SlideShare ist ein Scribd-Unternehmen logo
1 von 49
Downloaden Sie, um offline zu lesen
Preliminary draft. To be submitted to ApJ.


                                              Architecture and Dynamics of Kepler’s Candidate Multiple Transiting Planet
                                                                              Systems

                                               Jack J. Lissauer1 , Darin Ragozzine2 , Daniel C. Fabrycky3,4 , Natalie M. Batalha5 , William J.
arXiv:1102.0543v1 [astro-ph.EP] 2 Feb 2011




                                               Borucki1 , Stephen T. Bryson1 , Douglas A. Caldwell6 , Edward W. Dunham13 , Eric B. Ford7 ,
                                              Jonathan J. Fortney3 , Thomas N. Gautier III8 , Matthew J. Holman2 , Jon M. Jenkins1,6 , David
                                              G. Koch1 , David W. Latham3 , Geoffrey W. Marcy9 , Robert Morehead7 , Jason Rowe6 , Elisa V.
                                                             Quintana6 , Dimitar Sasselov2 , Avi Shporer10,11 , Jason H. Steffen12

                                                                                         Jack.Lissauer@nasa.gov


                                                                                               ABSTRACT


                                                          Borucki et al. (2011a) [ApJ submitted, hereafter B11] report on characteristics of
                                                      over 1200 candidate transiting planets orbiting nearly 1000 Kepler spacecraft target
                                                      stars detected in the first four months of spacecraft data. Included among these tar-
                                                      gets are 115 targets with two transiting planet candidates, 45 targets with three, 8
                                                      with four, and one each with five and six sets of transit signatures. We character-
                                                      ize herein the dynamical properties of these candidate multi-planet systems. We find
                                                      that virtually all systems are stable, as tested by numerical integration assuming a
                                                      mass-radius relationship. The distribution of observed period ratios is also clustered

                                                1
                                                    NASA Ames Research Center, Moffett Field, CA, 94035, USA
                                                2
                                                    Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
                                                3
                                                    Department of Astronomy & Astrophysics, University of California, Santa Cruz, CA 95064, USA
                                                4
                                                    Hubble Fellow
                                                5
                                                    Department of Physics and Astronomy, San Jose State University, San Jose, CA 95192, USA
                                                6
                                                    SETI Institute/NASA Ames Research Center, Moffett Field, CA, 94035, USA
                                                7
                                                    University of Florida, 211 Bryant Space Science Center, Gainesville, FL 32611, USA
                                                8
                                                    Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA
                                                9
                                                    Astronomy Department, UC Berkeley, Berkeley, CA 94720
                                               10
                                                Las Cumbres Observatory Global Telescope Network, 6740 Cortona Drive, Suite 102, Santa Barbara, CA 93117,
                                             USA
                                               11
                                                    Department of Physics, Broida Hall, University of California, Santa Barbara, CA 93106, USA
                                               12
                                                    Fermilab Center for Particle Astrophysics, P.O. Box 500, MS 127, Batavia, IL 60510
                                               13
                                                    Lowell Observatory, 1400 W. Mars Hill Road, Flagstaff, AZ 86001, USA
–2–


     just outside resonances, particularly the 2:1 resonance. Neither of these characteristics
     would emerge if the systems were significantly contaminated with false positives, and
     these combined with other considerations strongly suggest that the majority of these
     multi-candidate systems are true planetary systems. Using the observed multiplicity
     frequencies (e.g., the ratio of three-candidate systems to two-candidate systems), we
     find that a single population that matches the higher multiplicities underpredicts the
     number of singly-transiting systems. We provide constraints on the true multiplicity
     and mutual inclination distribution of the multi-candidate systems, which includes a
     population of systems with multiple small planets with low ( 5◦ ) inclinations. In all,
     multi-transiting systems from Kepler provide a large and rich dataset which can be
     used to powerfully test theoretical predictions of the formation and evolution of plan-
     etary systems. [This is a preliminary draft; some numbers may change slightly in the
     submitted version.]

     Subject headings: celestial mechanics – planetary systems



                                        1.    Introduction

     Kepler is a 0.95 m aperture space telescope using transit photometry to determine the fre-
quency and characteristics of planets and planetary systems (Borucki et al. 2010; Koch et al. 2010;
Jenkins et al. 2010; Caldwell et al. 2010). Besides searching for small planets in the habitable zone,
Kepler ’s ultra-precise long-duration photometry is also ideal for detecting systems with multiple
transiting planets. Borucki et al. (2011b) presented, for the first time, 5 Kepler systems with
multiple transiting planet candidates, and these candidate systems were analyzed by Steffen et al.
(2010). A system of three planets (Kepler-9=KOI-377, Holman et al. 2010; Torres et al. 2011) was
found, in which a resonant effect on the transit times allowed Kepler data to confirm two of the
planets and characterize the dynamics of the system. A compact system of six transiting planets
(Kepler-11, Lissauer et al. 2011), five of which were confirmed by their mutual gravitational inter-
actions exhibited through transit timing variations (TTVs), has also been reported. These systems
were highly anticipated because of their unique value in understanding the formation and evolution
of planetary systems Koch et al. (1996); Agol et al. (2005); Holman & Murray (2005); Fabrycky
(2009); Holman (2010); Ragozzine & Holman (2010).

     A catalog of all candidate planets evident in the first four and a half months of Kepler data
is presented in B11. They also summarize the observational data and follow-up programs being
conducted to confirm and characterize planetary candidates. Each object in this catalog is assigned
a vetting flag (reliability score), with 1 signifying a confirmed or validated planet (>98% confidence),
2 being a strong candidate (> 80% confidence), 3 being a somewhat weaker candidate (> 60%
confidence), and 4 signifying a candidate that has not received ground-based follow-up (this sample
also has an expected reliability of >60%). For most of the remainder of the paper, we refer to all
–3–


of these objects as “planets”, although 99% remain unvalidated and thus should more properly be
termed planet candidates. Ragozzine & Holman (2010) and Lissauer et al. (2011) point out that
systems with multiple planet candidates are less likely to be due to astrophysical false positives
than stars with single candidates, although the possibility that one of the candidate signals is due
to, e.g., a background eclipsing binary, is still non-negligible. Here we only examine the candidates
that are known to be in multiple systems; many of the apparently single systems are likely to have
additional planets that escape detection by being too small, too long-period, or too highly inclined
with respect to the line of sight. Ford et al. (2011, in preparation) discuss candidates exhibiting
TTVs that may be caused by non-transiting planets. B11 list five dozen planetary candidates with
radii Rp > 15 R⊕ (Earth radii), but none of these are in multiple candidate systems. Latham
et al. (2011, in preparation) compare the radius distribution of planets in multi-planet systems
with those of planets which don’t have observed transiting companions. We examine herein the
dynamical aspects of Kepler’s multi-planet systems.

     Ragozzine & Holman (2010) discussed in detail the value of systems with multiple transiting
planets. Besides their higher likelihood of being true planetary systems, multi-transiting systems
provide numerous other insights that are difficult or impossible to gain from single-transiting sys-
tems. These systems harness the power of radius measurements from transit photometry in combi-
nation with the illuminating properties of multi-planet orbital architecture. Observable interactions
between planets in these systems, seen in transit timing variations (TTVs), are much easier to char-
acterize, and our knowledge of both stellar and planetary parameters is improved. The true mutual
inclinations between planets in these systems, while not directly observable from the light curves
themselves, are much easier to obtain than in other systems, both individually and statistically (see
Section 6 and Lissauer et al. 2011). The distributions of planets in these systems, both in terms
of physical and orbital parameters, are invaluable for comparative planetology and provide unique
and invaluable insights into the processes of planet formation and evolution.

     It is thus very exciting that B11 report 115 doubles, 45 triples, 8 quads, 1 quintuple and 1
sextuple transiting system. While very few of these systems contain confirmed planets, it is very
likely that this set of multi-transiting candidates contains more than twice as many as planets as
those known in the pre-Kepler era.

     We begin by summarizing the characteristics of Kepler’s multi-planet systems in Section 2. The
reliability of the data set is discussed in Section 3. Sections 4 – 6 present our primary statistical
results: Section 4 presents an analysis of the long-term dynamical stability of candidate planetary
systems for nominal estimates of planetary masses and unobserved orbital parameters. Evidence of
orbital commensurabilities in among planet candidates is analyzed in Section 5. The characteristic
distributions of numbers of planets per system and mutual inclinations of planetary orbits are
discussed in Section 6. We compare the Kepler-11 planetary system (Lissauer et al. 2011) to other
Kepler candidates in Section 7. We conclude with a discussion for the implications of these data
and summarize the most important results presented herein.
–4–


                    2.   Characteristics of Kepler Multi-Planet Systems

     The light curves derived from Kepler data can be used to measure the radii and orbital periods
of planets. Periods are measured to high accuracy. The ratios of planetary radii to stellar radii
are also measured quite precisely, except for low signal to noise transits (e.g., planets much smaller
than their stars, those for which few transits are observed, and planets orbiting faint or variable
stars). Besides errors in estimated stellar radii, in some cases the planet’s radius is underestimated
because a significant fraction of the flux in the target aperture is provided by a star other than the
one being transited by the planet. This extra light can often be determined and corrected for by
analyzing displacements of the centroid of light from the target between in-transit and out-of-transit
observations (Bryson et al. 2010), high-resolution imaging and/or spectroscopic studies conducted
from the ground , although dilution by faint physically bound stars is difficult to detect (Torres et al.
2011). Spectroscopic studies can and in the future will be used to estimate the stellar radii, thereby
allowing for improved estimates of the planetary radii, and also the stellar masses.

     We list the key observed properties of Kepler two-planet systems in Table 1, three-planet
systems in Table 2, four-planet systems in Table 3, and those systems with more than four planets
in Table 4. Planet indexes given in these tables signify orbit order, with 1 being the planet with the
shortest period. These do not always correspond to the post-decimal point portion (.01, .02, etc.)
of the KOI number designation of these candidates given in B11, for which the numbers signify the
order in which the candidates were identified. These tables are laid out differently from one another
because of the difference in parameters that are important for systems with differing numbers of
planets. In addition to directly observed properties, these tables contain results of the dynamical
analyses discussed below.

     We plot the period and radius of each active Kepler planetary candidate observed during
the first four months of spacecraft operations, with special emphasis on multi-planet systems,
in Figure 1. The ratio(s) of orbital periods is a key factor in planetary system dynamics; the
distribution of these period ratios is plotted in Figure 3. Figure 4 shows a cumulative distribution
of the period ratio.

      To check whether the observed transiting planets are randomly distributed among host stars
we compared the observed distribution of planets per system to a statistically random Poisson
distribution, assuming there are 1229 planets orbiting some of the 150,000 stars Kepler monitors,
i.e., a Poisson distribution with a mean of 1229/150,000≈ 0.008. Such a Poisson distribution gives
an expectation of 1219 single planet systems, 5.0 doubles, 0.014 triples, 2.8 × 10−5 systems with
four planets and above. For single planet systems there are fewer observed systems than expected
from a random distribution, but, for systems with two and more planets there are far more systems
detected by Kepler than predicted randomly. This shows that transiting planets tend to “come in
packs” or cluster around stars, meaning that once a planet is detected around a given star, that star
is more likely to be orbited by an additional planet than a star with no detected planets. Some of
the observed clustering is attributable to SNR considerations, with stellar brightness, size, intrinsic
–5–


variability, multiplicity, crowdedness of field, and position on the Kepler detectors all contributing.
Quantifying the effects of SNR variations is beyond the scope of this paper, but they are clearly far
too small to account for the bulk of the observed clustering. Given the bias of the transit method
to detect planets on edge-on orbits, a possible interpretation of the clustering of transiting planets
is that planetary systems (or at least some fraction of them) tend to be coplanar, like the Solar
System.

     It is difficult to estimate the corresponding distribution for planets detected through radial
velocity (RV) surveys, since we do not know how many stars were spectroscopically surveyed for
planets for which none were found. However, a search through the Exoplanet Orbit Database1
(Wright et al. 2010) shows that 17% of the planetary systems detected by RV surveys are multi-
planet systems (have at least two planets), compared to 21% for Kepler. Including a linear RV
trend as evidence for an additional planetary companion, Wright et al. (2009) show that at least
28% of known RV planetary systems are multiple. Therefore, the clustering observed for transiting
planets detected by Kepler is likely to occur also for RV planets, and maybe even at a higher
fraction, which is expected given the bias of the transit detection method.



                                           3.   Reliability of the Sample

     Very few of the planetary candidates presented herein have been validated as true exoplan-
ets. As discussed in B11, the overwhelming majority of false positives are expected to come from
eclipsing binary stars. Some of these will be grazing eclipses of the target star, others eclipses of
fainter stars (physically associated to the target or background/faint foreground objects) within the
target aperture. Similarly, transits of fainter stars by large planets may also mimic small planets
transiting the target star. The vast majority of all active Kepler planet candidates were identified
using automated programs that do not discriminate between single candidates and systems of mul-
tiple planets. The 1% of single and multiple candidate systems that were identified in subjective
visual searches are too small to affect our results in any statistically significant manner, and even
the signs of such changes are unclear.

     The fact that planetary candidates are clustered around targets in the sense that there are far
more targets with more than one candidate than would be expected for randomly distributed can-
didates (section 2) suggests that the reliability of the multi-candidate systems is likely to be higher
than that of the single candidate targets. We know from radial velocity observations that planets
frequently come in multiple systems (Wright et al. 2009), whereas astrophysical false positives are
nearly random.

    Nonetheless, the presence of multiple candidates increases our confidence that one or both are
planets is real since the probability of two false positive signals is the product of the probabilities

  1
      exoplanet.org, as of January 2011.
–6–


of two relatively rare cases, significantly lowering the overall probability than none of the signals
is due to planets Ragozzine & Holman (2010). (Triple star systems would be dynamically unstable
for most of the orbital period ratios observed. Their dynamics would also give rise to very large
eclipse timing variations — e.g., Carter et al. 2011 — which are generally not seen.) Additionally,
the concentration of candidate pairings with period ratio near 1.5 and 2 (fig. 3) suggest that these
subsamples likely have an even larger fraction of true planets, since such concentrations would not
be seen for random eclipsing binaries. However, at present these qualitative factors have not been
quantified.

     Strong TTV signals are no more common for planets in observed multi-transiting candidates
than for single candidates (Ford et al. 2011, in preparation), suggesting that large TTVs are
often the result of stellar systems masquerading as planetary systems (Fabrycky et al. 2011, in
preparation, Steffen et al., 2011, in preparation). Nonetheless, it is possible that weaker TTVs
correlate with multi-transiting systems, but these require more data and very careful study to
reveal (e.g., those in the Kepler-11 planetary system, Lissauer et al. 2011).

     The ratio of durations given the observed periods of candidates in multi-transiting systems
can be used to potentially identify false positives (generated by a blend of two objects eclipsing
two stars of different densities) or as a probe of the eccentricity and inclination distributions. An
analysis of the corrected duration ratio as in (?) shows that essentially all pairs of candidate have
the expected duration ratio if they were both orbiting the same star. The consistency of duration
ratios is not a strong constraint, but the observed systems also pass this test.



                       4.   Long-term Stability of Planetary Systems

    Kepler measures planetary sizes, whereas masses are the key parameter for dynamical studies.
We convert sizes to masses using the following simple formula:
                                                  2.06
                                            Mp = Rp                                               (1)

where the planet’s mass, Mp , is expressed in Earth masses and its radius, Rp , in units of Earth’s
mean radius. The power-law of equation 1 is fit to Earth and Saturn; it slightly overestimates the
mass of Uranus (17.2 M⊕ vs. 14.5 M⊕ ) and slightly underestimates the mass of Neptune (16.2 M⊕
vs. 17.1 M⊕ ). (Note that Uranus is larger, but Neptune is more massive, so fitting both with a
monotonic mass-radius relation is not possible.)

    A convenient metric for the dynamical proximity of planets is the separation of their orbital
semi-major axes, ai+1 − ai , measured in units of their mutual Hill sphere radius
                                          mi + mi+1   1/3 (ai   + ai+1 )
                              RHi,i+1 =                                  ,                        (2)
                                             3M⋆                 2
where the index i = 1, N −1, with N being the number of planets in the system under consideration,
mi and ai are the planetary masses and semi-major axes, and M⋆ is the central star’s mass.
–7–


    The dynamics of two-planet systems are a special case of the 3-body problem that is amenable
to analytic treatment and simple numerically-derived scaling formulas. For instance, a pair of
planets initially on coplanar circular orbits with orbital separation
                                                          √
                                  ∆ ≡ (ao − ai )/RH > 2 3 ≈ 3.46                             (3)

cannot develop crossing orbits later, and are thus called “Hill stable” (Gladman 1993). Orbital
separations of all planet pairs in two planet systems for stellar masses given in B11 and planetary
masses given by Equation 1 are listed in Table 1. We see that they all obey this criterion.

     The actual dynamical stability of these systems also depends on their eccentricity and mutual
inclination (Veras & Armitage 2004), neither of which can be measured well from the transit data
alone. Circular coplanar orbits have the lowest angular momentum deficit (AMD), and thus are
the most stable, except for protective resonances that produce libration of angles and prevent close
approaches; a second exception exists for retrograde orbits (Smith & Lissauer 2009), but these are
implausible on cosmogonic grounds (Lissauer & Slartibartfast 2008). Hence, the Kepler data can
only be used to suggest that a pair of planets may be potentially unstable, particularly if their
semi-major axes are closer than the resonance overlap limit (Wisdom 1980; Malhotra 1998):
                                       aouter − ainner       Mp   2/7
                                   2                   < 1.5            .                          (4)
                                       ainner + aouter       M⋆
(This is derived only for one massive planet; the other is massless.) Conversely, if it is assumed
that both candidates are true planets, limits on the eccentricities and mutual inclinations can be
inferred from the requirement that the planetary systems are presumably stable for the age of the
system, which is generally of order 1011 times longer than the orbital timescale for the short-period
planets that represent the bulk of Kepler ’s candidates. However, we may turn the observational
problem around: the ensemble of duration measurements can be used to address the eccentricity
distribution (Ford et al. 2008); since multiplanet systems have eccentricities constrained by stability
requirements, they can be used as a check to those results (Moorhead et al. 2011, in preparation).

