.3827 44.6156
9048145 8.6678260 0.0000238 54970.3408 0.0331 299.3980 45.3182
9446824 4.2023346 0.0000087 55004.1071 0.0280 282.1844 46.0069
9451127 5.1174021 0.0000111 54967.5858 0.0287 284.8227 46.0137
9649222 5.9186193 0.0000138 54965.9239 0.0285 291.8585 46.3865
9719636 3.351570 0.000044 55000.7919 0.0223 295.5262 46.4571
10020423 7.4483776 0.0000192 54970.6936 0.0289 295.2979 46.9205
10295951 6.8108248 0.0000167 54955.424 0.062 299.1028 47.3778
10710755 4.8166114 0.0000102 54966.7629 0.0271 281.9364 48.0101
11200773 2.4895516 0.0000039 54965.0434 0.0184 296.4774 48.8537
11252617 4.4781198 0.0000096 55006.1732 0.0218 295.6914 48.9038
11404644 5.9025999 0.0000137 54969.388 0.042 292.7561 49.2646
11826400 5.8893723 0.0000135 54956.000 0.037 297.5134 50.0748

Download table as:  ASCIITypeset image

8.4. Circumbinary Planets

The search for circumbinary planets in the Kepler data includes looking for transits with multiple features (Deeg et al. 1998; Doyle et al. 2000). Transit patterns with multiple features are caused by a slowly moving object crossing in front of the EB; it is alternately silhouetted by the motion of the background binary stars as they orbit about each other producing predictable but non-periodic features of various shapes and depths. Emerging trends from studying these circumbinary planet systems show the planets orbital plane very close to the plane of the binary (in prograde motion) in addition to the host star. The orbits of the planets are in close proximity to the critical radius and the planet's sizes (in mass and/or radius) are smaller than that of Jupiter. A full discussion of these trends can be found in Welsh et al. (2014). There are currently 14 of these systems (Table 5). These systems are flagged with the "CBP" flag.

Table 5.  The Circumbinary Planets in the Kepler Sample

KIC Kepler # Period Period Error R.A.(°) Decl.(°) Citation
    (days) (days) (J2000) (J2000)  
12644769 Kepler-16b 41.0776 0.0002 289.075700 51.757400 Doyle et al. (2011)
8572936 Kepler-34b 27.7958 0.0001 296.435800 44.641600 Welsh et al. (2012)
9837578 Kepler-35b 20.7337 0.0001 294.497000 46.689800 Welsh et al. (2012)
6762829 Kepler-38b 18.7953 0.0001 286.830400 42.279200 Orosz et al. (2012b)
10020423 Kepler-47b,c 7.44838 0.00002 295.297900 46.920500 Orosz et al. (2012a)
4862625 Kepler-64b 20.0002 0.0001 298.215100 39.955100 Schwamb et al. (2013)
12351927 Kepler413b 10.11615 0.00003 288.510600 51.162500 Kostov et al. (2014)
9632895 Kepler-453b 27.3220 0.0001 283.241300 46.378500 Welsh et al. (2015)

Download table as:  ASCIITypeset image

8.5. Targets with Multiple Ephemerides

Sources with additional features (8.4) can be another sign of a stellar triple or multiple system (in addition to ETV signals described in Orosz 2015 and J. Orosz et al. 2016, in preparation). In this case the depth of the event is too deep to be the transit of a planet but is instead an eclipse by, or occultation of a third stellar body. A well known example is described in Carter et al. (2011). We have been looking for such features in the Catalog and have uncovered 14 systems exhibiting multiple, determinable periods (Table 6). These systems are flagged with the "M" flag in the Catalog. In some systems, extraneous events are observed whose ephemerides cannot be determined. In some cases the period is longer than the time baseline and two subsequent events have not been observed by Kepler. In other cases, eclipsing the inner-binary at different phases results in a nonlinear ephemeris with an indeterminable period. It is worth noting that without spectroscopy or ETVs that are in agreement that additional eclipse event is indeed related, these cases are not guaranteed to be multiple objects—some could be the blend of two independent binaries on the same pixel. These events and their properties are reported in Table 7.

Table 6.  The Systems Exhibiting Multiple Ephemerides in the Kepler Sample

KIC Period Period Error BJD0 BJD0 Error R.A.(°) Decl.(°)
  (days) (days) (−2400000) (days) (J2000) (J2000)
2856960 0.2585073 0.0000001 54964.658506 0.007310 292.3813 38.0767
2856960 204.256 0.002 54997.652563 0.369952 292.3813 38.0767
4150611 8.65309 0.00002 54961.005419 0.024746 289.7425 39.2671
4150611 1.522279 0.000002 54999.688801 0.003464 289.7425 39.2671
4150611 94.198 0.001 55029.333888 0.328165 289.7425 39.2671
5255552 32.4486 0.0002 54970.636491 0.116220 284.6931 40.4986
5897826 33.8042 0.0002 54967.628858 0.116761 297.4759 41.1143
5952403 0.905678 0.000001 54965.197892 0.014736 289.2874 41.2648
6665064 0.69837 0.00001 54964.697452 0.009707 281.6500 42.1321
6964043 5.36258 0.00002 55292.008176 0.308696 296.0145 42.4223
7289157 5.26581 0.00001 54969.976049 0.044130 293.9661 42.8373
7289157 242.713 0.002 54996.317389 0.055294 293.9661 42.8373
9007918 1.387207 0.000002 54954.746682 0.023254 286.0084 45.3560
11495766 8.34044 0.00002 55009.377729 0.046025 285.2253 49.4242

Download table as:  ASCIITypeset image

Table 7.  Properties of the Extraneous Events Found in the Kepler Sample

KIC Event Depth Event Width Start Time End Time
  (%) (days) (−240000) (240000)
6543674 0.96 2. 55023 55025
7222362 0.6 0.6 55280.9 55281.5
7222362 0.8 2 55307.5 55309.5
7222362 0.65 2 55975.5 55977.5
7668648 0.94 0.2 55501.1 55501.3
7668648 0.94 0.2 55905.5 55905.7
7668648 0.94 0.2 56104.4 56104.6
7668648 0.94 0.2 56303.7 56303.9
7670485 0.975 1.0 55663 55664

Download table as:  ASCIITypeset image

8.6. ETVs

In an EB, one normally expects the primary eclipses to be uniformly spaced in time. However, apsidal motion, mass transfer from one star to the other, or the presence of a third body in the system can give rise to changes in the orbital period, which in turn will change the time interval between consecutive eclipse events. These deviations will contain important clues as to the reason for the period change. A table of these systems and more information on their analysis can be found in Conroy et al. (2014) and J. Orosz et al. (2016, in preparation). ETV values and plots are provided for each system in the online Catalog.

8.7. Eclipse Depth Changes

We have come across 43 systems that exhibit eclipse depth changes (Figure 8). These systems were visually inspected and manually flagged. The depth variations shown here differ from QAM effects; the depth variations are long-term trends spanning multiple, sequential quarters. In 10319590 (see Figure 8), for example, the eclipses actually disappear. Although it is possible that some of these could be caused by gradual aperture movement or a source near the edge of a module leaking light, these variations are not quarter or season-dependent, and are much more likely to actually be physical. Physical causes of these long-term depth variations could include a rapid inclination or periastron change due to the presence of a third body or spot activity as shown in Croll et al. 2015). Table 8 lists these systems and are flagged by the "DV" (Depth Variation) flag.

Figure 8.

Figure 8. KIC 10319590 is a system undergoing eclipse depth variations.