     Systems with more than two planets have additional dynamically complexity, but are very
unlikely to be stable unless all neighboring pairs of planets satisfy equation 3. Nominal separations
between neighboring planet pairs in three planet systems are listed in Table 2, the separations in four
planet systems are listed in Table 3, and those for the 5 and 6 planet systems in Table 4. We find that
in these cases, equation 3 is satisfied by quite a wide factor. From suites of numerical integrations
Smith & Lissauer (2009), dynamical survival of systems with more than three comparably-spaced
planets for at least 1010 orbits has been shown only if the relative spacing between orbital semi-
major axes exceeds a critical number (∆crit         9) of mutual Hill-spheres. For each adjacent set
of 3 planets in our sample, we plot the ∆ separations of both the inner and the outer pairs in
Figure 9. We find that in some cases ∆ < ∆crit , but the other separation is quite a bit bigger than
∆crit . In our simulations of a population of systems in section 6, we impose the stability boundary
∆max + ∆min > 18, which accounts for this possibility, in addition to requiring that inequality 3 is
satisfied.
–8–


     We investigated long-term stability of all 55 systems with three or more planets using the hybrid
integrator within the Mercury package (Chambers 1999), run on the supercomputer Pleiades at
University of California, Santa Cruz. We set the switchover at 3 Hill radii, but in practice we
aborted simulations that violated this limit, so for the bulk of the simulation the Wisdom-Holman
n-body mapping (Wisdom & Holman 1991) was used, with a time step of 0.05 times as large as the
orbital period of the innermost planet. The simplest implementation (Nobili & Roxburgh 1986) of
general relativistic precession was used, an additional potential UGR = −3(GM⋆ /cr)2 , where G is
Newton’s constant, c is the speed of light, and r is the instantaneous distance from the star. More
sophisticated treatments of general relativity (Saha & Tremaine 1994) are not yet required, due to
the uncertainties of the masses of the planets and stars whose dynamics are being modeled. We
neglected precession due to tides or rotational flattening, which are only significant compared to
general relativity for Jupiter-size planets in few-day orbits (Ragozzine & Wolf 2009). The stellar
mass was derived from the R⋆ and log g estimates from color photometry, as given by B11. We
assumed initially circular and coplanar orbits that matched the observed periods and phases, and
chose the nominal masses of equation 1.

     For the triples and quadruples, we ran these integrations for 4 × 109 orbital periods of the
innermost planet, and found the nominal system to be stable for this span in nearly all cases; the
three exceptions are described below. An integration of the five-planet system KOI-500 has run
2 × 109 days (the inner planet’s period is 0.99 day), suggesting stability for astronomically rele-
vant timescales is possible. However, the short orbital periods and high planet number means this
conclusion needs to be revisited in future work, especially once masses are estimated observation-
ally. (We consider the near-resonances of KOI-500 in section 5.5.) The most populous transiting
planetary system discovered so far is the six-planet system Kepler-11. In the discovery paper,
Lissauer et al. (2011) reported a circular model, with transit time variations yielding an estimate
of the masses of the five inner planets. They also proposed two more models with planets on mod-
erate eccentricities, and slightly different masses, which also fit the transit timing signals. Now,
both those eccentric solutions have become unstable, at 610 Myr for the all-eccentric fit, and at
427 Myr for an integration that was a very slight offset from the b/c-eccentric fit. The all-circular
fit remains stable at 640 Myr, so we conclude that the fits of that paper are physically plausible,
but that introducing a small amount of eccentricity can compromise stability. Future work should
address how high the eccentricities could be, while requiring stability for several Gyr.

     Only three systems became unstable at the nominal masses, KOI-284, KOI-191, and KOI-730;
the first is described next, the latter two are described in section 5. KOI-284 has a pair of candidates
with a period ratio 1.0383. Both planets would need to have masses about that of the Earth or
smaller for the system to satisfy equation 3 and thus be Hill stable. But the vetting flags are 3
for both these candidates — meaning that they are not good candidates anyway — and the data
display significant correlated noise which may be responsible for these detections. We expect one
or both of these candidates are not planets, or if they are, they are not orbiting the same star.

    In our investigation, we also were able to correct a poorly fit radius of 3.2 RJup (accompanied
–9–


by a grazing impact parameter in the original model, trying to account for a “V” shape) for KOI-
1426, which resulted in too large a nominal mass, driving the system unstable. It is remarkable
that stability considerations allowed us to identify a poorly conditioned light curve fit. Once the
correction was made to 1.2 RJup , the nominal mass allowed the system to be stable.

     That so many of the systems passed basic stability measurements suggests that there are not
a large number of planets that with substantially higher densities than given by the simple formula
of equation (1). If these candidates are assumed to be orbiting the same stars, stability constraints
could potentially be used to place upper limits on the masses and densities.



                                         5.   Resonances

                            5.1.   Resonance Abundance Analysis

     The ratio of periods of multi-transiting candidates in the same system is significantly different
from ratios obtained by the ratio of randomly selected periods from all 408 multi-transiting can-
didates. Multiply-transiting planets provide better data for the measurement of resonances than
RV planetary systems for two reasons. First, periods are measured to very high accuracy, even
for systems with low SNR, so period ratios can be confidently determined without the harmonic
ambiguities that affect RV detections (Anglada-Escud´ et al. 2010; Dawson & Fabrycky 2010). Sec-
                                                      e
ond, resonances significantly enhance the signal of transit timing variations (Agol et al. 2005;
Holman & Murray 2005; Veras & Ford 2010), which can also be measured accurately, either to
demonstrate resonance occupation (as will be possible with the Kepler-9 system with future data;
Holman et al. 2010 or exclude it with confidence.

     Planets can be librating in a resonance even when they have apparent periods that are not
perfectly commensurate (Ragozzine & Holman 2010). Nevertheless, the period ratios of small tran-
siting resonant planets will generally be within a few percent of commensurability, unless masses or
eccentricities are relatively high. In the cumulative distribution of period ratios (Figure 3), we see
that few planets appear to be directly in the resonance (with the important exception of KOI-730).

     In the solar system, there is a known excess of satellite pairs near mean-motion resonance
(MMR) (Goldreich 1965) and theoretical models of planet formation and migration suggest that
this should be the case (e.g., Terquem & Papaloizou 2007). We can ask whether such an excess can
be seen in the sample of multiply transiting Kepler planetary systems. To do this, we divide the
period ratios of the multiply transiting systems into bins which surround the first and second-order
(j:j − 1 and j:j − 2) MMRs. The boundaries between these “neighborhoods” are chosen at the
intermediate, third-order MMRs. Thus, all candidate systems with period ratios less than 4:1 are
in some neighborhood. For example, the neighborhood of the 3:1 MMR runs between period ratios
of 5:2 and 4:1, for the 2:1 it runs from 7:4 to 5:2. With this method the first-order MMRs have
neighborhoods that are essentially twice as large as those for second-order MMRs.
– 10 –


    We define a statistic, ξ, that is a measure of the difference between an observed period ratio
and the MMR in its neighborhood. In order to treat each neighborhood uniformly, they are scaled
such that the ξ statistic runs from −1 to 1. For first-order MMRs, ξ is given by

                                           1                1
                                ξ1 ≡ 3        − Round                                             (5)
                                         P −1              P−1

where P is the observed period ratio (always greater than unity) and “Round” is the standard
rounding function that returns the integer nearest to its argument. Similarly, for second-order
MMRs, ξ is given by
                                        2                 2
                             ξ2 ≡ 3         − Round             .                           (6)
                                      P −1              P−1
Table 5 gives the number of period ratios taken from tables 1–4 that are found in each neighborhood.
The ξ statistic as a function of the period ratio, along with the data, is shown in Figure 6.

     We calculate ξ for each planetary period ratio and stack the results for all of the resonances
in each MMR order. The resulting distribution in ξ is compared to a randomly-generated sample
drawn from a uniform distribution in log P , a uniform distribution in P , and from a sample con-
structed by taking the ratios of a random set of the observed periods in the multiple candidate
systems. This third sample has a distribution that is very similar to the logarithmic distribution.
Figure 7 shows a histogram of the number of systems as a function of ξ for both the first and
second-order MMRs. We see that the results for the second-order MMRs have some qualitative
similarities to the first-order MMR, though the peaks and valleys are less prominent. In our tests
below we also consider the case where all first and second-order MMRs are stacked together.

     To test whether the observed distributions in ξ are consistent with those from our various
test samples, we took the absolute value of ξ for each collection of neighborhoods and used the
Kolmogorov-Smirnov test (KS test) to determine the probability that the observed sample is drawn
from the same distribution as our test sample. The results of these tests are shown in Table 6.
Figure 8 shows the probability density function (PDF) for the combined distribution of first and
second-order MMRs and the PDF for the logarithmically distributed sample.

     A few notable results of these tests include: 1) the combined first and second-order neighbor-
hoods are distinctly inconsistent with any of the trial period distributions, 2) the neighborhoods
surrounding the first-order MMRs are also very unlikely to originate from the test distributions, 3)
there is significant evidence that the neighborhood surrounding the 2:1 MMR alone is inconsistent
with the test distributions, and 4) there is a hint that second-order resonances are not be consistent
with the test distributions, particularly the distribution that is uniform in P . Additional data are
necessary to produce higher significance results for the second-order resonances.

     Given the distinct differences between our test distributions and the observed distribution,
a careful look at the histograms shown in Figure 7 shows that the most common location for a
pair of planets to reside is slightly exterior to the MMR (the planets are farther apart than the
resonance location) regardless of whether the resonance is first or second-order. Also, a slight
– 11 –


majority (just over 60%) of the period ratios have ξ statistics between -1/2 and +1/2 while all
of the test distributions have roughly 50% between these two values in ξ (note that ξ = ±1/2
corresponds to sixth-order resonances in first-order neighborhoods and twelfth-order resonances in
second-order neighborhoods—none of which are likely to be strong in these planet pairs though
they may play some minor role). There are very few examples of systems that lie slightly closer to
each other than the first-order resonances and for the second-order resonances only the 3:1 MMR
has systems just interior to it. Finally, there is a hint that the planets near the 2:1 MMR have a
range of orbital periods, while those near the 3:2 appear to have shorter periods and smaller radii
on average. While additional data are necessary to claim these statements with high confidence, the
observations are nonetheless interesting and merit some investigation to determine the mechanisms
that might produce such distributions.

     A preference for resonant period ratios cannot be exhibited by systems without direct dynam-
ical interaction. Thus, the preference for periods near resonances in our dataset is statistically
significant evidence that many of the systems are actually systems of three or more planets. (Stars
with such close orbits are not dynamically stable, thus dynamical interaction between objects with
period ratios 5 indicates low masses due to planets.) Assessing the probability of dynamical
proximity to resonance for the purpose of eliminating the false positive hypothesis in any partic-
ular system requires a specific investigation; here, we can only say that statistically many of the
near-resonant systems appear to be planetary in nature.

    The excess of pairs of planets wider than nominal resonances was predicted by (Terquem & Papaloizou
2007), who pointed out that tidal interactions with the star will, under some circumstances, break
resonances and will more strongly affect the inner planets. Interestingly KOI-730 seems to have
maintained stability and resonance occupation despite differential tidal evolution.



                            5.2.   Frequency of Resonant Systems

     Doppler planet searches suggest that roughly one third of multiple planet systems well-characterized
by Doppler observations are near a low-order period commensurability with one sixth near the 2:1
MMR (Wright et al. 2011). However, an approximate degeneracy with a single planet on an eccen-
tric orbit makes it difficult for Doppler observations to detect a low-mass planet in an interior 2:1
MMR (Giuppone et al. 2009; Anglada-Escud´ et al. 2010). Among Doppler-discovered systems,
                                              e
only one quarter of the pairs near the 2:1 MMR have an inner planet significantly less massive
than the outer planet. The two exceptions (GJ 876 b&c, µ Ara e&b) benefitted from an unusually
large (≥ 100) number of Doppler observations. Thus, it is possible that the true rate of planetary
systems near the 2:1 MMR is significantly greater than one sixth.

     Kepler observations find that at least ∼ 16% of multiple transiting planet candidate systems
contain at least one pair of transiting planets close to a 2:1 period commensurability (1.83 ≤ Pout /Pin ≤
2.18). Even using our narrow range of period ratios, the true rate of systems near a 2:1 period
– 12 –


commensurability could be significantly greater than ∼ 16%, since not all planets will transit. If
we assume that planets near the 2:1 MMR are in the low inclination regime, then the outer planet
should transit ≃ 63% of the time, implying that the true rate of detectable planets near the 2:1
MMR (if both had a favorable inclination) is at least ∼ 25%. If a significant fraction of these
systems are not in the low inclination regime, then the true rate of pairs of planets near the 2:1
MMR would be even larger. This mildly suggests that radial velocity surveys could be missing some
2:1 pairs due to observational degeneracy. To consider this possibility further requires investigating
the expected mass ratios of planets near the 2:1.

     Based on the B11 catalog, neighboring transiting planet candidates tend to have similar radii,
and in the majority of cases, the outer planet is slightly larger. This is consistent with the typical
outcome of planet formation models that predict more distant planets can more easily grow to
larger sizes.

     The distribution of planetary radii ratios (Rp,out /Rp,in ) for neighboring pairs of transiting
planet candidates near a 2:1 period commensurability is concentrated between 0.9 and 1.25 with
a tail extending to larger ratios (see Fig. 10, solid curve). For reasonable assumptions of a mass-
radius ratio (Eqn. 1), this implies that most neighboring transiting planet candidate systems have
masses within 20% of each other. Fortunately, such pairs are not significantly affected by the radial
velocity observational degeneracy with a single eccentric planet. However, the tail of ∼ 20% of
systems with a radii ratio exceeding 1.7 are more problematic, as the Doppler signatures of would
differ from that of a single eccentric planet by less than 20% of the velocity amplitude of the outer
planet.

    Based on the planet radii ratio distribution from Kepler, we estimate that the abundance of
neighboring planets near the 2:1 MMR could be ∼ 20% greater than suggested by Doppler surveys
due to the difficulty in distinguishing the Doppler signature of such systems from a single eccentric
planet. A more precise determination of the frequency of resonant or near-resonant systems based
on Kepler and/or Doppler surveys is left for future work.



                   5.3.   KOI-730: A Multi-Resonant Candidate System

     While few nearly exact mean motion resonances are evident in the sample of Kepler planetary
candidates, one system stands out as exceptional: the periods of the four candidates in KOI-
730 satisfy the ratio 6:4:4:3 to ∼ 1 part in 1000 or better. This system is the first to show
evidence for extrasolar “Trojan” or co-orbital planets, which have been suggested to be theoretically
possible (Laughlin & Chambers 2002; Go´dziewski & Konacki 2006; Smith & Lissauer 2010). In
                                           z
Kepler data, the two co-orbital candidates began separated by ∼118◦ , with the trailing candidate
reducing the gap at the rate of ∼1◦ per month. The numerical integration (section 4) matching
the periods and phases in B11, circular orbits, and nominal masses went unstable at 25 Myr,
precipitated by a close encounter between the coorbital planets.
– 13 –


     This system will be difficult to study because of the faintness of the target star (Kepler magni-
tude 15.34) and its location within the field which implies that it cannot be observed by Kepler dur-
ing winter quarters (Q4, Q8, ...). Nonetheless, the remarkably commensurate period ratios of these
four candidates give us strong confidence that they all will eventually be confirmed as planets. The
dynamics of this class of planetary systems are analyzed by Fabrycky et al. (2011; in preparation),
whose results further strengthen our confidence in these candidates.



                   5.4.   KOI-191: A System Protected by Resonance?

     Two of the planets in the KOI-191 system have a period ratio of 1.258. One of those planets
has a ∼ 1RJup size. Its mass would need to be 17M⊙ to satisfy equation 3. Alternatively, the
two planets could be locked in 5:4 mean-motion resonance that prevents close approaches even if
orbits cross. Neptune and Pluto are locked in 3:2 mean-motion resonance of this type. Relatively
fast precession, on the order of 100 orbital periods, would be needed as well, as the observed period
ratio is significantly different from the nominal resonance location.



                5.5.   KOI-500: (Near-)Resonant Five-Candidate System

     KOI-500 is a 5-candidate system with periods 0.986779 d, 3.072166 d, 4.645353 d, 7.053478
d, and 9.521696 d. Neighboring pairs of the outer four of these planets are all near two-body
mean motion resonances, but unless there is an unexpectedly large amount of apse precession,
they do not appear to be locked in librating two-body resonances. Perhaps this is a consequence
of diverging tidal evolution, as predicted by Papaloizou & Terquem (2010). The three adjacent
period spacings between these four planets contribute to the statistics of period ratios just wide of
resonance (negative ξ). Moreover, the spacing of these systems from exact resonance,

                                    Poffset = 1/(1/Pout − 1/Pin )                                 (7)

is equal to ∼ 190 days in all cases. The combination of mean motions 2n2 − 5n3 + 3n4 ≈ 1.6 × 10−5
and 2n3 − 6n4 + 4n5 ≈ 1.3 × 10−5 are so small that we suspect two 3-body Laplace-like resonances
and/or a 4-planet resonance may be active and controlling the dynamics of this system.