Standard image High-resolution image

Table 8.  The Systems Exhibiting Eclipse Depth Variations in the Kepler Sample

KIC Period Period Error BJD0 BJD0 Error R.A.(°) Decl.(°)
  (days) (days) (−2400000) (days) (J2000) (J2000)
1722276 569.95 0.01 55081.923700 0.035212 291.6967 37.2380
2697935 21.5134 0.0001 55008.469440 287.4679 37.9666
2708156 1.891272 0.000003 54954.335595 0.046299 290.2871 37.9365
3247294 67.419 0.000 54966.433454 0.046494 294.4399 38.3808
3867593 73.332 0.000 55042.964810 0.063167 295.6975 38.9020
3936357 0.3691536 0.0000002 54953.852697 0.020054 285.4109 39.0412
4069063 0.504296 0.000000 54964.906342 0.010072 294.5585 39.1377
4769799 21.9293 0.0001 54968.505532 0.061744 297.3174 39.8780
5130380 19.9830 0.0002 55009.918408 0.080312 299.1550 40.2881
5217781 564.41 0.01 55022.074166 0.109098 298.9473 40.3755
5310387 0.4416691 0.0000003 54953.664664 0.026196 299.8789 40.4824
5653126 38.4969 0.0002 54985.816152 0.067928 299.7020 40.8963
5771589 10.73914 0.00003 54962.116764 0.038736 284.5855 41.0096
6148271 1.7853 0.0001 54966.114761 0.040979 298.2297 41.4598
6197038 9.75171 0.00003 54961.776580 0.098446 289.3008 41.5265
6205460 3.72283 0.00001 54956.455222 0.098339 291.9803 41.5594
6432059 0.769740 0.000001 54964.754300 0.012775 288.0399 41.8184
6629588 2.264471 0.000003 54966.783103 0.042077 297.7556 42.0091
7289157 5.26581 0.00001 54969.976049 0.044130 293.9661 42.8373
7375612 0.1600729 0.0000001 54953.638704 0.008521 295.4670 42.9279
7668648 27.8186 0.0001 54963.315401 0.054044 286.2770 43.3391
7670617 24.7038 0.0001 54969.128128 0.044240 287.1845 43.3671
7955301 15.3244 0.0001 54968.272901 0.420237 290.1863 43.7239
8023317 16.5790 0.0001 54979.733478 0.052909 289.9703 43.8205
8122124 0.2492776 0.0000001 54964.612833 0.014220 299.0195 43.9576
8365739 2.38929 0.00000 54976.979013 0.028860 291.8875 44.3456
8758716 0.1072049 0.0000000 54953.672989 0.006195 293.8519 44.9494
8938628 6.86222 0.00002 54966.603088 0.018955 285.6626 45.2177
9214715 265.300 0.003 55149.923156 0.055997 290.3912 45.6816
9715925 6.30820 0.00003 54998.931972 0.016542 294.1644 46.4240
9834257 15.6514 0.0001 55003.014878 0.036964 293.2332 46.6002
9944907 0.613440 0.000000 54964.826351 0.021577 288.8836 46.8698
10014830 3.03053 0.00001 54967.124438 0.091901 293.1763 46.9224
10223616 29.1246 0.0001 54975.144652 0.093130 297.3907 47.2557
10268809 24.7090 0.0001 54971.999951 0.034276 289.6806 47.3178
10319590 21.3205 0.0001 54965.716743 0.085213 281.7148 47.4144
10743597 81.195 0.001 54936.094491 0.029846 296.3823 48.0453
10855535 0.1127824 0.0000000 54964.629315 0.006374 289.2119 48.2031
10919564 0.4621374 0.0000003 54861.987523 0.037827 291.5974 48.3933
11465813 670.70 0.01 55285.816598 0.302424 296.6986 49.3165
11558882 73.921 0.001 54987.661506 0.044135 291.7399 49.5708
11869052 20.5455 0.0001 54970.909481 0.022169 294.4019 50.1721
12062660 2.92930 0.00001 54998.936258 0.043481 292.0709 50.5468

Note. Those reported without ${\mathrm{BJD}}_{0}$ errors are also heartbeat stars.

Download table as:  ASCIITypeset image

8.8. Single Eclipse Events

Systems exhibiting a primary and/or secondary eclipse but lack a repeat of either one are shown in Table 9. For these 32 systems no ephemeris, ETV, or period error is determined. These are flagged with the "L" (long) flag and are available from the database under http://keplerEBs.villanova.edu/search but are not included in the EB Catalog.

Table 9.  The Systems With No Repeating Events (Long) in the Kepler Sample

KIC Period Event Width BJD0 R.A.(°) Decl.(°)
  (%) (days) (−2400000) (J2000) (J2000)
2162635 0.996 1.2 55008 291.9776 37.5326
3346436 0.84 0.9 55828 292.4968 38.4627
3625986 0.75 12 55234 284.7664 38.7935
4042088 0.983 0.4 55449 286.8895 39.1074
4073089 0.67 0.7 56222 295.4901 39.1059
4585946 0.89 0.8 54969 297.3779 39.6685
4755159 0.9 0.8 55104 294.1376 39.8821
5109854 0.87 0.5 55125 294.6938 40.292
5125633 0.92 0.4 55503 298.1539 40.2474
5456365 0.95 0.7 56031 294.1284 40.6823
5480825 0.992 1.5 55194 299.5369 40.6708
6751029 0.88 0.3 55139 281.3179 42.2645
6889430 0.87 2 55194 298.3608 42.3808
7200282 0.945 0.4 55632 291.7615 42.7566
7222362 0.6 0.6 55284 297.5835 42.7721
7282080 0.965 0.2 55539 291.8008 42.8907
7288354 0.965 0.2 55235 293.7429 42.862
7533340 0.98 0.8 55819 293.4575 43.1115
7732233 0.8 0.2 55306 282.2236 43.4657
7875441 0.93 1 55534 284.9639 43.6527
7944566 0.8 0.5 55549 285.1083 43.7168
7971363 0.965 1.5 55507 295.6284 43.7571
8056313 0.88 1.5 56053 299.735 43.8779
8648356 0.98 0.75 55357 299.1302 44.7447
9466335 0.93 35 55334 292.4365 46.0956
9702891 0.82 1.1 55062 288.4751 46.4646
9730194 0.9 1 55277 298.7442 46.4508
9970525 0.9988 0.3 54972 298.7442 46.8301
10058021 0.985 0.3 55433 283.7151 47.0321
10403228 0.955 2 55777 291.2267 47.55
10613792 0.975 0.4 55773 296.8537 47.8918
11038446 0.94 0.3 55322 295.7166 48.5419

Note. These systems do not have periods.