                            6.   Coplanarity of Planetary Systems

     The orbits of the planets in our Solar System lie close to the same plane. The distribution
of planetary orbits around the invariable plane of the Solar System has a Rayleigh distribution
width of σi ≃ 1.5◦ ; if Mercury is excluded, then σi ≃ 1◦ . The strong level of co-planarity has been
recognized as an important constraint on models of planet formation for over 250 years (Kant 1755;
Laplace 1796) and our current understanding is that growth within a dissipative protoplanetary
– 14 –


disk generally yields circular orbits and low relative inclinations (e.g., Safronov 1969; Lissauer
1993). Some mechanisms for explaining the higher eccentricity distribution of large exoplanets
are likely to increase inclinations as well, such as planet-planet scattering (e.g., Chatterjee et al.
2008; Juri´ & Tremaine 2008). Whatever the mechanism, it is clear that measuring the inclination
           c
distribution of typical planetary systems will create an important test for planet formation theories.

     The interplay between effects due to the protoplanetary disk, distant perturbers, and planet-
planet scattering will leave a unique fingerprint in the coupled eccentricity-inclination distribution
of exoplanetary systems and the exoplanet inclination distribution is thus of immense value. Yet
neither transit observations, nor the radial velocity technique can directly measure the true mutual
inclination between planets observed in multiple systems, except in unusually fortuitous circum-
stances. This continues to be true in multi-transiting systems: even though the inclinations to
the line of sight of all transiting planets must be small, the orbits could be rotated around the
line of sight and mutually inclined to each other. Indirect constraints, however, can be obtained
for systems with multiple transiting planets that make them the best probe of mutual inclinations
around main sequence stars. For example, the lack of transit duration variations of the planets of
Kepler-9 (Holman et al. 2010) and Kepler-11 (Lissauer et al. 2011) already place interesting lim-
its on mutual inclinations to be 10◦ , while the inclinations to the line-of-sight of the Kepler-10
planets (Batalha et al. 2011, in prep.) require a mutual inclination 5◦ .

     Furthermore, given the large number of multi-candidate systems identified in B11, we can
approach the question of the inclination distribution of the population statistically. The more
mutually inclined a given pair of planets is, the smaller the probability that multiple planets
transit (Ragozzine & Holman 2010), and thus the smaller the fraction of multi-planet systems that
we would expect to see. Therefore, an investigation of coplanarity requires also addressing the
number of planets expected per planetary system. Of course, the question of planetary multiplicity
is also inherently interesting to investigate. Here we attempt to provide a self-consistent rigorous
estimate of the multiplicity and coplanarity of typical systems using a statistical approach.



                              6.1.   Simulated Population Model

                                     6.1.1.   Model Background

     The number of observed transiting planets per star is clearly affected by multiple observational
biases, most notably that the probability of transiting decreases with increasing orbital period.
With the current observations, we have little to no insight on the non-transiting planets in these
systems, though eventually investigations of transit timing and duration variations (or the lack
thereof) will be able to place additional limits on such planets (Steffen et al. 2010, Ford et al. in
preparation). If we knew about the presence of every planet, transiting or not, then the problem
of determining the typical inclination would be much easier.
– 15 –


     A useful way to address the problem of non-detections is to create a forward model of transiting
planet detections based on a minimal number of assumptions and a relatively small number of
tunable parameters and to compare the output with the properties of the observed systems. The
model described here attempts use Kepler ’s observed frequency of planet multiplicity to determine
typical values for the true multiplicity of planetary systems as well as the mutual inclination between
planets in the same system. After choosing a distribution for the number of planets assigned to a
random Kepler star (characterized by a parameter Np ) and their relative inclinations (characterized
by σi ), the model computes how many of these planets would be detected, requiring a geometrical
alignment that leads to transit and a signal-to-noise ratio (SNR) large enough to be detected. This
process is done for over 100000 planetary systems and the results compared to frequency of observed
multi-transiting systems. Though imperfections in the model and observations will only provide an
approximate answer, Kepler enables for the first serious attempt at using observations to measure
the value of Np and σi for a large number of systems.

     To isolate the true multiplicity and inclination distributions, other aspects of the simulated
systems (such as radius and period distributions) must be consistent with the true underlying
distribution. This requires some interpretation of the observed distribution and the identification
of potential biases. At the end of Kepler ’s mission, a simulated population that attempts to match
all the observations could be employed; here, we hold the period and radius distributions fixed and
they are pre-selected in a way that the output of the simulation will closely match the observed
distribution.

     A good example of this process is the choice of the period distribution. The difference between
the observed and true underlying period distribution is almost entirely determined by the geometric
bias of observing planets in transit that scales as 1/a ∝ P −2/3 . As discussed in B11 and confirmed by
our investigations, after correcting for the geometric bias and discarding very short periods (below
3 days) and large planets (above 6 Earth radii), the remaining distribution is very well described
by a uniform distribution in log(P ). Hence, our simulator uses a period distribution uniform in
log(P ) for a pre-specified range of periods: 3-125 days. An improvement could be placed on this
process that takes into account the observed relationship (generally expected from planet formation
theories), that larger period planets have larger radii (Figure 10), but this covariance is not included
in the simulated planetary systems discussed below.

     Estimating the true radius distribution of Kepler planets is more complicated. A simple his-
togram of planet radii shows that above ∼2 Earth radii, the observed population follows approxi-
mately a 1/r 2 distribution (B10, B11). Below ∼1.5-2 Earth radii, the number of planets decreases
significantly, which is at least partly due to small planets escaping detection due to low SNR. The
1/r 2 distribution is very sensitive to the minimum allowed radius; to avoid this, we draw radii from
the distribution of observed planetary radii around stars with Kepler magnitude greater than 14
and through-Q5 SNR (from Table 1 in B11) greater than 16. A histogram of this distribution, and
the full distribution of all detected planets, is shown in Figure 11. In particular, our simulated
planets have radii drawn from this distribution between the limits of 1.5 and 6 Earth radii; except
– 16 –


at the smallest radii, this nearly approximates the 1/r 2 distribution observed in B11.

    Finally, an inspection of the B11 SNR ratios (which are calculated from Quarters 1-5) suggest
that completeness effects begin to dominate below a SNR of ∼16, corresponding to a SNR of 9.2
from Quarters 1-2. We require that our simulated planets match this SNR in order to be detectable.
We did not consider the FOP vetting flag in our choice of systems as a large fraction of the multiples
have not yet been investigated in detail.

     Applying these same cuts to Kepler ’s observed multi-candidate systems from B11 results in
437, 63, 17, 1, and 1 systems with 1, 2, 3, 4, and 5 transiting planets, respectively. It is possible
that some small fraction of the observed planets could be false positives and our simulation does
not attempt to account for this. Note that the sixth planet of Kepler-11, planet g, had a transit
that fell into the data gap between Quarters 1 and 2, so this system counts as a quintuple for
comparison to the simulated population (a minor difference that does not affect our results).



                                     6.1.2.   Model Description

     Under the constraints described above, we want to create simulated planetary populations that
will be compared to the Kepler observations. In particular, for each simulated population, we will
compared the number of systems with i detectable transiting planets for 1 ≤ i ≤ 5 to the observed
distribution from Kepler .

      The model starts by taking the full stellar population from the 161836 Q2 target stars in the
“EX”oplanet program (the list of which is available from MAST), with the assumed masses and
radii from the Kepler Input Catalog (Latham et al. 2005, Brown et al. in preparation). After
randomly choosing a star, the model then assigns a number of planets, Np . Three distributions
were considered for the number of planets: a “uniform” distribution, a Poisson distribution, and an
exponential distribution. In the exponential distribution, the probability of having i + 1 planets is
a factor of α smaller than the probability of having i planets, with α a free parameter. This model
created a very poor fit to the data (particularly the triples to doubles ratio) for any value of α or
σi , so we do not consider it further.

     The “uniform” distribution parameter Np,U is the mean number of planets per system, which
is a fixed number of planets per star if Np,U is an integer or a random mix of two fixed values if
Np,U is not an integer. For example, if Np,U = 3.25, 75% of systems would be assigned 3 planets
and 25% would be assigned 4 planets. The Poisson distribution parameter Np,λ assigns the number
of planets per system based on a Poisson distribution with width Np,λ . The frequency of stars with
planetary systems is considered separately (see below), so we require the number of planets to be
non-zero; for Np,λ 1, the distribution differs somewhat from a true Poisson distribution.

    Once the each star is hosting a specific non-zero number of planets, the orbital and physical
properties of these planets are assigned. As discussed above, the period of each planet is selected
– 17 –


by drawing from a uniform distribution in log(P ) with a minimum of 3 and a maximum of 125
days. Although the period ratios in multiple systems deviate from a random distribution due to
clustering near resonance (Section 5), this is a minor effect compared to the overall geometric
bias. Each planetary radius is assigned based on the Kepler observed radii independently of orbital
period, as described above.

     As the observed multiples are almost entirely stable on long timescales (Section 4), there is a
desire to impose a proximity constraint that rejects simulated multiples that are too closely packed.
This will require assigned a mass to each planet as well. We follow the mass-radius relation used
above (eq. 1), i.e., M = R2.06 with masses and radii in units of Earth masses and radii. The mutual
Hill separation between each pair of planets, ∆ can then be calculated as described in equation 3.

     In practice, the systems are built one planet at a time until all the planets assigned to this star
(based on Np ) are selected. The first planet is assigned its period, radius, and corresponding mass.
Starting with the second planet, a new period, radius, and mass are assigned and ∆ calculated. If
any two planets have ∆ < 3.46 or if ∆inner + ∆outer < 18 for any three consecutive planets, then the
planet is rejected on stability grounds and the process begins again until all the planets have been
assigned. Very rarely, over the wide parameter space we are considering, large numbers of planets
( 10) are assigned which will not fit within the period range and stability criteria; in this case, as
many planets as can fit are given to the star.

     Inclinations with respect to a reference plane are drawn with a Rayleigh distribution (appro-
priate for this kind of geometric problem) of width σi and random nodal longitudes are assigned.
In general, we assume that eccentricities are zero, and our investigations that considered non-zero
eccentricities indicated that, statistically speaking, the observed multiplicities did not change sig-
nificantly, though we did not consider the possibility that eccentricities and inclinations may be
correlated.

     At this point, the full orbital and physical characteristics of the planetary systems have been
determined and we can begin to assess observability. First, these planetary systems are then rotated
by a uniform random rotation matrix in a manner equivalent to choosing a random point on a sphere
for the direction of the normal to the reference plane. In the new random orientation, the impact
parameter of each planet is calculated. If the planets are transiting, the SNR of a single transit is
calculated assuming a box-shaped transit with depth of (Rp /R∗ )2 and duration calculated from the
equation in (Kipping 2010). A random epoch is assigned and the number of transits in a 127-day
period (corresponding to the duration of Q1-Q2) is calculated. To account for the duty cycle of
92%, each of these transits can be independently and randomly lost with 8% probability. The
total signal is then calculated as the signal for a single transit times the square root of the number
of observed transits. This is compared to the estimated noise over the course of the duration of
the transit, calculated from the TMCDPP values for this star from Quarter 2 (see Kepler Data
Release Notes). As described above, we require that our simulated planets have SNR of 9.2 in
the simulated Quarters 1 and 2 in order to be detected. We find that approximately 30-50% of
– 18 –


simulated transiting planets are “missed” due to insufficient SNR.

     The output of the model is the number of stars with N detectable transiting planets, for all N .
The next step is a statistical comparison between a variety of simulated populations (with different
choices for Np and σi ) with the observed distribution from Kepler . An assessment of whether
a particular simulation is rejectable is done using a G-test (similar to the categorical χ2 test) by
calculating the G-statistic:
                                        G=      2Oi ln(Oi /Ei )                                   (8)
                                             i

where Oi is the number of systems with i planets observed by Kepler and Ei is the expected number
of such systems from the simulated population, which is scaled so that the total values of O and
E are the same. The values Oi and Ei always refer to the number of systems with i observed and
expected detectable transiting planets. Note that this statistic gives the most weight to categories
with large numbers of objects and no weight to the categories where there are no observed or
expected systems (i.e. i > 5). The G-statistic grows as the two populations become more different.
Though the value of the G-statistic follows the χ2 distribution when O and E are large enough, it
is more robust to compare the G value to a suite of G-statistics from a Monte Carlo simulation.
Using the expected frequencies Ei , 100 random Monte Carlo populations are created with the true
underlying distribution and their G-statistics are calculated. The significance of the observed value
of G is calculated by comparing it to the Monte Carlo generated G-statistics. For example, if the
observed value of G is 10 and this is larger than 95% of the Monte Carlo values of G, then the
hypothesis that the population observed by Kepler is the same as the simulated population can be
rejected with 95% confidence.



                                 6.2.   Results and Discussion

     The simulated population assumes that all 161836 Q2 target stars have planetary systems.
B11 report that the expected number of planets per star based on the Kepler observations is about
0.4, though this describes planets of all types (unlike the limited size range we are considering)
and converting this to the fraction of stars with planetary systems requires dividing by the typical
multiplicity. Our investigation could address the frequency of planetary systems by adding a new
free parameter fp and fitting to Oi with 0 ≤ i ≤ 5. However, this fit would be equivalent to: fitting
a scaled population of systems that all contained planets to Oi , i > 0; calculating the scaled number
of systems that had planets but did not transit, E0 ; and setting E0 = (1 − fp )O0 . That is, we can
return to the question of the frequency of planetary systems after we find simulated populations
that match the observed frequencies of systems where at least one transiting planet is detected.

     The first result is that there is no simulated population that adequately fits the observed
distribution Oi for 1 ≤ i ≤ 5. The only population that adequately describes the singly-transiting
systems has a uniform distribution of two planets per star (Np,U = 2) and an inclination width of
σi = 4◦ . While the G-statistic shows that this solution has high confidence, clearly any population
– 19 –


that has only 2 planets per star cannot explain the large number of triples and higher-multiplicities
seen in the Kepler observations. The “good fit” here is biased by the larger weight that the G-
statistic gives to categories with more members and that categories with no observations have
no weight. However, this model is useful to point out as one that correctly reproduces the ratio
of doubles to singles by requiring a typical inclination dispersion of σi = 4◦ , corresponding to a
Rayleigh distribution with a mean of ¯ ≡ σi π/2 ≃ 5◦ .
                                      i

      All other simulated populations produce relatively fewer singly-transiting systems than are
observed. The difficulty can be seen qualitatively by comparing the ratio of doubly-transiting
systems to singly-transiting systems (O2 /O1 = 14.5%) to the ratio of triply-transiting systems to
doubly-transiting systems (O3 /O2 = 25.8%). These two ratios are quite different, yet the proba-
bility of an additional planet transiting depends only on the semi-major axis ratio and the mutual
inclination (Ragozzine & Holman 2010). We have also found that the probability that a planet
passed our detection threshold is more-or-less independent of multiplicity, i.e., the detectability can
be estimated on a planet-by-planet basis. (In detail, systems with high multiplicities tend to have
planets with longer periods which lowers the SNR somewhat.) The simulated populations have
both the geometric and detectability probabilities decreasing at an approximately constant rate as
additional planets are added, leading to ratios Ei+1 /Ei that are nearly constant.

     Unlike in the single homogeneous simulated populations, the observed ratios are not similar
and Kepler seems to be seeing an “excess” of singly-transiting systems. Note that this cannot be
attributed to the fact that “hot Jupiters” appear to be singletons (Latham et al. 2011, in prep),
since we are considering objects that are Neptune size or smaller. While it is possible that the
inability to create a simulated population that matches the observations of multiples by Kepler is
due to a shortcoming in the model, a theoretically motivated alternative is that the true planet
population that has a bimodal distribution of characteristic inclination, with some have small
inclinations and typically more planets than the other group. This is strongly suggested for distant
planets (e.g., Levison et al. 1998) and may be present in the close-in population we are modeling
here due to a distinction between systems that have had large dynamical scatterings in the past
versus those systems which have remained relatively calm.

      As we are only attempting to match 5 multiplicity frequencies Oi , we cannot justify adding a
full second population to our fits since this would require 4 highly degenerate parameters. Instead,
we suppose that there may be a excess population of singly-transiting systems, and the simulated
populations are compared again to the observations, but are now fit only to the observed population
with two or more transiting planets (i.e., 2 ≤ i ≤ 5) using the same G-statistic method described
above. Most of the constraint on this population is from the ratio of O3 /O2 . Note that a second
planetary population with a large inclination dispersion would also contribute slightly to the number
of doubles and possibly triples, but that we are ignoring this contribution.

    The best fit (not rejectable at the 93% level) in this case is from a population of systems with
Np,U = 3.25 (i.e., 75% of systems with 3 planets and 25% of systems with 4 planets) and with a
– 20 –


Rayleigh inclination width of σi = 2◦ , implying a mean inclination of ¯ ≃ 2.5◦ . This population
                                                                           i
produced Ei = (62.1, 17.5, 1.4, 0) for i from 2-5, which is an excellent match to the observed values
Oi = (63, 17, 1, 1). (Recall that the expected numbers are scaled so that the total number of planets
is the same as the observations.) When scaled to the observed values, this population produced
only E1 = 130.1 singly-transiting systems, suggesting that over half of the observed singles cannot
be explained in a simple model that matches the higher multiplicities.

      Note that these fits have only 1 degree of freedom (4 categories - 1 from scaling - 2 parameters),
so it is not a surprise that a wide variety of models provide moderately good fits to the data. Figure
12 gives a contour plot of the acceptability of simulated populations as a function of Np,U and σi . As
expected, increasing the number of planets per system requires the inclination dispersion to increase
in order to provide a good match to the Kepler data, which can be demonstrated quantitatively.
These fits also cannot explain the large number of singly-transiting systems: the better the fit to
Oi , 2 ≤ i ≤ 5, the lower the expected number of singly-transiting systems.

     Some of the simulated populations where the number of planets is taken from a Poisson dis-
tribution also give moderately good fits. The best is rejectable at the 70% level and has Np,λ = 2.5
with σi = 4◦ . This simulation can (rarely) produce systems with more than 4 detectable transiting
planets, Ei = (65.3, 14.3, 1.9, 0.4) for i from 2-5, but underproduces the number of triply-transiting
systems. A contour plot showing these the acceptability levels as a function of Np,λ and σi is shown
in Figure 13. This figure also demonstrates the trend that more planets require a larger inclination
dispersion to match the Kepler observations.

     It is expected that not every star has planets in this size and period range that would have
been detected by Kepler ; the Sun is an example of such a star. The population of doubles that best
matched the observations for 1 ≤ i ≤ 5 required E0 =9699 stars, i.e., ∼6% of the Kepler sample. The
best-fit population for 2 ≤ i ≤ 5 described above (Np , U = 3.25, σi = 2◦ ) required only E0 = 3225
stars (∼2%) to have, on average, 3.25 planets (within the period and radius limits) to reproduce the
observations seen by Kepler . The Poisson distributed simulated population (Np , λ = 2.5, σi = 4◦
required 4925 stars (∼3%). (Other simulated populations that were adequate fits as shown in
Figures 12 and 13 had similar values of E0 .) In these latter cases, an additional set of stars with
planets are needed to account for the additional singly-transiting systems. Nevertheless, these
numbers appear quite small and fall about a factor of 2 short of the completeness calculations
reported in B11, even after accounting for the fact that we are considering a different size range
and that the number of stars with planetary systems is the average number of planets with stars
(calculated by B11) divided by the typical multiplicity. The source of this discrepancy is unclear.

     In multiple systems with significant width in the inclination distribution, a substantial fraction
of observers will see at least one planet in transit, even if the periods are somewhat large. Our
simulations suggest that only a small fraction (∼5%) of the ∼162000 stars observed by Kepler have
large numbers of planets with periods less than 125 days and radii between 1.5 and 6 Earth radii.
As our completeness SNR cutoff is just an approximation to the actual fraction of planetary signals
– 21 –


that Kepler will detect among all actual planetary transits, the true fraction of stars with planets
will be different from this estimate, with a more detailed assessment of completeness to be done by
the Kepler Mission in the future. For example, the large errors and, in some cases, biases in the
KIC will propagate into our simulations to some degree, though the ratio of observed multiplicities
which drive the fits above are not as strongly affected as the estimate of the number of stars with
planetary systems.

     With these caveats, Kepler is providing incredibly powerful insight into the formation of plan-
etary systems. A single population model does not match the observed distribution, suggesting a
superposition of two populations. A small fraction of stars have multiple (3 or more) super-Earth or
Neptune size planets with periods less than 125 days. It appears that many of the doubly-transiting
systems and less than half of the singly-transiting systems are actually systems with 3-4 planets,
with non-transiting planets potentially observable through TTVs (Ford et al. 2011, in preparation).
Systems with large values of ∆ in Tables 1 and 2 are good candidates for systems where additional
non-transiting planets may be missing in between.

     We have also provided the first estimate of the inclination dispersion that has been given for a
large sample of observed planetary systems. To match the numbers of higher multiples observed by
Kepler requires a population with relatively low inclination dispersion; mean inclinations greater
than 5◦ are poor fits. It does not appear that all of the multiples observed by Kepler come from
a population of purely coplanar systems. Systems with multiple super-Earth and Neptune size
planets with periods less than 125 days at low relative inclinations suggest interactions with a
protoplanetary disk that induce migration while damping inclinations.

     There are good prospects for improving our understanding of the planetary populations ob-
served by Kepler . Besides continued observations which will increase the detectability of planets
and find new candidates in multiple systems, a more detailed population simulator could be de-
veloped. For example, though the errors in the impact parameters of Kepler planets are usually
relatively large, they do contain information on relative inclinations. This can be used both in a
statistical population sense as well as for individual systems.

     In some systems, the detection of or upper limits to Transit Duration Variations (TDVs) will
constrain the unobserved component of the relative inclinations of planets. Potential mutual events
in Kepler multi-planet systems will be discussed and studied in more detail by Ragozzine et al.
(2011, in preparation). Measurements of the Rossiter-McLaughlin effect for planets within multi-
planet systems to measure true mutual inclination would constrain orbital inclinations to star’s
equator, though this is observationally challenging for most Kepler stars (Ragozzine & Holman
2010). Each measurement of the true mutual inclination in individual systems helps fill in the
picture for the typical inclination distribution of planetary systems.
– 22 –


                7.   How Rare are Planetary Systems Similar to Kepler-11

     The Kepler-11 (= KOI-157) planetary system, with six transiting planets, is clearly an outlier
to the number/frequency distribution for detected transiting planets. Moreover, the period ratio
of the two inner planets is only 1.264, which is the 4th lowest ratio among the 238 neighboring
pairs of transiting Kepler exoplanets (Figure 3). The smallest period ratio is 1.001, between the
candidate co-orbital planets in KOI-730 (Section 5.3). The second smallest period ratio, 1.038, is
between two weak (vetting flag 3) candidates in the KOI-284 system; this system would be unstable
for any reasonable planetary masses if indeed both of these candidates are actually planets in orbit
about the same star (section 4). The third smallest period ratio, 1.258, is for a pair of candidates in
the KOI-191 system that stability considerations strongly suggest are locked in a 5:4 mean motion
resonance that prevents close approaches (section 5.4). Thus, Kepler-11b and Kepler-11c may well
have the smallest period ratio of any non-resonant pair of planets in the entire candidate list.
Additionally, the five inner planets of Kepler-11 travel on orbits that lie quite close to one another,
both in terms of period ratio and absolute distance. Kepler-11 clearly does not possess a “typical”
planetary system, but are systems of this type quite rare, or simply somewhat scarce?

     The two candidate most analogous to Kepler-11 are KOI-500 and KOI-707. With 5 planet
candidates, KOI-500 is second in abundance to Kepler-11 and four of these candidates are close to
one another in terms of period ratio (Table 4). The target star is significantly smaller than the
Sun (∼0.74 R⊙ ; B11), and the orbital periods of the candidates are shorter, so if confirmed the
planets in this system would be most closely spaced in terms of physical distance between orbits
of any several planetary systems known. Moreover, four of the planets orbiting KOI-500 appear
to be locked in three-planet resonances (Section 5.5), whereas no analogous situation is observed
in Kepler-11. Unfortunately, KOI-500 is quite faint, so it will be more difficult to study than is
Kepler-11. The four candidates in KOI-707 have periods within the range of the periods of the 5
inner planets observed in the Kepler-11 system. Periods ratios of neighboring candidates in the
four planet KOI-117 are all less than two, but nonetheless significantly larger than in Kepler-11
and KOI-707. None of the six other four candidate systems is nearly as closely spaced.

     Furthermore, the better-fitting population simulations described in 6 rarely produced 5-planet
systems, suggesting that Kepler-11 is not just a natural extension of Kepler multiple systems.
Although its solitary nature makes determining a lower bound on planetary systems like Kepler-11
an ill-posed question (Kepler could have just gotten lucky), we can estimate the number of Kepler-
11 systems that would not be seen in transit probabilitistically. If all the Kepler-11 planets are
coplanar (which is slightly inconsistent with the observed impact parameters), then the probability
of seeing all six planets transit is the same as seeing Kepler-11g transit, which is 1.2%, suggesting
that there are dozens of systems similar to Kepler-11, unless we were observing from a particularly
favorable position.
– 23 –


                                         8.   Conclusions

     Analysis of the first four months of Kepler data reveals 170 targets with more than one tran-
siting planet candidate (B11). While the vast majority of these candidates have yet to be validated
as true planets, we expect that the fidelity of this subsample of Kepler planet candidates to be high
(Section ), and the small fraction of false positives not to affect the robustness of our statistical
results.

    Our major conclusions are:

  1. The large number of multiple planet systems observed by Kepler show that flat multi-planet
     systems are common in short-period orbits around other stars. This result holds for planets
     in the size range of ∼ 1.5 – 6 R⊕ , but not for giant planets. Not enough data are yet available
     to assess its viability for Earth-size and smaller worlds, nor for planets with orbital periods
     longer than a few months. The abundance and distribution of multiply transiting systems
     implies that the mean number of planets per star with at least one planet of this type is ∼3-4.
     Thus, only ∼3% of stars have planets within these in size-period limits.

  2. Almost all candidate systems survived long-term dynamical integrations that assume circular,
     planar orbits and a mass-radius relationship derived from planets within our Solar System 1.
     This bolsters the evidence for the fidelity of the sample.

  3. Mean-motion orbital resonances and near resonances are favored over random statistical prob-
     ability, but only a small fraction of multiple candidates are in or very near such resonances.
     In particular, resonances seem to be less common among nearly coplanar planets with short
     periods and radii a few times that of Earth than among the typically more massive and dis-
     tant planets detected by RV (Wright et al. 2011). Nonetheless, we note that a few systems
     are likely to have particularly interesting resonances configurations (Section 5).

  4. Taken together, the properties of the observed systems suggest that the majority are free from
     false positives. It is unlikely that all of the candidates in multiple system are false positives,
     but there may be systems where one of the candidates is not a transiting planet around the
     same star as the other candidate(s). Those very near mean-motion resonances are most likely
     to be true planetary systems.

  5. No single population of multiple planets with a single inclination dispersion can describe all the
     Kepler transiting systems. It appears that Kepler systems come from two populations which
     produce an excess in singly-transiting systems. Among systems with two or more transiting
     planets, the inclination dispersion appears to have a mean between 1-5◦ , suggesting relatively
     low mutual inclinations similar to the Solar System. Many more Kepler targets are multi-
     planet systems where additional non-transiting planets are present and these may eventually
     show TTVs.
– 24 –


     The rich population of multi-transiting systems discovered by Kepler and reported by B11
have immense value both as individuals and as a population for improving our understanding of
the formation and evolution of planetary systems. The Kepler spacecraft is scheduled to continue to
return data on these multi-planet system for the remainder of its mission, and the longer temporal
baseline afforded by these data will allow for the discovery of more planets and more accurate
measurements of the planets and their interactions.


    Kepler was competitively selected as the tenth Discovery mission. Funding for this mission is
provided by NASA’s Science Mission Directorate. The authors thank the many people who gave
so generously of their time to make the Kepler mission a success. D. C. F. acknowledge NASA
support through Hubble Fellowship grants #HF-51272.01-A and #HF-51267.01-A, respectively,
awarded by STScI, operated by AURA under contract NAS 5-26555. We thank Avi Loeb, Josh
Carter, and others for valuable discussions.



                                         REFERENCES

Agol, E., Steffen, J., Sari, R., & Clarkson, W. 2005, MNRAS, 359, 567

Anglada-Escud´, G., L´pez-Morales, M., & Chambers, J. E. 2010, ApJ, 709, 168
             e       o

Borucki, W. J., Kepler, & Team. 2011a, ApJ

—. 2011b, ApJ, in press

Borucki, W. J., Koch, D., Basri, G., Batalha, N., Brown, T., Caldwell, D., Caldwell, J., Christensen-
      Dalsgaard, J., Cochran, W. D., DeVore, E., Dunham, E. W., Dupree, A. K., Gautier, T. N.,
      Geary, J. C., Gilliland, R., Gould, A., Howell, S. B., Jenkins, J. M., Kondo, Y., Latham,
      D. W., Marcy, G. W., Meibom, S., Kjeldsen, H., Lissauer, J. J., Monet, D. G., Morrison,
      D., Sasselov, D., Tarter, J., Boss, A., Brownlee, D., Owen, T., Buzasi, D., Charbonneau, D.,
      Doyle, L., Fortney, J., Ford, E. B., Holman, M. J., Seager, S., Steffen, J. H., Welsh, W. F.,
      Rowe, J., Anderson, H., Buchhave, L., Ciardi, D., Walkowicz, L., Sherry, W., Horch, E.,
      Isaacson, H., Everett, M. E., Fischer, D., Torres, G., Johnson, J. A., Endl, M., MacQueen,
      P., Bryson, S. T., Dotson, J., Haas, M., Kolodziejczak, J., Van Cleve, J., Chandrasekaran,
      H., Twicken, J. D., Quintana, E. V., Clarke, B. D., Allen, C., Li, J., Wu, H., Tenenbaum,
      P., Verner, E., Bruhweiler, F., Barnes, J., & Prsa, A. 2010, Science, 327, 977

Bryson, S. T., Tenenbaum, P., Jenkins, J. M., Chandrasekaran, H., Klaus, T., Caldwell, D. A.,
      Gilliland, R. L., Haas, M. R., Dotson, J. L., Koch, D. G., & Borucki, W. J. 2010, ApJ, 713,
      L97

Caldwell, D. A., Kolodziejczak, J. J., Van Cleve, J. E., Jenkins, J. M., Gazis, P. R., Argabright,
      V. S., Bachtell, E. E., Dunham, E. W., Geary, J. C., Gilliland, R. L., Chandrasekaran, H.,
– 25 –


      Li, J., Tenenbaum, P., Wu, H., Borucki, W. J., Bryson, S. T., Dotson, J. L., Haas, M. R.,
      & Koch, D. G. 2010, ApJ, 713, L92

Chambers, J. E. 1999, MNRAS, 304, 793

Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008, ApJ, 686, 580

Dawson, R. I., & Fabrycky, D. C. 2010, ApJ, 722, 937

Fabrycky, D. C. 2009, in IAU Symposium, Vol. 253, IAU Symposium, 173–179

Ford, E. B., Quinn, S. N., & Veras, D. 2008, ApJ, 678, 1407

Giuppone, C. A., Tadeu dos Santos, M., Beaug´, C., Ferraz-Mello, S., & Michtchenko, T. A. 2009,
                                            e
     ApJ, 699, 1321

Gladman, B. 1993, Icarus, 106, 247

Goldreich, P. 1965, MNRAS, 130, 159

Go´dziewski, K., & Konacki, M. 2006, ApJ, 647, 573
  z

Holman, M. J. 2010, in EAS Publications Series, Vol. 42, EAS Publications Series, ed.
     K. Go˙ dziewski, A. Niedzielski, & J. Schneider, 39–54
          z

Holman, M. J., Fabrycky, D. C., Ragozzine, D., Ford, E. B., Steffen, J. H., Welsh, W. F., Lissauer,
     J. J., Latham, D. W., Marcy, G. W., Walkowicz, L. M., Batalha, N. M., Jenkins, J. M.,
     Rowe, J. F., Cochran, W. D., Fressin, F., Torres, G., Buchhave, L. A., Sasselov, D. D.,
     Borucki, W. J., Koch, D. G., Basri, G., Brown, T. M., Caldwell, D. A., Charbonneau, D.,
     Dunham, E. W., Gautier, T. N., Geary, J. C., Gilliland, R. L., Haas, M. R., Howell, S. B.,
     Ciardi, D. R., Endl, M., Fischer, D., F¨r´sz, G., Hartman, J. D., Isaacson, H., Johnson,
                                              ue
     J. A., MacQueen, P. J., Moorhead, A. V., Morehead, R. C., & Orosz, J. A. 2010, Science,
     330, 51

Holman, M. J., & Murray, N. W. 2005, Science, 307, 1288

Jenkins, J. M., Caldwell, D. A., Chandrasekaran, H., Twicken, J. D., Bryson, S. T., Quintana,
      E. V., Clarke, B. D., Li, J., Allen, C., Tenenbaum, P., Wu, H., Klaus, T. C., Middour,
      C. K., Cote, M. T., McCauliff, S., Girouard, F. R., Gunter, J. P., Wohler, B., Sommers, J.,
      Hall, J. R., Uddin, A. K., Wu, M. S., Bhavsar, P. A., Van Cleve, J., Pletcher, D. L., Dotson,
      J. A., Haas, M. R., Gilliland, R. L., Koch, D. G., & Borucki, W. J. 2010, ApJ, 713, L87

Juri´, M., & Tremaine, S. 2008, ApJ, 686, 603
    c

Kant, I. 1755, Allgemeine Naturgeschichte und Theorie des Himmels

Kipping, D. M. 2010, MNRAS, 407, 301
– 26 –


Koch, D., Borucki, W., Cullers, K., Dunham, E., Webster, L., Miers, T., & Reitsema, H. 1996,
      J. Geophys. Res., 101, 9297

Koch, D. G., Borucki, W. J., Basri, G., Batalha, N. M., Brown, T. M., Caldwell, D., Christensen-
      Dalsgaard, J., Cochran, W. D., DeVore, E., Dunham, E. W., Gautier, T. N., Geary, J. C.,
      Gilliland, R. L., Gould, A., Jenkins, J., Kondo, Y., Latham, D. W., Lissauer, J. J., Marcy,
      G., Monet, D., Sasselov, D., Boss, A., Brownlee, D., Caldwell, J., Dupree, A. K., Howell,
      S. B., Kjeldsen, H., Meibom, S., Morrison, D., Owen, T., Reitsema, H., Tarter, J., Bryson,
      S. T., Dotson, J. L., Gazis, P., Haas, M. R., Kolodziejczak, J., Rowe, J. F., Van Cleve,
      J. E., Allen, C., Chandrasekaran, H., Clarke, B. D., Li, J., Quintana, E. V., Tenenbaum, P.,
      Twicken, J. D., & Wu, H. 2010, ApJ, 713, L79

Laplace, P. 1796, Exposition du Systeme du Monde

Latham, D. W., Brown, T. M., Monet, D. G., Everett, M., Esquerdo, G. A., & Hergenrother, C. W.
      2005, in Bulletin of the American Astronomical Society, Vol. 37, Bulletin of the American
      Astronomical Society, 1340

Laughlin, G., & Chambers, J. E. 2002, AJ, 124, 592

Levison, H. F., Lissauer, J. J., & Duncan, M. J. 1998, AJ, 116, 1998

Lissauer, J. J. 1993, ARA&A, 31, 129

Lissauer, J. J., Fabrycky, D. C., Ford, E. B., & et al. 2011, Nature, 470, 53

Lissauer, J. J., & Slartibartfast. 2008, in Astronomical Society of the Pacific Conference Series,
       Vol. 398, Astronomical Society of the Pacific Conference Series, ed. D. Fischer, F. A. Rasio,
       S. E. Thorsett, & A. Wolszczan, 491

Malhotra, R. 1998, in Astronomical Society of the Pacific Conference Series, Vol. 149, Solar System
      Formation and Evolution, ed. D. Lazzaro, R. Vieira Martins, S. Ferraz-Mello, & J. Fernandez
      , 37

Nobili, A., & Roxburgh, I. W. 1986, in IAU Symposium, Vol. 114, Relativity in Celestial Mechanics
       and Astrometry. High Precision Dynamical Theories and Observational Verifications, ed.
       J. Kovalevsky & V. A. Brumberg, 105–110

Ragozzine, D., & Holman, M. J. 2010, preprint (arxiv:1006.3727)

Ragozzine, D., & Wolf, A. S. 2009, ApJ, 698, 1778

Safronov, V. S. 1969, Evolution of the protoplanetary cloud and formation of the Earth and planets.