Download table as:  ASCIITypeset image

8.9. Additional Interesting Classes

The Catalog contains many interesting classes of objects which we encourage the community to explore. Some additional classes, which will only be listed (for brevity), include: total eclipsers, occultation binaries, reflection-effect binaries, red giants, pulsators, and short-period detached systems. These systems can be found using the search criteria provided at http://keplerEBs.villanova.edu/search.

9. SPECTROSCOPIC FOLLOW-UP

Modeling light curves of detached EB stars can only yield relative sizes in binary systems. To obtain the absolute scale, we need RV data. In case of double-lined spectroscopic binaries (SB2), we can obtain the mass ratio q, the projected semimajor axis $a\mathrm{sin}i$ and the center-of-mass (systemic) velocity vγ. In addition, SB2 RV data also constrain orbital elements, most notably eccentricity e and argument of periastron ω. Using SB2 light curve and RV data in conjunction, we can derive masses and radii of individual components. Using luminosities from light curve data, we can derive distances to these systems. When the luminosity of one star dominates the spectrum, only one RV curve can be extracted from spectra. These are single-lined spectroscopic binaries (SB1), and for those we cannot (generally) obtain a full solution. The light curves of semi-detached and overcontact systems can constrain the absolute scale by way of ellipsoidal variations: continuous flux variations due to the changing projected cross-section of the stars with respect to line of sight as a function of orbital phase. These depend on stellar deformation, which in turn depends on the absolute scale of the system. While possible, the photometric determination of the q and $a\mathrm{sin}i$ is significantly less precise than the SB2 modeling, and the parameter space is plagued by parameter correlations and solution degeneracy. Thus, obtaining spectroscopic data for as many EBs as possible remains a uniquely important task to obtain the absolute scale and the distances to these systems.

We successfully proposed for 30 nights at the Kitt Peak National Observatory's 4-m telescope to acquire high resolution (R ≥ 20,000), moderate signal-to-noise ratio (${\rm{S}}/{\rm{N}}\geqslant 15$) spectra using the echelle spectrograph. To maximize the scientific yield given the follow-up time requirement, we prioritized all targets into the following groups:

  • 1.  
    Well-detached EBs in near-circular orbits. These are the prime sources for calibrating the MLRT relationships across the main sequence. Because of the separation, the components are only marginally distorted and are assumed to have evolved independently from one another. Circular orbits indicate that the ages of these systems are of the order of circularization time (Zahn & Bouchet 1989; Meibom & Mathieu 2005). Modeling these systems with state-of-the-art modeling codes (i.e., WD: Wilson 1979, and its derivatives; ELC: Orosz & Hauschildt 2000; PHOEBE: Prša & Zwitter 2005) that account for a range of physical circumstances allows us to determine their parameters to a very high accuracy. These results can then be used to map the physical properties of these main sequence components to the spectral type determined from spectroscopy. While deconvolution artifacts might raise some concern (see Figure 3 left), in practice they become vanishingly small for orbital periods $\gtrsim 1$-day.
  • 2.  
    Low-mass main sequence EBs. By combining photometric observations during the eclipses and high-R spectroscopy, we can test the long standing discrepancy between the theoretical and observational mass–radius relations at the bottom of the main-sequence, namely that the observed radii of low-mass stars are up to 15% larger than predicted by structure models. It has been suggested that this discrepancy may be related to strong stellar magnetic fields, which are not properly accounted for in current theoretical models. All previously well-characterized low-mass main-sequence EBs have periods of a few days or less, and their components are therefore expected to be rotating rapidly as a result of tidal synchronization, thus generating strong magnetic fields. Stars in the binaries with longer orbital periods, which are expected to have weaker magnetic fields, may better match the assumptions of theoretical stellar models. Spectroscopy can provide evidence of stellar chromospheric activity, which is statistically related to ages, thus discriminating between young systems settling onto the main sequence and the older ones already on the main sequence.
  • 3.  
    EBs featuring total eclipses. A select few EBs with total eclipses allow us to determine the inclination and the radii to an even higher accuracy, typically a fraction of a percent. Coupled with RVs, we can obtain the absolute scale of the system and parameters of those systems with unprecedented accuracy.
  • 4.  
    EBs with intrinsic variations. Binarity is indiscriminate to spectral and luminosity types. Thus, components can be main-sequence stars, evolved, compact, or intrinsically variable—such as pulsators (δ-Sct, RR Lyr, ...), spots, etc. These components are of prime astrophysical interest to asteroseismology, since we can compare fundamental parameters derived from binarity to those derived from asteroseismic scaling relations (Huber et al. 2014).
  • 5.  
    EBs exhibiting ETVs. These variations can be either dynamical or due to the light time effect. The periodic changes of the orbital period are indicative of tertiary components (Conroy et al. 2014). Understanding the frequency of tertiaries in binary systems is crucial because many theories link the third component with the tightening of binary orbits via Kozai–Lidov mechanism and/or periastron interactions. Careful studies of statistical properties of ETVs detected in EBs may shed light on the origins of binarity. In addition, the uncertainty of fundamental parameters derived from multiple stellar systems is an order of magnitude smaller than that of binary stars (Carter et al. 2011; Doyle et al. 2011).