Saha, P., & Tremaine, S. 1994, AJ, 108, 1962

Smith, A. W., & Lissauer, J. J. 2009, Icarus, 201, 381
– 27 –


—. 2010, Celestial Mechanics and Dynamical Astronomy, 107, 487

Steffen, J. H., Batalha, N. M., Borucki, W. J., Buchhave, L. A., Caldwell, D. A., Cochran, W. D.,
      Endl, M., Fabrycky, D. C., Fressin, F., Ford, E. B., Fortney, J. J., Haas, M. J., Holman,
      M. J., Howell, S. B., Isaacson, H., Jenkins, J. M., Koch, D., Latham, D. W., Lissauer, J. J.,
      Moorhead, A. V., Morehead, R. C., Marcy, G., MacQueen, P. J., Quinn, S. N., Ragozzine,
      D., Rowe, J. F., Sasselov, D. D., Seager, S., Torres, G., & Welsh, W. F. 2010, ApJ, 725,
      1226

Terquem, C., & Papaloizou, J. C. B. 2007, ApJ, 654, 1110

Torres, G., Fressin, F., Batalha, N. M., Borucki, W. J., Brown, T. M., Bryson, S. T., Buchhave,
       L. A., Charbonneau, D., Ciardi, D. R., Dunham, E. W., Fabrycky, D. C., Ford, E. B.,
       Gautier, III, T. N., Gilliland, R. L., Holman, M. J., Howell, S. B., Isaacson, H., Jenkins,
       J. M., Koch, D. G., Latham, D. W., Lissauer, J. J., Marcy, G. W., Monet, D. G., Prsa, A.,
       Quinn, S. N., Ragozzine, D., Rowe, J. F., Sasselov, D. D., Steffen, J. H., & Welsh, W. F.
       2011, ApJ, 727, 24

Veras, D., & Armitage, P. J. 2004, Icarus, 172, 349

Veras, D., & Ford, E. B. 2010, ApJ, 715, 803

Wisdom, J. 1980, AJ, 85, 1122

Wisdom, J., & Holman, M. 1991, AJ, 102, 1528

Wright, J. T., Fakhouri, O., Marcy, G. W., Han, E., Feng, Y., Johnson, J. A., Howard, A. W.,
      Valenti, J. A., Anderson, J., & Piskunov, N. 2010, ArXiv e-prints

Wright, J. T., Upadhyay, S., Marcy, G. W., Fischer, D. A., Ford, E. B., & Johnson, J. A. 2009,
      ApJ, 693, 1084

Wright, J. T., Veras, D., Ford, E. B., Johnson, J. A., Marcy, G. W., Howard, A. W., Isaacson, H.,
      Fischer, D. A., Spronck, J., Anderson, J., & Valenti, J. 2011, preprint (arxiv:1101.1097)




   This preprint was prepared with the AAS L TEX macros v5.2.
                                            A
– 28 –




Table 1.   Characteristics of Systems with Two Transiting Planets

            KOI #   Rp,1 (R⊕ )   Rp,2 (R⊕ )    P2 /P1     ∆

             72        1.30        2.29       54.083733   90.8
             82        6.84        3.70        1.565732   7.6
             89        4.36        5.46        1.269541   4.8
             112       3.68        1.71       13.770977   55.7
             115       3.37        2.20        1.316598   7.4
             116       4.73        4.81        3.230779   21.6
             123       2.26        2.50        3.274199   34.1
             124       2.33        2.83        2.499341   25.5
             139       5.66        1.18       67.267271   54.4
             150       3.41        3.69        3.398061   26.3
             153       3.17        3.07        1.877382   14.2
             209       7.55        4.86        2.702205   14.8
             220       2.62        0.67        1.703136   17.3
             222       2.06        1.68        2.026806   20.5
             223       2.75        2.40       12.906148   55.8
             232       3.58        1.55        2.161937   20.1
             244       4.53        2.65        2.038993   15.5
             260       1.19        2.19        9.554264   70.3
             270       0.90        1.03        2.676454   49.1
             271       1.99        1.82        1.654454   17.5
             279       4.90        2.09        1.846201   13.4
             282       2.77        0.86        3.252652   36.0
             291       1.45        1.05        3.876548   57.0
             313       3.10        2.17        2.220841   21.0
             314       1.95        1.57        1.675518   14.8
             339       1.47        1.08        3.240242   50.6
             341       3.32        2.26        1.525757   11.0
             343       2.19        1.64        2.352438   28.6
             386       3.40        2.90        2.462733   21.7
             401       6.24        6.61        5.480100   23.3
             416       2.95        2.82        4.847001   35.8
             431       3.59        3.50        2.485534   19.6
             433       5.78        13.40      81.440719   28.7
             440       2.80        2.24        3.198296   30.0
             442       1.86        1.43        7.815905   67.9
             446       2.31        1.69        1.708833   15.5
             448       2.33        3.79        4.301984   28.2
             456       3.12        1.66        3.179075   31.6
             459       3.69        1.28        2.810266   26.6
             464       7.08        2.66       10.908382   33.2
             474       2.34        2.32        2.648401   28.7
             475       2.36        2.63        1.871903   17.0
             490       2.29        2.25        1.685884   15.1
             497       2.49        1.72        2.981091   33.8
             508       3.76        3.52        2.101382   16.3
             509       2.67        2.86        2.750973   25.6
– 29 –




            Table 1—Continued

KOI #   Rp,1 (R⊕ )   Rp,2 (R⊕ )    P2 /P1     ∆

  510      2.67         2.67       2.172874   20.6
  518      2.37         1.91       3.146856   32.3
  523      7.31         2.72       1.340780   4.9
  534      2.05         1.38       2.339334   28.7
  543      1.89         1.53       1.371064   11.1
  551      2.12         1.79       2.045843   23.5
  555      1.51         2.27      23.366026   77.3
  564      2.35         4.99       6.073030   34.0
  573      3.15         2.10       2.908328   28.3
  584      1.58         1.51       2.138058   28.5
  590      2.10         2.22       4.451198   44.3
  597      2.60         1.40       8.272799   59.7
  612      3.51         3.62       2.286744   18.1
  638      4.76         4.15       2.838628   19.5
  645      2.64         2.48       2.796998   28.9
  657      1.59         1.90       4.001270   42.5
  658      1.53         2.22       1.698143   18.1
  663      1.88         1.75       7.369380   52.0
  672      3.98         4.48       2.595003   18.8
  676      4.48         2.94       3.249810   20.4
  691      2.88         1.29       1.828544   18.8
  693      1.80         1.72       1.837696   22.2
  700      3.05         1.93       3.297027   32.2
  708      2.19         1.71       2.262657   27.3
  736      2.64         1.96       2.789104   24.7
  738      3.30         2.86       1.285871   6.2
  749      1.98         1.35       1.357411   11.0
  752      2.69         3.39       5.734874   39.3
  775      2.11         2.47       2.079976   17.9
  787      2.87         2.19       1.284008   7.0
  800      2.73         2.52       2.659942   26.3
  837      1.83         1.45       1.919049   22.3
  841      4.00         4.91       2.042803   13.0
  842      2.78         3.14       2.835637   22.9
  853      2.87         2.12       1.767103   15.2
  869      3.19         4.33       4.845276   30.2
  870      3.62         3.45       1.519922   8.9
  877      2.50         2.32       2.021803   17.0
  881      2.54         3.92      10.793324   43.8
  896      3.85         2.81       2.574418   20.6
  904      2.11         2.96      12.635882   50.2
  936      3.46         2.04      10.601824   40.6
  938      2.89         1.28       9.512348   57.5
  945      2.04         2.64       1.575035   13.6
  954      2.35         2.45       4.550137   40.9
 1015      2.36         1.63       2.305817   27.5
– 30 –




            Table 1—Continued

KOI #   Rp,1 (R⊕ )   Rp,2 (R⊕ )    P2 /P1     ∆

 1060     1.54          1.23       2.545144   39.7
 1089     9.64          5.69       7.094116   22.6
 1102     3.22          0.94       1.513919   12.3
 1113     2.65          2.81       3.217321   30.7
 1151     1.17          1.20       1.420286   16.6
 1163     1.90          1.77       2.729492   33.8
 1198     2.04          1.53       1.561840   16.0
 1203     2.51          2.44       2.256556   23.0
 1215     2.21          2.12       1.905318   20.8
 1221     5.03          5.31       1.694236   10.0
 1236     2.78          1.67       5.807534   49.5
 1241     10.43         6.99       2.039957   9.3
 1278     1.57          2.39       3.575571   40.0
 1301     1.86          2.32       2.954666   32.3
 1307     2.96          2.81       2.204782   20.1
 1360     2.73          2.29       2.520266   24.0
 1364     2.86          2.74       2.952861   27.1
 1396     2.53          1.86       1.790301   17.8
 1475     1.79          2.18       5.910858   43.3
 1486     8.45          2.38       8.433622   27.9
 1589     2.23          2.28       1.476360   12.0
 1590     2.75          1.90      10.943201   55.8
 1596     2.26          3.45      17.785655   54.0
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems
Architecture and dynamics of kepler's candidate multiple transiting planet systems

Weitere ähnliche Inhalte

Was ist angesagt?

VABodiut_S2574608_Astrobiology_SETI
VABodiut_S2574608_Astrobiology_SETIVABodiut_S2574608_Astrobiology_SETI
VABodiut_S2574608_Astrobiology_SETIVictor Andrei Bodiut
 
Transiting circumbinary planets_kepler-34b_and_35b
Transiting circumbinary planets_kepler-34b_and_35bTransiting circumbinary planets_kepler-34b_and_35b
Transiting circumbinary planets_kepler-34b_and_35bSérgio Sacani
 
Science 2011-lehner-955-8
Science 2011-lehner-955-8Science 2011-lehner-955-8
Science 2011-lehner-955-8Sérgio Sacani
 
Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...
Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...
Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...Sérgio Sacani
 
A closely packed system of low mass, low-density planets transiting kepler-11
A closely packed system of low mass, low-density planets transiting kepler-11A closely packed system of low mass, low-density planets transiting kepler-11
A closely packed system of low mass, low-density planets transiting kepler-11Sérgio Sacani
 
Kic 9632895 the_10_kepler_transiting_circumbinary_planet
Kic 9632895 the_10_kepler_transiting_circumbinary_planetKic 9632895 the_10_kepler_transiting_circumbinary_planet
Kic 9632895 the_10_kepler_transiting_circumbinary_planetSérgio Sacani
 
Delivery of dark_material_to_vesta_via_carbonaceous_chondritic_impacts
Delivery of dark_material_to_vesta_via_carbonaceous_chondritic_impactsDelivery of dark_material_to_vesta_via_carbonaceous_chondritic_impacts
Delivery of dark_material_to_vesta_via_carbonaceous_chondritic_impactsSérgio Sacani
 
The harps n-rocky_planet_search_hd219134b_transiting_rocky_planet
The harps n-rocky_planet_search_hd219134b_transiting_rocky_planetThe harps n-rocky_planet_search_hd219134b_transiting_rocky_planet
The harps n-rocky_planet_search_hd219134b_transiting_rocky_planetSérgio Sacani
 
Observational evidence for_a_dark_side_to_ngc5128_globular_cluster_system
Observational evidence for_a_dark_side_to_ngc5128_globular_cluster_systemObservational evidence for_a_dark_side_to_ngc5128_globular_cluster_system
Observational evidence for_a_dark_side_to_ngc5128_globular_cluster_systemSérgio Sacani
 
Chapter 10 Tests
Chapter 10 TestsChapter 10 Tests
Chapter 10 Testsdlsupport
 
A dynamical signature_of_multiple_stellar_populations_in_47_tucanae
A dynamical signature_of_multiple_stellar_populations_in_47_tucanaeA dynamical signature_of_multiple_stellar_populations_in_47_tucanae
A dynamical signature_of_multiple_stellar_populations_in_47_tucanaeSérgio Sacani
 
Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5
Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5
Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5Sérgio Sacani
 
Chapter 15 Test
Chapter 15 TestChapter 15 Test
Chapter 15 Testdlsupport
 
Chapter 01 Tests
Chapter 01 TestsChapter 01 Tests
Chapter 01 Testsdlsupport
 
Isolated compact elliptical_galaxies_stellar_systems_that_ran_away
Isolated compact elliptical_galaxies_stellar_systems_that_ran_awayIsolated compact elliptical_galaxies_stellar_systems_that_ran_away
Isolated compact elliptical_galaxies_stellar_systems_that_ran_awaySérgio Sacani
 
The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...
The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...
The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...Eric Roe
 
SPS Colloquium Talk 2011
SPS Colloquium Talk 2011SPS Colloquium Talk 2011
SPS Colloquium Talk 2011Ryan Laird
 

Was ist angesagt? (20)

VABodiut_S2574608_Astrobiology_SETI
VABodiut_S2574608_Astrobiology_SETIVABodiut_S2574608_Astrobiology_SETI
VABodiut_S2574608_Astrobiology_SETI
 
Transiting circumbinary planets_kepler-34b_and_35b
Transiting circumbinary planets_kepler-34b_and_35bTransiting circumbinary planets_kepler-34b_and_35b
Transiting circumbinary planets_kepler-34b_and_35b
 
Science 2011-lehner-955-8
Science 2011-lehner-955-8Science 2011-lehner-955-8
Science 2011-lehner-955-8
 
Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...
Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...
Hats 6b a_warm_saturn_transiting_an_early_m_dwarf_star_and_a_set_of_empirical...
 
A closely packed system of low mass, low-density planets transiting kepler-11
A closely packed system of low mass, low-density planets transiting kepler-11A closely packed system of low mass, low-density planets transiting kepler-11
A closely packed system of low mass, low-density planets transiting kepler-11
 
Kic 9632895 the_10_kepler_transiting_circumbinary_planet
Kic 9632895 the_10_kepler_transiting_circumbinary_planetKic 9632895 the_10_kepler_transiting_circumbinary_planet
Kic 9632895 the_10_kepler_transiting_circumbinary_planet
 
Delivery of dark_material_to_vesta_via_carbonaceous_chondritic_impacts
Delivery of dark_material_to_vesta_via_carbonaceous_chondritic_impactsDelivery of dark_material_to_vesta_via_carbonaceous_chondritic_impacts
Delivery of dark_material_to_vesta_via_carbonaceous_chondritic_impacts
 
The harps n-rocky_planet_search_hd219134b_transiting_rocky_planet
The harps n-rocky_planet_search_hd219134b_transiting_rocky_planetThe harps n-rocky_planet_search_hd219134b_transiting_rocky_planet
The harps n-rocky_planet_search_hd219134b_transiting_rocky_planet
 
A new class_of_planet
A new class_of_planetA new class_of_planet
A new class_of_planet
 
Observational evidence for_a_dark_side_to_ngc5128_globular_cluster_system
Observational evidence for_a_dark_side_to_ngc5128_globular_cluster_systemObservational evidence for_a_dark_side_to_ngc5128_globular_cluster_system
Observational evidence for_a_dark_side_to_ngc5128_globular_cluster_system
 
Chapter 10 Tests
Chapter 10 TestsChapter 10 Tests
Chapter 10 Tests
 
Cheng apj fullres
Cheng apj fullresCheng apj fullres
Cheng apj fullres
 
A dynamical signature_of_multiple_stellar_populations_in_47_tucanae
A dynamical signature_of_multiple_stellar_populations_in_47_tucanaeA dynamical signature_of_multiple_stellar_populations_in_47_tucanae
A dynamical signature_of_multiple_stellar_populations_in_47_tucanae
 
Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5
Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5
Chandra deep observation_of_xdcpj004402033_a_massive_galaxy_cluster_at_z_1_5
 
Chapter 15 Test
Chapter 15 TestChapter 15 Test
Chapter 15 Test
 
Chapter 01 Tests
Chapter 01 TestsChapter 01 Tests
Chapter 01 Tests
 
Clarkson
ClarksonClarkson
Clarkson
 
Isolated compact elliptical_galaxies_stellar_systems_that_ran_away
Isolated compact elliptical_galaxies_stellar_systems_that_ran_awayIsolated compact elliptical_galaxies_stellar_systems_that_ran_away
Isolated compact elliptical_galaxies_stellar_systems_that_ran_away
 
The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...
The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...
The Search for Distant Objects in the Solar System Using Spacewatch - Astrono...
 
SPS Colloquium Talk 2011
SPS Colloquium Talk 2011SPS Colloquium Talk 2011
SPS Colloquium Talk 2011
 

Andere mochten auch

Imprints of fast rotating massive stars in the galactic bulge
Imprints of fast rotating massive stars in the galactic bulgeImprints of fast rotating massive stars in the galactic bulge
Imprints of fast rotating massive stars in the galactic bulgeSérgio Sacani
 
Herschel galactic plane_survey_the_global_distribution_of_ism_gas_componnent
Herschel galactic plane_survey_the_global_distribution_of_ism_gas_componnentHerschel galactic plane_survey_the_global_distribution_of_ism_gas_componnent
Herschel galactic plane_survey_the_global_distribution_of_ism_gas_componnentSérgio Sacani
 
The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240
The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240
The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240Sérgio Sacani
 
High pre eruptive water contents preserved in lunar melt inclusions
High pre eruptive water contents preserved in lunar melt inclusionsHigh pre eruptive water contents preserved in lunar melt inclusions
High pre eruptive water contents preserved in lunar melt inclusionsSérgio Sacani
 
Characteristics of planetary candidates observed by kepler, ii analysis of t...
Characteristics of planetary candidates observed by kepler, ii  analysis of t...Characteristics of planetary candidates observed by kepler, ii  analysis of t...
Characteristics of planetary candidates observed by kepler, ii analysis of t...Sérgio Sacani
 
Titan obliquity as evidence for a subsurface ocean
Titan obliquity as evidence for a subsurface oceanTitan obliquity as evidence for a subsurface ocean
Titan obliquity as evidence for a subsurface oceanSérgio Sacani
 
Herschel far infrared_spectroscopy_of_the_galactic_center
Herschel far infrared_spectroscopy_of_the_galactic_centerHerschel far infrared_spectroscopy_of_the_galactic_center
Herschel far infrared_spectroscopy_of_the_galactic_centerSérgio Sacani
 
Hd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bang
Hd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bangHd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bang
Hd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bangSérgio Sacani
 
A giant planet around a metal poor star of extragalactic origin
A giant planet around a metal poor star of extragalactic originA giant planet around a metal poor star of extragalactic origin
A giant planet around a metal poor star of extragalactic originSérgio Sacani
 
Spatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_b
Spatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_bSpatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_b
Spatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_bSérgio Sacani
 
538352main sts134 presskit_508
538352main sts134 presskit_508538352main sts134 presskit_508
538352main sts134 presskit_508Sérgio Sacani
 
Stellar variability in_open_clusters_new_class_in_ngc3766
Stellar variability in_open_clusters_new_class_in_ngc3766Stellar variability in_open_clusters_new_class_in_ngc3766
Stellar variability in_open_clusters_new_class_in_ngc3766Sérgio Sacani
 
One armed spiral_waves_in_galaxy_simulations_with_computer_rotating_stars
One armed spiral_waves_in_galaxy_simulations_with_computer_rotating_starsOne armed spiral_waves_in_galaxy_simulations_with_computer_rotating_stars
One armed spiral_waves_in_galaxy_simulations_with_computer_rotating_starsSérgio Sacani
 
A population of_fast_radio_bursts_ar_cosmological_distances
A population of_fast_radio_bursts_ar_cosmological_distancesA population of_fast_radio_bursts_ar_cosmological_distances
A population of_fast_radio_bursts_ar_cosmological_distancesSérgio Sacani
 
Herbig haro objects_in_the_vicinity_of_ngc2023
Herbig haro objects_in_the_vicinity_of_ngc2023Herbig haro objects_in_the_vicinity_of_ngc2023
Herbig haro objects_in_the_vicinity_of_ngc2023Sérgio Sacani
 
Dissecting x ray_emitting_gas_around_the_center_of_our_galaxy
Dissecting x ray_emitting_gas_around_the_center_of_our_galaxyDissecting x ray_emitting_gas_around_the_center_of_our_galaxy
Dissecting x ray_emitting_gas_around_the_center_of_our_galaxySérgio Sacani
 

Andere mochten auch (17)

Imprints of fast rotating massive stars in the galactic bulge
Imprints of fast rotating massive stars in the galactic bulgeImprints of fast rotating massive stars in the galactic bulge
Imprints of fast rotating massive stars in the galactic bulge
 
Herschel galactic plane_survey_the_global_distribution_of_ism_gas_componnent
Herschel galactic plane_survey_the_global_distribution_of_ism_gas_componnentHerschel galactic plane_survey_the_global_distribution_of_ism_gas_componnent
Herschel galactic plane_survey_the_global_distribution_of_ism_gas_componnent
 
The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240
The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240
The exceptional soft_x_ray_halo_of_the_galaxy_merger_ngc6240
 
High pre eruptive water contents preserved in lunar melt inclusions
High pre eruptive water contents preserved in lunar melt inclusionsHigh pre eruptive water contents preserved in lunar melt inclusions
High pre eruptive water contents preserved in lunar melt inclusions
 
Characteristics of planetary candidates observed by kepler, ii analysis of t...
Characteristics of planetary candidates observed by kepler, ii  analysis of t...Characteristics of planetary candidates observed by kepler, ii  analysis of t...
Characteristics of planetary candidates observed by kepler, ii analysis of t...
 
Titan obliquity as evidence for a subsurface ocean
Titan obliquity as evidence for a subsurface oceanTitan obliquity as evidence for a subsurface ocean
Titan obliquity as evidence for a subsurface ocean
 
Herschel far infrared_spectroscopy_of_the_galactic_center
Herschel far infrared_spectroscopy_of_the_galactic_centerHerschel far infrared_spectroscopy_of_the_galactic_center
Herschel far infrared_spectroscopy_of_the_galactic_center
 
Hd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bang
Hd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bangHd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bang
Hd140283 a star_in_the_solar_neighborhood_that_formed_shortly_after_the_big_bang
 
A giant planet around a metal poor star of extragalactic origin
A giant planet around a metal poor star of extragalactic originA giant planet around a metal poor star of extragalactic origin
A giant planet around a metal poor star of extragalactic origin
 
Spatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_b
Spatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_bSpatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_b
Spatially resolved images_of_dust_belt_around_the_planet_hosting_subgiant_κ_cr_b
 
538352main sts134 presskit_508
538352main sts134 presskit_508538352main sts134 presskit_508
538352main sts134 presskit_508
 
Stellar variability in_open_clusters_new_class_in_ngc3766
Stellar variability in_open_clusters_new_class_in_ngc3766Stellar variability in_open_clusters_new_class_in_ngc3766
Stellar variability in_open_clusters_new_class_in_ngc3766
 
Bright quasar 3c273
Bright quasar 3c273Bright quasar 3c273
Bright quasar 3c273
 
One armed spiral_waves_in_galaxy_simulations_with_computer_rotating_stars
One armed spiral_waves_in_galaxy_simulations_with_computer_rotating_starsOne armed spiral_waves_in_galaxy_simulations_with_computer_rotating_stars
One armed spiral_waves_in_galaxy_simulations_with_computer_rotating_stars
 
A population of_fast_radio_bursts_ar_cosmological_distances
A population of_fast_radio_bursts_ar_cosmological_distancesA population of_fast_radio_bursts_ar_cosmological_distances
A population of_fast_radio_bursts_ar_cosmological_distances
 
Herbig haro objects_in_the_vicinity_of_ngc2023
Herbig haro objects_in_the_vicinity_of_ngc2023Herbig haro objects_in_the_vicinity_of_ngc2023
Herbig haro objects_in_the_vicinity_of_ngc2023
 
Dissecting x ray_emitting_gas_around_the_center_of_our_galaxy
Dissecting x ray_emitting_gas_around_the_center_of_our_galaxyDissecting x ray_emitting_gas_around_the_center_of_our_galaxy
Dissecting x ray_emitting_gas_around_the_center_of_our_galaxy
 

Ähnlich wie Architecture and dynamics of kepler's candidate multiple transiting planet systems

Evidence for reflected_lightfrom_the_most_eccentric_exoplanet_known
Evidence for reflected_lightfrom_the_most_eccentric_exoplanet_knownEvidence for reflected_lightfrom_the_most_eccentric_exoplanet_known
Evidence for reflected_lightfrom_the_most_eccentric_exoplanet_knownSérgio Sacani
 
Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...
Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...
Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...Sérgio Sacani
 
GrantProposalSethKrantzler_Fra
GrantProposalSethKrantzler_FraGrantProposalSethKrantzler_Fra
GrantProposalSethKrantzler_FraSeth Krantzler
 
Kepler 47 a-transiting_circumbinary
Kepler 47 a-transiting_circumbinaryKepler 47 a-transiting_circumbinary
Kepler 47 a-transiting_circumbinarySérgio Sacani
 
Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...
Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...
Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...Sérgio Sacani
 
A nearby m_star_with_three_transiting_super-earths_discovered_by_k2
A nearby m_star_with_three_transiting_super-earths_discovered_by_k2A nearby m_star_with_three_transiting_super-earths_discovered_by_k2
A nearby m_star_with_three_transiting_super-earths_discovered_by_k2Sérgio Sacani
 
Submillimeter follow up_of_wise_selected_hyperluminous_galaxies
Submillimeter follow up_of_wise_selected_hyperluminous_galaxiesSubmillimeter follow up_of_wise_selected_hyperluminous_galaxies
Submillimeter follow up_of_wise_selected_hyperluminous_galaxiesSérgio Sacani
 
The kepler 10_planetary_system_revisited_by_harps
The kepler 10_planetary_system_revisited_by_harpsThe kepler 10_planetary_system_revisited_by_harps
The kepler 10_planetary_system_revisited_by_harpsSérgio Sacani
 
A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...
A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...
A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...Sérgio Sacani
 
Three super earths_orbiting_hd7924
Three super earths_orbiting_hd7924Three super earths_orbiting_hd7924
Three super earths_orbiting_hd7924Sérgio Sacani
 
The first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wiseThe first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wiseSérgio Sacani
 
The first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wiseThe first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wiseSérgio Sacani
 
Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...
Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...
Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...Sérgio Sacani
 
Alma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxies
Alma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxiesAlma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxies
Alma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxiesSérgio Sacani
 
Measurements of a_massive_galaxy_cluster
Measurements of a_massive_galaxy_clusterMeasurements of a_massive_galaxy_cluster
Measurements of a_massive_galaxy_clusterSérgio Sacani
 
The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...
The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...
The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...Sérgio Sacani
 
A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...
A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...
A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...Sérgio Sacani
 

Ähnlich wie Architecture and dynamics of kepler's candidate multiple transiting planet systems (20)

Batalha n kepler-10b
Batalha n kepler-10bBatalha n kepler-10b
Batalha n kepler-10b
 
Evidence for reflected_lightfrom_the_most_eccentric_exoplanet_known
Evidence for reflected_lightfrom_the_most_eccentric_exoplanet_knownEvidence for reflected_lightfrom_the_most_eccentric_exoplanet_known
Evidence for reflected_lightfrom_the_most_eccentric_exoplanet_known
 
Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...
Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...
Kepler 62 a_five_planet_system_with_planets_of_1_4_and_1_6_earth_radii_in_the...
 
GrantProposalSethKrantzler_Fra
GrantProposalSethKrantzler_FraGrantProposalSethKrantzler_Fra
GrantProposalSethKrantzler_Fra
 
Kepler 47 a-transiting_circumbinary
Kepler 47 a-transiting_circumbinaryKepler 47 a-transiting_circumbinary
Kepler 47 a-transiting_circumbinary
 
Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...
Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...
Four new planets_around_giant_stars_and_the_mass_metallicity_correlation_of_p...
 
A nearby m_star_with_three_transiting_super-earths_discovered_by_k2
A nearby m_star_with_three_transiting_super-earths_discovered_by_k2A nearby m_star_with_three_transiting_super-earths_discovered_by_k2
A nearby m_star_with_three_transiting_super-earths_discovered_by_k2
 
vernstromcv
vernstromcvvernstromcv
vernstromcv
 
Submillimeter follow up_of_wise_selected_hyperluminous_galaxies
Submillimeter follow up_of_wise_selected_hyperluminous_galaxiesSubmillimeter follow up_of_wise_selected_hyperluminous_galaxies
Submillimeter follow up_of_wise_selected_hyperluminous_galaxies
 
The kepler 10_planetary_system_revisited_by_harps
The kepler 10_planetary_system_revisited_by_harpsThe kepler 10_planetary_system_revisited_by_harps
The kepler 10_planetary_system_revisited_by_harps
 
Teletransportacion marcikic2003
Teletransportacion  marcikic2003Teletransportacion  marcikic2003
Teletransportacion marcikic2003
 
A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...
A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...
A Search for Technosignatures Around 11,680 Stars with the Green Bank Telesco...
 
Three super earths_orbiting_hd7924
Three super earths_orbiting_hd7924Three super earths_orbiting_hd7924
Three super earths_orbiting_hd7924
 
The first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wiseThe first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wise
 
The first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wiseThe first hyper_luminous_infrared_galaxy_discovered_by_wise
The first hyper_luminous_infrared_galaxy_discovered_by_wise
 
Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...
Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...
Kepler 1647b the_largest_and_longest_period_kepler_transiting_circumbinary_pl...
 
Alma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxies
Alma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxiesAlma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxies
Alma observations of_sptdiscovered_strongly_lensed_dusty_star_forming_galaxies
 
Measurements of a_massive_galaxy_cluster
Measurements of a_massive_galaxy_clusterMeasurements of a_massive_galaxy_cluster
Measurements of a_massive_galaxy_cluster
 
The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...
The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...
The discovery of_lensed_radio_and_x-ray_sources_behind_the_frontier_fields_cl...
 
A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...
A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...
A SPectroscopic Survey of Biased Halos in the Reionization Era (ASPIRE): JWST...
 

Mehr von Sérgio Sacani

Biogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune Waterworlds
Biogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune WaterworldsBiogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune Waterworlds
Biogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune WaterworldsSérgio Sacani
 
Asymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 b
Asymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 bAsymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 b
Asymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 bSérgio Sacani
 
Formation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disksFormation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disksSérgio Sacani
 
Nightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43b
Nightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43bNightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43b
Nightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43bSérgio Sacani
 
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...Sérgio Sacani
 
Disentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOSTDisentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOSTSérgio Sacani
 
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...Sérgio Sacani
 
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...Sérgio Sacani
 
Isotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on IoIsotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on IoSérgio Sacani
 
Hubble Asteroid Hunter III. Physical properties of newly found asteroids
Hubble Asteroid Hunter III. Physical properties of newly found asteroidsHubble Asteroid Hunter III. Physical properties of newly found asteroids
Hubble Asteroid Hunter III. Physical properties of newly found asteroidsSérgio Sacani
 
Observational constraints on mergers creating magnetism in massive stars
Observational constraints on mergers creating magnetism in massive starsObservational constraints on mergers creating magnetism in massive stars
Observational constraints on mergers creating magnetism in massive starsSérgio Sacani
 
Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...
Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...
Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...Sérgio Sacani
 
The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...
The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...
The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...Sérgio Sacani
 
Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...
Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...
Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...Sérgio Sacani
 
First Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPR
First Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPRFirst Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPR
First Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPRSérgio Sacani
 
The Sun’s differential rotation is controlled by high- latitude baroclinicall...
The Sun’s differential rotation is controlled by high- latitude baroclinicall...The Sun’s differential rotation is controlled by high- latitude baroclinicall...
The Sun’s differential rotation is controlled by high- latitude baroclinicall...Sérgio Sacani
 
Hydrogen Column Density Variability in a Sample of Local Compton-Thin AGN
Hydrogen Column Density Variability in a Sample of Local Compton-Thin AGNHydrogen Column Density Variability in a Sample of Local Compton-Thin AGN
Hydrogen Column Density Variability in a Sample of Local Compton-Thin AGNSérgio Sacani
 
Huygens - Exploring Titan A Mysterious World
Huygens - Exploring Titan A Mysterious WorldHuygens - Exploring Titan A Mysterious World
Huygens - Exploring Titan A Mysterious WorldSérgio Sacani
 
The Radcliffe Wave Of Milk Way is oscillating
The Radcliffe Wave Of Milk Way  is oscillatingThe Radcliffe Wave Of Milk Way  is oscillating
The Radcliffe Wave Of Milk Way is oscillatingSérgio Sacani
 
Thermonuclear explosions on neutron stars reveal the speed of their jets
Thermonuclear explosions on neutron stars reveal the speed of their jetsThermonuclear explosions on neutron stars reveal the speed of their jets
Thermonuclear explosions on neutron stars reveal the speed of their jetsSérgio Sacani
 

Mehr von Sérgio Sacani (20)

Biogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune Waterworlds
Biogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune WaterworldsBiogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune Waterworlds
Biogenic Sulfur Gases as Biosignatures on Temperate Sub-Neptune Waterworlds
 
Asymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 b
Asymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 bAsymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 b
Asymmetry in the atmosphere of the ultra-hot Jupiter WASP-76 b
 
Formation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disksFormation of low mass protostars and their circumstellar disks
Formation of low mass protostars and their circumstellar disks
 
Nightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43b
Nightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43bNightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43b
Nightside clouds and disequilibrium chemistry on the hot Jupiter WASP-43b
 
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
Discovery of an Accretion Streamer and a Slow Wide-angle Outflow around FUOri...
 
Disentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOSTDisentangling the origin of chemical differences using GHOST
Disentangling the origin of chemical differences using GHOST
 
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
All-domain Anomaly Resolution Office U.S. Department of Defense (U) Case: “Eg...
 
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
PossibleEoarcheanRecordsoftheGeomagneticFieldPreservedintheIsuaSupracrustalBe...
 
Isotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on IoIsotopic evidence of long-lived volcanism on Io
Isotopic evidence of long-lived volcanism on Io
 
Hubble Asteroid Hunter III. Physical properties of newly found asteroids
Hubble Asteroid Hunter III. Physical properties of newly found asteroidsHubble Asteroid Hunter III. Physical properties of newly found asteroids
Hubble Asteroid Hunter III. Physical properties of newly found asteroids
 
Observational constraints on mergers creating magnetism in massive stars
Observational constraints on mergers creating magnetism in massive starsObservational constraints on mergers creating magnetism in massive stars
Observational constraints on mergers creating magnetism in massive stars
 
Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...
Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...
Observation of Gravitational Waves from the Coalescence of a 2.5–4.5 M⊙ Compa...
 