We acquired and reduced multi-epoch spectra for 611 systems within this program; typically 2–3 spectra per target were acquired near the quadratures. The spectra are available for download from the EB Catalog website. Extrapolating from these systems, we anticipate around 30% of the Kepler sample of EBs presented in the Catalog to be SB2 systems. The detailed analysis of the acquired spectra is the topic of an upcoming paper (C. Johnston et al. 2016, in preparation).

10. CATALOG ANALYSIS

In Papers I and II we noted a non-uniform distribution of EB occurrence rates as a function of galactic latitude. This was surprising at first, since the implication that EB stars are not uniformly distributed in space is not immediately obvious. Galaxy population observations and models, however, have long predicted this behavior: the stellar population at lower galactic latitudes (thin disk) contains notably younger stars that are on average larger (i.e., contain more giants in the magnitude-limited sample) than the older, sparser population of the thick disk and halo (Prša et al. 2015). In consequence, the geometric probability of eclipses increases toward the galactic disk. Figure 9 depicts this distribution for the complete sample of EBs in the current Catalog. Galactic latitudes are divided into ∼1° bins and EBs within those bins are counted. To get the area-corrected occurrence rates, their number is divided by the number of all stars observed by Kepler within the same bins. The latitude-dependent occurrence rate is immediately obvious. To the jth bin we assigned a corresponding error of $\sqrt{{N}_{j}}/{N}_{\mathrm{tot},j}$, where Nj is the number of detected EBs and ${N}_{\mathrm{tot},j}$ is the number of all stars in that bin. Unfortunately, Kepler's latitude span is not sufficient to fit an actual galactic population model to the number of EBs detected, so we resort to a toy proxy using a simple exponential fit: ${dp}/{db}(b)=A\mathrm{exp}(-C(b-B))+D$. We performed a least squares fit to derive parameters of the exponential, depicted by a solid line in Figure 9. This toy model predicts that the largest expected occurrence rate of EBs is $A+D\sim 2.2\%$, while the smallest expected occurrence rate is $D\sim 0.9\%$. Parameter B drives the galactic latitude offset ($b\sim 4^\circ $) where the exponential rise of EBs in this magnitude-limited sample is still a reasonable approximation, beyond which the disk opaqueness causes the number of EBs to level off and the exponential model to fail. Finally, parameter C provides an estimate of thin-to-thick disk transition, determined predominantly by the decrease in the number of giants in the magnitude-limited sample.

Figure 9.

Figure 9. EB occurrence rate (dp/db) as a function of galactic latitude (b). Each latitude bin is area-corrected to give true occurrence rates. Uncertainties are estimated as $\sqrt{{N}_{j}}/{N}_{\mathrm{tot},j}$, where Nj is the number of detected EBs and ${N}_{\mathrm{tot},j}$ is the total number of targets observed by Kepler in the jth bin. The solid line represents the exponential fit to the data, with parameters annotated in the figure. The parameters A and D are occurrence rates, C is the thin-to-thick disk transition value, and B is degrees of galactic latitude. The results of this toy model indicate that the occurrence rate span of EBs ranges between 0.9% and 2.2%.

Standard image High-resolution image

10.1. Catalog Completeness

With the primary Kepler mission having ended after 4 years of service, the longest orbital periods of EBs in the Catalog are ∼1000 days. Catalog completeness at those periods is challenged by our ability to detect every single eclipse event, which can be made difficult by the small eclipse amplitudes, data gaps and other intrinsic and extrinsic contributions to background and noise. Adding to this is the increasingly low probability of eclipses at the larger orbital separations that accompany longer periods, and of course the increasing probability of entirely missed eclipses for orbital periods longer than the observing window. On the other hand, completeness should be ∼100% for short period EBs ($P\sim 1$ day) because of the high geometrical probability of eclipses and because even non-eclipsing systems manifest as ellipsoidal variables. The overall Catalog completeness is thus predominantly a function of orbital period and S/N of the eclipses. Here we only estimate completeness and defer an in-depth study that derives the underlying orbital period distribution of all binaries from the Besançon model of the Galaxy to A. Prša et al. (2016, in preparation).

To estimate Catalog completeness, we start with the observed period distribution. Figure 10 depicts the distribution of orbital periods. Two features are particularly interesting: the excess of short period binaries ($P\sim 0.3$ day) and the gradual drop-off of longer period binaries.

Figure 10.

Figure 10. Distribution of orbital periods of all Cataloged EBs.

Standard image High-resolution image

The short-period excess is a well-known feature of EB stars: at short periods, proximity effects become pronounced, most notably ellipsoidal variations, which enable us to detect binary stars even in the absence of eclipses. This also drives the overall probability of detection sharply upwards at shorter periods. A much broader, less pronounced, longer period peak at ∼2–3 days is typically attributed to Kozai–Lidov Cycles and Tidal Friction (KCTF; Fabrycky & Tremaine 2007). The Kozai–Lidov mechanism describes interactions with a more distant third companion on an eccentric orbit, where angular momentum is exchanged between the inner binary eccentricity and the outer orbit inclination. When the components of the inner binary are sufficiently close, tidal friction dissipates energy and tightens the pair. The high occurrence rate of stellar triples (Gies et al. 2012; Rappaport et al. 2013; Conroy et al. 2014) gives further credibility to this model.

The gradual drop-off at the long period end is due to two main contributions. The dominant contribution is the geometrical probability of eclipses. Since the eccentricity distribution of Kepler EBs can be estimated and the pool of stellar masses and radii can be inferred from the bulk properties of all observed Kepler targets, computing this correction is a tractable problem. The second contribution is due to Kepler's duty cycle. The satellite observed a single patch of the sky, but observations were interrupted by regular quarterly rolls and data downlink, and by unexpected events that put the telescope into safe mode. As a result, the actual duty cycle of observations was ∼92%. The dead module and targets observed only for a subset of quarters further impact completeness. We discuss each of these effects in turn and derive corresponding corrections that we use to estimate Catalog completeness. The corrections are depicted in Figure 11.

Figure 11.

Figure 11. Corrections to EB detection rates as a function of orbital period. Geometrical correction is dominant across all orbital periods, while the duty cycle correction becomes important at the long period end. The solid curve for the duty cycle correction is computed by requiring at least two eclipse events to be observed, while the dashed curve shows the duty cycle correction for at least three observed eclipse events. The combined correction is the product of individual corrections.

Standard image High-resolution image

10.1.1. Geometrical Correction

The probability of eclipses is determined by a simple relation: $\mathrm{cos}i\leqslant ({\rho }_{1}+{\rho }_{2})/\delta $, where i is orbital inclination, ${\rho }_{1}$ and ${\rho }_{2}$ are fractional radii, and δ is the relative instantaneous separation. The right-hand side becomes progressively smaller with increasing orbital periods because of Kepler's 3rd law, whereas the $\mathrm{cos}i$ term is distributed uniformly, so the probability of eclipses drops. For eccentric orbits, the probability of eclipse at superior and inferior conjunction is ${p}_{\mathrm{sup},\mathrm{inf}}\quad =\quad ({\rho }_{1}+{\rho }_{2})(1\pm e\mathrm{sin}\omega )/(1-{e}^{2})$. To compute it, we take all ∼200,000 stars observed by Kepler and create a pool based on their effective temperatures and surface gravities as reported in the Kepler Input Catalog (KIC; Brown et al. 2011). From this pool we draw pairs of stars and use the mass–radius–temperature–log(g) relationship from Torres et al. (2010) to determine the masses and radii of the drawn stars. We then place these stars in orbits with a predefined orbital period in the $-1\leqslant \mathrm{log}P\leqslant 3$ range, and compute the semimajor axes. We also account for ellipsoidal variability on the short period end and non-zero eccentricity on the long period end. The geometrical correction is depicted in Figure 11 in black.

10.1.2. Duty Cycle Correction

While Kepler observed approximately 200,000 stars for most of the mission, which targets were observed in which quarters varies at the 10% level for various reasons, such as compensating for missing modules and variations in target lists over time. To account for these variations and the specific duty cycle, we performed a detailed analysis that leads to a global window function. As an approximation to the search of our algorithm, we use the Transiting Planet Search (TPS) deweighting vector (Burke et al. 2015) to determine which cadences are actually observed by Kepler. This correctly accounts for the exact gaps in the data.