The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...
The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...
The SAMI Galaxy Sur v ey: galaxy spin is more strongly correlated with stella...
 
Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...
Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...
Is Betelgeuse Really Rotating? Synthetic ALMA Observations of Large-scale Con...
 
First Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPR
First Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPRFirst Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPR
First Direct Imaging of a Kelvin–Helmholtz Instability by PSP/WISPR
 
The Sun’s differential rotation is controlled by high- latitude baroclinicall...
The Sun’s differential rotation is controlled by high- latitude baroclinicall...The Sun’s differential rotation is controlled by high- latitude baroclinicall...
The Sun’s differential rotation is controlled by high- latitude baroclinicall...
 
Hydrogen Column Density Variability in a Sample of Local Compton-Thin AGN
Hydrogen Column Density Variability in a Sample of Local Compton-Thin AGNHydrogen Column Density Variability in a Sample of Local Compton-Thin AGN
Hydrogen Column Density Variability in a Sample of Local Compton-Thin AGN
 
Huygens - Exploring Titan A Mysterious World
Huygens - Exploring Titan A Mysterious WorldHuygens - Exploring Titan A Mysterious World
Huygens - Exploring Titan A Mysterious World
 
The Radcliffe Wave Of Milk Way is oscillating
The Radcliffe Wave Of Milk Way  is oscillatingThe Radcliffe Wave Of Milk Way  is oscillating
The Radcliffe Wave Of Milk Way is oscillating
 
Thermonuclear explosions on neutron stars reveal the speed of their jets
Thermonuclear explosions on neutron stars reveal the speed of their jetsThermonuclear explosions on neutron stars reveal the speed of their jets
Thermonuclear explosions on neutron stars reveal the speed of their jets
 

Architecture and dynamics of kepler's candidate multiple transiting planet systems