For every target, we use the online stellar table to identify the quarters for which observations were taken. We then scan over periods ranging from 0.5 to 1500 days, with a trial period corresponding to integer number of Kepler cadences (30 minutes). For each period, we also scan over every possible phase (again, in integer cadences) and determine the fraction of phases where Kepler acquired data on this target. This creates an individual window function for each target that corresponds to the probability that an EB with a random phase would have had at least two observations as a function of period, which is the minimum requirement to determine the ephemeris of the system. We then sum these individual window functions to determine, for every period, how many Kepler targets would have two observations. The correction as a function of period is depicted in Figure 11 in solid blue.

It is instructive to compare this window function to the theoretical window function if we assume a uniform observing completeness of 92%. Assuming that an EB was observed for all quarters, the probability of observing any single eclipse is 92%, and the probability for an eclipse to fall into a data gap is 8%. The effect is smaller for intermediate period binaries and diminishes for shorter period systems. A binary with an orbital period of ∼1000 days will have 2 primary eclipse events in the data, a binary with an orbital period of ∼250 days will have 5 eclipse events, and a ∼100-day binary will have 14 events. Thus, the probability of a significant fraction of 14 events falling into data gaps is negligibly small, while the effect on 5 events, and especially 2 events, can be dire.

The probability that two or more eclipses were observed is calculated by a binomial expression that depends on the duty cycle and on the number of eclipses, which in turn depends on the orbital period. To compute that probability, it is simpler to compute the probability that no eclipses are detected or that only one eclipse is detected, and take the complement: ${p}_{\mathrm{detection}}=1-{p}_{0}-{p}_{1}$ $=\quad 1-{(0.08)}^{N}-N$ ${(0.92)}^{1}{(0.08)}^{N-1}$, where N is the number of eclipses. Kepler accumulated 1460 days of data, so N can be written as $1+\mathrm{int}(1460/P)$, where P is the orbital period. We expect a cascading correction with a discrete jump at every period that changes the integer division value. This correction is depicted in Figure 11 in dashed blue.

If we were to require a detection of three eclipses, we would need to subtract another binomial term from the probability equation, ${p}_{2}=0.5N{(N-1)(0.92)}^{N-2}{(0.08)}^{2}$, which further reduces the detection rate. This correction is depicted in Figure 11 in dash–dotted blue.

10.1.3. Combined Correction

The combined correction, ${\epsilon }_{{\rm{C}}}$, is a product of individual corrections: ${\epsilon }_{{\rm{C}}}={\epsilon }_{{\rm{G}}}{\epsilon }_{\mathrm{DC}}$. If we knew what the underlying distribution of all (not just eclipsing) binary stars is, then multiplying that distribution with ${\epsilon }_{{\rm{C}}}$ would provide us with a theoretical prediction for the observed number of EBs by Kepler. The actual number of EBs observed by Kepler, divided by the predicted number of EBs observed by Kepler, is then the measure of Catalog completeness.

We do not know the underlying distribution of binary stars, but we can approximate it for the purpose of this estimate. Several works, most notably Duquennoy & Mayor (1991) and Kroupa & Burkert (2001), report a log-Gaussian distribution with the peak well in excess of 10,000 days. We are thus sampling a far tail of the left wing, and we can assume that the distribution in the $1.3\leqslant \mathrm{log}P\leqslant 3$ range is locally38 linear, ${dB}/d(\mathrm{log}P)=a\mathrm{log}P\quad +\quad b$. By making that assumption, we can then: (1) set a and b to some reasonable initial values; (2) derive the EB distribution function $d\mathrm{EB}/d(\mathrm{log}P)={\epsilon }_{{\rm{C}}}\quad \times \quad {dB}/d(\mathrm{log}P);$ (3) compare this theoretical prediction with the observed distribution, and (4) iterate a and b to obtain the best-fit values on the linear $\mathrm{log}P$ range. Figure 12 depicts this comparison: the top plot shows the comparison between the observed EB period distribution (blue bars) and theoretical prediction in the $1.3\leqslant \mathrm{log}P\leqslant 3$ range (solid red line), and the bottom plot shows the comparison between the backwards-projected ${dB}/d(\mathrm{log}P)=d\mathrm{EB}/d(\mathrm{log}P)/{\epsilon }_{{\rm{C}}}$ (gray bars) and the assumed underlying linear distribution (red line). To get the completeness estimates, we divide the integral of the observed trend with the integral of the predicted trend. We do not take the short period excess into account because of its extrinsic causes. The annotations in the plots give the best-fit parameters and the derived completeness rates. The completeness of cataloged EBs is ${\eta }_{\mathrm{EB}}=89.1\%\pm 3.5\%$, and completeness of all binaries is ${\eta }_{\mathrm{bin}}=80\%\pm 11\%$.

Figure 12.

Figure 12. Completeness estimates for the EB Catalog (top), and projected completeness estimates for the binary star population (bottom). The observed distribution of orbital periods is depicted in blue, and the theoretical distribution of orbital periods, derived from a linear model of the underlying binary log-period distribution, is depicted in red. The distribution depicted in gray is the predicted occurrence rate of all binary stars based on Kepler data, compared to the best-fit linear model.

Standard image High-resolution image

11. SUMMARY

This revision of the Catalog contains new additions consisting of rejected KOIs, previously misidentified false positives, proprietary systems, updated long period systems, EBs identified from other systems, heartbeat stars, and Planet Hunters' systems. Detrended and phased long-cadence data are hosted and available for the public, as well as an updated period for each system adjusted to account for ETVs. We also provide a deconvolved polyfit which serves as a better approximation of the actual light curve and a classification parameter representing the morphology of the phased light curve. A Catalog analysis including a completeness study is provided.

An online version of the Catalog is maintained at http://keplerEBs.villanova.edu. This Catalog lists the KIC, ephemeris, morphology, principle parameters, polyfit data, ETV data, raw data, and an array of figures displaying the raw time domain, detrended data, and phased light curves of each system along with period frequency, ETV, and diagnostic analysis plots. The online Catalog also provides a visualization tool to further exploit this data set. It is our hope that the Catalog will serve the EB community as a bridge between the raw Kepler data and in-depth scientific modeling.

All of the data presented in this paper were obtained from the Multimission Archive at the Space Telescope Science Institute (MAST). STScI is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. Support for MAST for non-Hubble Space Telescope data is provided by the NASA Office of Space Science via grant NNX09AF08G and by other grants and contracts. Funding for this Discovery Mission is provided by NASAs Science Mission Directorate. Spectroscopic follow-up data are made available through NOAO survey program 11A-0022. This work is funded in part by the NASA/SETI subcontract 08-SC-1041 and NSF RUI AST-05-07542. B.Q. was supported by an appointment to the NASA Postdoctoral Program at the Ames Research Center, administered by Oak Ridge Associated Universities through a contract with NASA. T.S.B. acknowledges support from ADAP14-0245 and ADAP12-0172. A.D. has been supported by the Postdoctoral Fellowship Programme of the Hungarian Academy of Sciences, the János Bolyai Research Scholarship of the Hungarian Academy of Sciences, Lendület-2009 Young Researchers Programme of the Hungarian Academy of Sciences, the European Communitys Seventh Framework Programme (FP7/2007-2013) under grant agreement no. 269194 (IRSES/ASK) and no. 312844 (SPACEINN). A. D. has also been supported by the Hungarian National Research, Development and Innovation Office—NKFIH K-1157709.

Facility: Kepler - The Kepler Mission.

APPENDIX: THE ONLINE CATALOG

The online Catalog provides a searchable database of all the EBs found in the entire Kepler data set along with downloadable content and visualization tools. The online version of the Catalog is currently maintained at http://keplerEBs.villanova.edu. The homepage presents the user with a view of the entire EB catalog. Along the top tool-bar is the tab, "Search" which provides a page where the Catalog can be filtered, sorted, and exported with a variety of options: ID numbers, eclipse properties, morphology parameter, location elements, ETV properties, effective temperatures, and flags with the ability to export the results to something other than the default HTML table, if desirable. For a complete list of search options, units and explanations, and exporting formats please see the online Help section.

For individual EB entries, the online Catalog provides a summary of the EB's physical properties, analytic plots, and provides the time-series data for download in various formats. In addition to the raw data, the polyfit data used to detrend the time-series data and the ETV data, along with the suite of diagnostic figures are available for download. The figures provided for each EB entry are: raw data, detrended data, polyfit, ETV, and TPF plots. If available, the individual entry page shows additional spectral observations available for download and any publications concerning that EB entry.

The Catalog currently maintained at http://keplerEBs.villanova.edu will continue to evolve as additional data is ingested and new techniques enhance our research capabilities. A static version of the online Catalog associated with this paper is maintained at MAST https://archive.stsci.edu/kepler/eclipsing_binaries.html.

Footnotes

  • 37 
  • 38 

    Tidal interaction and the Kozai–Lidov mechanism prohibit this toy model from working below $\mathrm{log}P\sim 1.3$, which is why we do not attempt to model the short period end; since it is reasonable to assume that the Catalog is essentially complete on that end, this deviation from the model bears no significant impact on our completeness result.

Please wait… references are loading.
10.3847/0004-6256/151/3/68