  • 1. Preliminary draft. To be submitted to ApJ. Architecture and Dynamics of Kepler’s Candidate Multiple Transiting Planet Systems Jack J. Lissauer1 , Darin Ragozzine2 , Daniel C. Fabrycky3,4 , Natalie M. Batalha5 , William J. arXiv:1102.0543v1 [astro-ph.EP] 2 Feb 2011 Borucki1 , Stephen T. Bryson1 , Douglas A. Caldwell6 , Edward W. Dunham13 , Eric B. Ford7 , Jonathan J. Fortney3 , Thomas N. Gautier III8 , Matthew J. Holman2 , Jon M. Jenkins1,6 , David G. Koch1 , David W. Latham3 , Geoffrey W. Marcy9 , Robert Morehead7 , Jason Rowe6 , Elisa V. Quintana6 , Dimitar Sasselov2 , Avi Shporer10,11 , Jason H. Steffen12 Jack.Lissauer@nasa.gov ABSTRACT Borucki et al. (2011a) [ApJ submitted, hereafter B11] report on characteristics of over 1200 candidate transiting planets orbiting nearly 1000 Kepler spacecraft target stars detected in the first four months of spacecraft data. Included among these tar- gets are 115 targets with two transiting planet candidates, 45 targets with three, 8 with four, and one each with five and six sets of transit signatures. We character- ize herein the dynamical properties of these candidate multi-planet systems. We find that virtually all systems are stable, as tested by numerical integration assuming a mass-radius relationship. The distribution of observed period ratios is also clustered 1 NASA Ames Research Center, Moffett Field, CA, 94035, USA 2 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA 3 Department of Astronomy & Astrophysics, University of California, Santa Cruz, CA 95064, USA 4 Hubble Fellow 5 Department of Physics and Astronomy, San Jose State University, San Jose, CA 95192, USA 6 SETI Institute/NASA Ames Research Center, Moffett Field, CA, 94035, USA 7 University of Florida, 211 Bryant Space Science Center, Gainesville, FL 32611, USA 8 Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA 9 Astronomy Department, UC Berkeley, Berkeley, CA 94720 10 Las Cumbres Observatory Global Telescope Network, 6740 Cortona Drive, Suite 102, Santa Barbara, CA 93117, USA 11 Department of Physics, Broida Hall, University of California, Santa Barbara, CA 93106, USA 12 Fermilab Center for Particle Astrophysics, P.O. Box 500, MS 127, Batavia, IL 60510 13 Lowell Observatory, 1400 W. Mars Hill Road, Flagstaff, AZ 86001, USA
  • 2. –2– just outside resonances, particularly the 2:1 resonance. Neither of these characteristics would emerge if the systems were significantly contaminated with false positives, and these combined with other considerations strongly suggest that the majority of these multi-candidate systems are true planetary systems. Using the observed multiplicity frequencies (e.g., the ratio of three-candidate systems to two-candidate systems), we find that a single population that matches the higher multiplicities underpredicts the number of singly-transiting systems. We provide constraints on the true multiplicity and mutual inclination distribution of the multi-candidate systems, which includes a population of systems with multiple small planets with low ( 5◦ ) inclinations. In all, multi-transiting systems from Kepler provide a large and rich dataset which can be used to powerfully test theoretical predictions of the formation and evolution of plan- etary systems. [This is a preliminary draft; some numbers may change slightly in the submitted version.] Subject headings: celestial mechanics – planetary systems 1. Introduction Kepler is a 0.95 m aperture space telescope using transit photometry to determine the fre- quency and characteristics of planets and planetary systems (Borucki et al. 2010; Koch et al. 2010; Jenkins et al. 2010; Caldwell et al. 2010). Besides searching for small planets in the habitable zone, Kepler ’s ultra-precise long-duration photometry is also ideal for detecting systems with multiple transiting planets. Borucki et al. (2011b) presented, for the first time, 5 Kepler systems with multiple transiting planet candidates, and these candidate systems were analyzed by Steffen et al. (2010). A system of three planets (Kepler-9=KOI-377, Holman et al. 2010; Torres et al. 2011) was found, in which a resonant effect on the transit times allowed Kepler data to confirm two of the planets and characterize the dynamics of the system. A compact system of six transiting planets (Kepler-11, Lissauer et al. 2011), five of which were confirmed by their mutual gravitational inter- actions exhibited through transit timing variations (TTVs), has also been reported. These systems were highly anticipated because of their unique value in understanding the formation and evolution of planetary systems Koch et al. (1996); Agol et al. (2005); Holman & Murray (2005); Fabrycky (2009); Holman (2010); Ragozzine & Holman (2010). A catalog of all candidate planets evident in the first four and a half months of Kepler data is presented in B11. They also summarize the observational data and follow-up programs being conducted to confirm and characterize planetary candidates. Each object in this catalog is assigned a vetting flag (reliability score), with 1 signifying a confirmed or validated planet (>98% confidence), 2 being a strong candidate (> 80% confidence), 3 being a somewhat weaker candidate (> 60% confidence), and 4 signifying a candidate that has not received ground-based follow-up (this sample also has an expected reliability of >60%). For most of the remainder of the paper, we refer to all
  • 3. –3– of these objects as “planets”, although 99% remain unvalidated and thus should more properly be termed planet candidates. Ragozzine & Holman (2010) and Lissauer et al. (2011) point out that systems with multiple planet candidates are less likely to be due to astrophysical false positives than stars with single candidates, although the possibility that one of the candidate signals is due to, e.g., a background eclipsing binary, is still non-negligible. Here we only examine the candidates that are known to be in multiple systems; many of the apparently single systems are likely to have additional planets that escape detection by being too small, too long-period, or too highly inclined with respect to the line of sight. Ford et al. (2011, in preparation) discuss candidates exhibiting TTVs that may be caused by non-transiting planets. B11 list five dozen planetary candidates with radii Rp > 15 R⊕ (Earth radii), but none of these are in multiple candidate systems. Latham et al. (2011, in preparation) compare the radius distribution of planets in multi-planet systems with those of planets which don’t have observed transiting companions. We examine herein the dynamical aspects of Kepler’s multi-planet systems. Ragozzine & Holman (2010) discussed in detail the value of systems with multiple transiting planets. Besides their higher likelihood of being true planetary systems, multi-transiting systems provide numerous other insights that are difficult or impossible to gain from single-transiting sys- tems. These systems harness the power of radius measurements from transit photometry in combi- nation with the illuminating properties of multi-planet orbital architecture. Observable interactions between planets in these systems, seen in transit timing variations (TTVs), are much easier to char- acterize, and our knowledge of both stellar and planetary parameters is improved. The true mutual inclinations between planets in these systems, while not directly observable from the light curves themselves, are much easier to obtain than in other systems, both individually and statistically (see Section 6 and Lissauer et al. 2011). The distributions of planets in these systems, both in terms of physical and orbital parameters, are invaluable for comparative planetology and provide unique and invaluable insights into the processes of planet formation and evolution. It is thus very exciting that B11 report 115 doubles, 45 triples, 8 quads, 1 quintuple and 1 sextuple transiting system. While very few of these systems contain confirmed planets, it is very likely that this set of multi-transiting candidates contains more than twice as many as planets as those known in the pre-Kepler era. We begin by summarizing the characteristics of Kepler’s multi-planet systems in Section 2. The reliability of the data set is discussed in Section 3. Sections 4 – 6 present our primary statistical results: Section 4 presents an analysis of the long-term dynamical stability of candidate planetary systems for nominal estimates of planetary masses and unobserved orbital parameters. Evidence of orbital commensurabilities in among planet candidates is analyzed in Section 5. The characteristic distributions of numbers of planets per system and mutual inclinations of planetary orbits are discussed in Section 6. We compare the Kepler-11 planetary system (Lissauer et al. 2011) to other Kepler candidates in Section 7. We conclude with a discussion for the implications of these data and summarize the most important results presented herein.
  • 4. –4– 2. Characteristics of Kepler Multi-Planet Systems The light curves derived from Kepler data can be used to measure the radii and orbital periods of planets. Periods are measured to high accuracy. The ratios of planetary radii to stellar radii are also measured quite precisely, except for low signal to noise transits (e.g., planets much smaller than their stars, those for which few transits are observed, and planets orbiting faint or variable stars). Besides errors in estimated stellar radii, in some cases the planet’s radius is underestimated because a significant fraction of the flux in the target aperture is provided by a star other than the one being transited by the planet. This extra light can often be determined and corrected for by analyzing displacements of the centroid of light from the target between in-transit and out-of-transit observations (Bryson et al. 2010), high-resolution imaging and/or spectroscopic studies conducted from the ground , although dilution by faint physically bound stars is difficult to detect (Torres et al. 2011). Spectroscopic studies can and in the future will be used to estimate the stellar radii, thereby allowing for improved estimates of the planetary radii, and also the stellar masses. We list the key observed properties of Kepler two-planet systems in Table 1, three-planet systems in Table 2, four-planet systems in Table 3, and those systems with more than four planets in Table 4. Planet indexes given in these tables signify orbit order, with 1 being the planet with the shortest period. These do not always correspond to the post-decimal point portion (.01, .02, etc.) of the KOI number designation of these candidates given in B11, for which the numbers signify the order in which the candidates were identified. These tables are laid out differently from one another because of the difference in parameters that are important for systems with differing numbers of planets. In addition to directly observed properties, these tables contain results of the dynamical analyses discussed below. We plot the period and radius of each active Kepler planetary candidate observed during the first four months of spacecraft operations, with special emphasis on multi-planet systems, in Figure 1. The ratio(s) of orbital periods is a key factor in planetary system dynamics; the distribution of these period ratios is plotted in Figure 3. Figure 4 shows a cumulative distribution of the period ratio. To check whether the observed transiting planets are randomly distributed among host stars we compared the observed distribution of planets per system to a statistically random Poisson distribution, assuming there are 1229 planets orbiting some of the 150,000 stars Kepler monitors, i.e., a Poisson distribution with a mean of 1229/150,000≈ 0.008. Such a Poisson distribution gives an expectation of 1219 single planet systems, 5.0 doubles, 0.014 triples, 2.8 × 10−5 systems with four planets and above. For single planet systems there are fewer observed systems than expected from a random distribution, but, for systems with two and more planets there are far more systems detected by Kepler than predicted randomly. This shows that transiting planets tend to “come in packs” or cluster around stars, meaning that once a planet is detected around a given star, that star is more likely to be orbited by an additional planet than a star with no detected planets. Some of the observed clustering is attributable to SNR considerations, with stellar brightness, size, intrinsic
  • 5. –5– variability, multiplicity, crowdedness of field, and position on the Kepler detectors all contributing. Quantifying the effects of SNR variations is beyond the scope of this paper, but they are clearly far too small to account for the bulk of the observed clustering. Given the bias of the transit method to detect planets on edge-on orbits, a possible interpretation of the clustering of transiting planets is that planetary systems (or at least some fraction of them) tend to be coplanar, like the Solar System. It is difficult to estimate the corresponding distribution for planets detected through radial velocity (RV) surveys, since we do not know how many stars were spectroscopically surveyed for planets for which none were found. However, a search through the Exoplanet Orbit Database1 (Wright et al. 2010) shows that 17% of the planetary systems detected by RV surveys are multi- planet systems (have at least two planets), compared to 21% for Kepler. Including a linear RV trend as evidence for an additional planetary companion, Wright et al. (2009) show that at least 28% of known RV planetary systems are multiple. Therefore, the clustering observed for transiting planets detected by Kepler is likely to occur also for RV planets, and maybe even at a higher fraction, which is expected given the bias of the transit detection method. 3. Reliability of the Sample Very few of the planetary candidates presented herein have been validated as true exoplan- ets. As discussed in B11, the overwhelming majority of false positives are expected to come from eclipsing binary stars. Some of these will be grazing eclipses of the target star, others eclipses of fainter stars (physically associated to the target or background/faint foreground objects) within the target aperture. Similarly, transits of fainter stars by large planets may also mimic small planets transiting the target star. The vast majority of all active Kepler planet candidates were identified using automated programs that do not discriminate between single candidates and systems of mul- tiple planets. The 1% of single and multiple candidate systems that were identified in subjective visual searches are too small to affect our results in any statistically significant manner, and even the signs of such changes are unclear. The fact that planetary candidates are clustered around targets in the sense that there are far more targets with more than one candidate than would be expected for randomly distributed can- didates (section 2) suggests that the reliability of the multi-candidate systems is likely to be higher than that of the single candidate targets. We know from radial velocity observations that planets frequently come in multiple systems (Wright et al. 2009), whereas astrophysical false positives are nearly random. Nonetheless, the presence of multiple candidates increases our confidence that one or both are planets is real since the probability of two false positive signals is the product of the probabilities 1 exoplanet.org, as of January 2011.
  • 6. –6– of two relatively rare cases, significantly lowering the overall probability than none of the signals is due to planets Ragozzine & Holman (2010). (Triple star systems would be dynamically unstable for most of the orbital period ratios observed. Their dynamics would also give rise to very large eclipse timing variations — e.g., Carter et al. 2011 — which are generally not seen.) Additionally, the concentration of candidate pairings with period ratio near 1.5 and 2 (fig. 3) suggest that these subsamples likely have an even larger fraction of true planets, since such concentrations would not be seen for random eclipsing binaries. However, at present these qualitative factors have not been quantified. Strong TTV signals are no more common for planets in observed multi-transiting candidates than for single candidates (Ford et al. 2011, in preparation), suggesting that large TTVs are often the result of stellar systems masquerading as planetary systems (Fabrycky et al. 2011, in preparation, Steffen et al., 2011, in preparation). Nonetheless, it is possible that weaker TTVs correlate with multi-transiting systems, but these require more data and very careful study to reveal (e.g., those in the Kepler-11 planetary system, Lissauer et al. 2011). The ratio of durations given the observed periods of candidates in multi-transiting systems can be used to potentially identify false positives (generated by a blend of two objects eclipsing two stars of different densities) or as a probe of the eccentricity and inclination distributions. An analysis of the corrected duration ratio as in (?) shows that essentially all pairs of candidate have the expected duration ratio if they were both orbiting the same star. The consistency of duration ratios is not a strong constraint, but the observed systems also pass this test. 4. Long-term Stability of Planetary Systems Kepler measures planetary sizes, whereas masses are the key parameter for dynamical studies. We convert sizes to masses using the following simple formula: 2.06 Mp = Rp (1) where the planet’s mass, Mp , is expressed in Earth masses and its radius, Rp , in units of Earth’s mean radius. The power-law of equation 1 is fit to Earth and Saturn; it slightly overestimates the mass of Uranus (17.2 M⊕ vs. 14.5 M⊕ ) and slightly underestimates the mass of Neptune (16.2 M⊕ vs. 17.1 M⊕ ). (Note that Uranus is larger, but Neptune is more massive, so fitting both with a monotonic mass-radius relation is not possible.) A convenient metric for the dynamical proximity of planets is the separation of their orbital semi-major axes, ai+1 − ai , measured in units of their mutual Hill sphere radius mi + mi+1 1/3 (ai + ai+1 ) RHi,i+1 = , (2) 3M⋆ 2 where the index i = 1, N −1, with N being the number of planets in the system under consideration, mi and ai are the planetary masses and semi-major axes, and M⋆ is the central star’s mass.
  • 7. –7– The dynamics of two-planet systems are a special case of the 3-body problem that is amenable to analytic treatment and simple numerically-derived scaling formulas. For instance, a pair of planets initially on coplanar circular orbits with orbital separation √ ∆ ≡ (ao − ai )/RH > 2 3 ≈ 3.46 (3) cannot develop crossing orbits later, and are thus called “Hill stable” (Gladman 1993). Orbital separations of all planet pairs in two planet systems for stellar masses given in B11 and planetary masses given by Equation 1 are listed in Table 1. We see that they all obey this criterion. The actual dynamical stability of these systems also depends on their eccentricity and mutual inclination (Veras & Armitage 2004), neither of which can be measured well from the transit data alone. Circular coplanar orbits have the lowest angular momentum deficit (AMD), and thus are the most stable, except for protective resonances that produce libration of angles and prevent close approaches; a second exception exists for retrograde orbits (Smith & Lissauer 2009), but these are implausible on cosmogonic grounds (Lissauer & Slartibartfast 2008). Hence, the Kepler data can only be used to suggest that a pair of planets may be potentially unstable, particularly if their semi-major axes are closer than the resonance overlap limit (Wisdom 1980; Malhotra 1998): aouter − ainner Mp 2/7 2 < 1.5 . (4) ainner + aouter M⋆ (This is derived only for one massive planet; the other is massless.) Conversely, if it is assumed that both candidates are true planets, limits on the eccentricities and mutual inclinations can be inferred from the requirement that the planetary systems are presumably stable for the age of the system, which is generally of order 1011 times longer than the orbital timescale for the short-period planets that represent the bulk of Kepler ’s candidates. However, we may turn the observational problem around: the ensemble of duration measurements can be used to address the eccentricity distribution (Ford et al. 2008); since multiplanet systems have eccentricities constrained by stability requirements, they can be used as a check to those results (Moorhead et al. 2011, in preparation). Systems with more than two planets have additional dynamically complexity, but are very unlikely to be stable unless all neighboring pairs of planets satisfy equation 3. Nominal separations between neighboring planet pairs in three planet systems are listed in Table 2, the separations in four planet systems are listed in Table 3, and those for the 5 and 6 planet systems in Table 4. We find that in these cases, equation 3 is satisfied by quite a wide factor. From suites of numerical integrations Smith & Lissauer (2009), dynamical survival of systems with more than three comparably-spaced planets for at least 1010 orbits has been shown only if the relative spacing between orbital semi- major axes exceeds a critical number (∆crit 9) of mutual Hill-spheres. For each adjacent set of 3 planets in our sample, we plot the ∆ separations of both the inner and the outer pairs in Figure 9. We find that in some cases ∆ < ∆crit , but the other separation is quite a bit bigger than ∆crit . In our simulations of a population of systems in section 6, we impose the stability boundary ∆max + ∆min > 18, which accounts for this possibility, in addition to requiring that inequality 3 is satisfied.
  • 8. –8– We investigated long-term stability of all 55 systems with three or more planets using the hybrid integrator within the Mercury package (Chambers 1999), run on the supercomputer Pleiades at University of California, Santa Cruz. We set the switchover at 3 Hill radii, but in practice we aborted simulations that violated this limit, so for the bulk of the simulation the Wisdom-Holman n-body mapping (Wisdom & Holman 1991) was used, with a time step of 0.05 times as large as the orbital period of the innermost planet. The simplest implementation (Nobili & Roxburgh 1986) of general relativistic precession was used, an additional potential UGR = −3(GM⋆ /cr)2 , where G is Newton’s constant, c is the speed of light, and r is the instantaneous distance from the star. More sophisticated treatments of general relativity (Saha & Tremaine 1994) are not yet required, due to the uncertainties of the masses of the planets and stars whose dynamics are being modeled. We neglected precession due to tides or rotational flattening, which are only significant compared to general relativity for Jupiter-size planets in few-day orbits (Ragozzine & Wolf 2009). The stellar mass was derived from the R⋆ and log g estimates from color photometry, as given by B11. We assumed initially circular and coplanar orbits that matched the observed periods and phases, and chose the nominal masses of equation 1. For the triples and quadruples, we ran these integrations for 4 × 109 orbital periods of the innermost planet, and found the nominal system to be stable for this span in nearly all cases; the three exceptions are described below. An integration of the five-planet system KOI-500 has run 2 × 109 days (the inner planet’s period is 0.99 day), suggesting stability for astronomically rele- vant timescales is possible. However, the short orbital periods and high planet number means this conclusion needs to be revisited in future work, especially once masses are estimated observation- ally. (We consider the near-resonances of KOI-500 in section 5.5.) The most populous transiting planetary system discovered so far is the six-planet system Kepler-11. In the discovery paper, Lissauer et al. (2011) reported a circular model, with transit time variations yielding an estimate of the masses of the five inner planets. They also proposed two more models with planets on mod- erate eccentricities, and slightly different masses, which also fit the transit timing signals. Now, both those eccentric solutions have become unstable, at 610 Myr for the all-eccentric fit, and at 427 Myr for an integration that was a very slight offset from the b/c-eccentric fit. The all-circular fit remains stable at 640 Myr, so we conclude that the fits of that paper are physically plausible, but that introducing a small amount of eccentricity can compromise stability. Future work should address how high the eccentricities could be, while requiring stability for several Gyr. Only three systems became unstable at the nominal masses, KOI-284, KOI-191, and KOI-730; the first is described next, the latter two are described in section 5. KOI-284 has a pair of candidates with a period ratio 1.0383. Both planets would need to have masses about that of the Earth or smaller for the system to satisfy equation 3 and thus be Hill stable. But the vetting flags are 3 for both these candidates — meaning that they are not good candidates anyway — and the data display significant correlated noise which may be responsible for these detections. We expect one or both of these candidates are not planets, or if they are, they are not orbiting the same star. In our investigation, we also were able to correct a poorly fit radius of 3.2 RJup (accompanied
  • 9. –9– by a grazing impact parameter in the original model, trying to account for a “V” shape) for KOI- 1426, which resulted in too large a nominal mass, driving the system unstable. It is remarkable that stability considerations allowed us to identify a poorly conditioned light curve fit. Once the correction was made to 1.2 RJup , the nominal mass allowed the system to be stable. That so many of the systems passed basic stability measurements suggests that there are not a large number of planets that with substantially higher densities than given by the simple formula of equation (1). If these candidates are assumed to be orbiting the same stars, stability constraints could potentially be used to place upper limits on the masses and densities. 5. Resonances 5.1. Resonance Abundance Analysis The ratio of periods of multi-transiting candidates in the same system is significantly different from ratios obtained by the ratio of randomly selected periods from all 408 multi-transiting can- didates. Multiply-transiting planets provide better data for the measurement of resonances than RV planetary systems for two reasons. First, periods are measured to very high accuracy, even for systems with low SNR, so period ratios can be confidently determined without the harmonic ambiguities that affect RV detections (Anglada-Escud´ et al. 2010; Dawson & Fabrycky 2010). Sec- e ond, resonances significantly enhance the signal of transit timing variations (Agol et al. 2005; Holman & Murray 2005; Veras & Ford 2010), which can also be measured accurately, either to demonstrate resonance occupation (as will be possible with the Kepler-9 system with future data; Holman et al. 2010 or exclude it with confidence. Planets can be librating in a resonance even when they have apparent periods that are not perfectly commensurate (Ragozzine & Holman 2010). Nevertheless, the period ratios of small tran- siting resonant planets will generally be within a few percent of commensurability, unless masses or eccentricities are relatively high. In the cumulative distribution of period ratios (Figure 3), we see that few planets appear to be directly in the resonance (with the important exception of KOI-730). In the solar system, there is a known excess of satellite pairs near mean-motion resonance (MMR) (Goldreich 1965) and theoretical models of planet formation and migration suggest that this should be the case (e.g., Terquem & Papaloizou 2007). We can ask whether such an excess can be seen in the sample of multiply transiting Kepler planetary systems. To do this, we divide the period ratios of the multiply transiting systems into bins which surround the first and second-order (j:j − 1 and j:j − 2) MMRs. The boundaries between these “neighborhoods” are chosen at the intermediate, third-order MMRs. Thus, all candidate systems with period ratios less than 4:1 are in some neighborhood. For example, the neighborhood of the 3:1 MMR runs between period ratios of 5:2 and 4:1, for the 2:1 it runs from 7:4 to 5:2. With this method the first-order MMRs have neighborhoods that are essentially twice as large as those for second-order MMRs.
  • 10. – 10 – We define a statistic, ξ, that is a measure of the difference between an observed period ratio and the MMR in its neighborhood. In order to treat each neighborhood uniformly, they are scaled such that the ξ statistic runs from −1 to 1. For first-order MMRs, ξ is given by 1 1 ξ1 ≡ 3 − Round (5) P −1 P−1 where P is the observed period ratio (always greater than unity) and “Round” is the standard rounding function that returns the integer nearest to its argument. Similarly, for second-order MMRs, ξ is given by 2 2 ξ2 ≡ 3 − Round . (6) P −1 P−1 Table 5 gives the number of period ratios taken from tables 1–4 that are found in each neighborhood. The ξ statistic as a function of the period ratio, along with the data, is shown in Figure 6. We calculate ξ for each planetary period ratio and stack the results for all of the resonances in each MMR order. The resulting distribution in ξ is compared to a randomly-generated sample drawn from a uniform distribution in log P , a uniform distribution in P , and from a sample con- structed by taking the ratios of a random set of the observed periods in the multiple candidate systems. This third sample has a distribution that is very similar to the logarithmic distribution. Figure 7 shows a histogram of the number of systems as a function of ξ for both the first and second-order MMRs. We see that the results for the second-order MMRs have some qualitative similarities to the first-order MMR, though the peaks and valleys are less prominent. In our tests below we also consider the case where all first and second-order MMRs are stacked together. To test whether the observed distributions in ξ are consistent with those from our various test samples, we took the absolute value of ξ for each collection of neighborhoods and used the Kolmogorov-Smirnov test (KS test) to determine the probability that the observed sample is drawn from the same distribution as our test sample. The results of these tests are shown in Table 6. Figure 8 shows the probability density function (PDF) for the combined distribution of first and second-order MMRs and the PDF for the logarithmically distributed sample. A few notable results of these tests include: 1) the combined first and second-order neighbor- hoods are distinctly inconsistent with any of the trial period distributions, 2) the neighborhoods surrounding the first-order MMRs are also very unlikely to originate from the test distributions, 3) there is significant evidence that the neighborhood surrounding the 2:1 MMR alone is inconsistent with the test distributions, and 4) there is a hint that second-order resonances are not be consistent with the test distributions, particularly the distribution that is uniform in P . Additional data are necessary to produce higher significance results for the second-order resonances. Given the distinct differences between our test distributions and the observed distribution, a careful look at the histograms shown in Figure 7 shows that the most common location for a pair of planets to reside is slightly exterior to the MMR (the planets are farther apart than the resonance location) regardless of whether the resonance is first or second-order. Also, a slight
  • 11. – 11 – majority (just over 60%) of the period ratios have ξ statistics between -1/2 and +1/2 while all of the test distributions have roughly 50% between these two values in ξ (note that ξ = ±1/2 corresponds to sixth-order resonances in first-order neighborhoods and twelfth-order resonances in second-order neighborhoods—none of which are likely to be strong in these planet pairs though they may play some minor role). There are very few examples of systems that lie slightly closer to each other than the first-order resonances and for the second-order resonances only the 3:1 MMR has systems just interior to it. Finally, there is a hint that the planets near the 2:1 MMR have a range of orbital periods, while those near the 3:2 appear to have shorter periods and smaller radii on average. While additional data are necessary to claim these statements with high confidence, the observations are nonetheless interesting and merit some investigation to determine the mechanisms that might produce such distributions. A preference for resonant period ratios cannot be exhibited by systems without direct dynam- ical interaction. Thus, the preference for periods near resonances in our dataset is statistically significant evidence that many of the systems are actually systems of three or more planets. (Stars with such close orbits are not dynamically stable, thus dynamical interaction between objects with period ratios 5 indicates low masses due to planets.) Assessing the probability of dynamical proximity to resonance for the purpose of eliminating the false positive hypothesis in any partic- ular system requires a specific investigation; here, we can only say that statistically many of the near-resonant systems appear to be planetary in nature. The excess of pairs of planets wider than nominal resonances was predicted by (Terquem & Papaloizou 2007), who pointed out that tidal interactions with the star will, under some circumstances, break resonances and will more strongly affect the inner planets. Interestingly KOI-730 seems to have maintained stability and resonance occupation despite differential tidal evolution. 5.2. Frequency of Resonant Systems Doppler planet searches suggest that roughly one third of multiple planet systems well-characterized by Doppler observations are near a low-order period commensurability with one sixth near the 2:1 MMR (Wright et al. 2011). However, an approximate degeneracy with a single planet on an eccen- tric orbit makes it difficult for Doppler observations to detect a low-mass planet in an interior 2:1 MMR (Giuppone et al. 2009; Anglada-Escud´ et al. 2010). Among Doppler-discovered systems, e only one quarter of the pairs near the 2:1 MMR have an inner planet significantly less massive than the outer planet. The two exceptions (GJ 876 b&c, µ Ara e&b) benefitted from an unusually large (≥ 100) number of Doppler observations. Thus, it is possible that the true rate of planetary systems near the 2:1 MMR is significantly greater than one sixth. Kepler observations find that at least ∼ 16% of multiple transiting planet candidate systems contain at least one pair of transiting planets close to a 2:1 period commensurability (1.83 ≤ Pout /Pin ≤ 2.18). Even using our narrow range of period ratios, the true rate of systems near a 2:1 period
  • 12. – 12 – commensurability could be significantly greater than ∼ 16%, since not all planets will transit. If we assume that planets near the 2:1 MMR are in the low inclination regime, then the outer planet should transit ≃ 63% of the time, implying that the true rate of detectable planets near the 2:1 MMR (if both had a favorable inclination) is at least ∼ 25%. If a significant fraction of these systems are not in the low inclination regime, then the true rate of pairs of planets near the 2:1 MMR would be even larger. This mildly suggests that radial velocity surveys could be missing some 2:1 pairs due to observational degeneracy. To consider this possibility further requires investigating the expected mass ratios of planets near the 2:1. Based on the B11 catalog, neighboring transiting planet candidates tend to have similar radii, and in the majority of cases, the outer planet is slightly larger. This is consistent with the typical outcome of planet formation models that predict more distant planets can more easily grow to larger sizes. The distribution of planetary radii ratios (Rp,out /Rp,in ) for neighboring pairs of transiting planet candidates near a 2:1 period commensurability is concentrated between 0.9 and 1.25 with a tail extending to larger ratios (see Fig. 10, solid curve). For reasonable assumptions of a mass- radius ratio (Eqn. 1), this implies that most neighboring transiting planet candidate systems have masses within 20% of each other. Fortunately, such pairs are not significantly affected by the radial velocity observational degeneracy with a single eccentric planet. However, the tail of ∼ 20% of systems with a radii ratio exceeding 1.7 are more problematic, as the Doppler signatures of would differ from that of a single eccentric planet by less than 20% of the velocity amplitude of the outer planet. Based on the planet radii ratio distribution from Kepler, we estimate that the abundance of neighboring planets near the 2:1 MMR could be ∼ 20% greater than suggested by Doppler surveys due to the difficulty in distinguishing the Doppler signature of such systems from a single eccentric planet. A more precise determination of the frequency of resonant or near-resonant systems based on Kepler and/or Doppler surveys is left for future work. 5.3. KOI-730: A Multi-Resonant Candidate System While few nearly exact mean motion resonances are evident in the sample of Kepler planetary candidates, one system stands out as exceptional: the periods of the four candidates in KOI- 730 satisfy the ratio 6:4:4:3 to ∼ 1 part in 1000 or better. This system is the first to show evidence for extrasolar “Trojan” or co-orbital planets, which have been suggested to be theoretically possible (Laughlin & Chambers 2002; Go´dziewski & Konacki 2006; Smith & Lissauer 2010). In z Kepler data, the two co-orbital candidates began separated by ∼118◦ , with the trailing candidate reducing the gap at the rate of ∼1◦ per month. The numerical integration (section 4) matching the periods and phases in B11, circular orbits, and nominal masses went unstable at 25 Myr, precipitated by a close encounter between the coorbital planets.
  • 13. – 13 – This system will be difficult to study because of the faintness of the target star (Kepler magni- tude 15.34) and its location within the field which implies that it cannot be observed by Kepler dur- ing winter quarters (Q4, Q8, ...). Nonetheless, the remarkably commensurate period ratios of these four candidates give us strong confidence that they all will eventually be confirmed as planets. The dynamics of this class of planetary systems are analyzed by Fabrycky et al. (2011; in preparation), whose results further strengthen our confidence in these candidates. 5.4. KOI-191: A System Protected by Resonance? Two of the planets in the KOI-191 system have a period ratio of 1.258. One of those planets has a ∼ 1RJup size. Its mass would need to be 17M⊙ to satisfy equation 3. Alternatively, the two planets could be locked in 5:4 mean-motion resonance that prevents close approaches even if orbits cross. Neptune and Pluto are locked in 3:2 mean-motion resonance of this type. Relatively fast precession, on the order of 100 orbital periods, would be needed as well, as the observed period ratio is significantly different from the nominal resonance location. 5.5. KOI-500: (Near-)Resonant Five-Candidate System KOI-500 is a 5-candidate system with periods 0.986779 d, 3.072166 d, 4.645353 d, 7.053478 d, and 9.521696 d. Neighboring pairs of the outer four of these planets are all near two-body mean motion resonances, but unless there is an unexpectedly large amount of apse precession, they do not appear to be locked in librating two-body resonances. Perhaps this is a consequence of diverging tidal evolution, as predicted by Papaloizou & Terquem (2010). The three adjacent period spacings between these four planets contribute to the statistics of period ratios just wide of resonance (negative ξ). Moreover, the spacing of these systems from exact resonance, Poffset = 1/(1/Pout − 1/Pin ) (7) is equal to ∼ 190 days in all cases. The combination of mean motions 2n2 − 5n3 + 3n4 ≈ 1.6 × 10−5 and 2n3 − 6n4 + 4n5 ≈ 1.3 × 10−5 are so small that we suspect two 3-body Laplace-like resonances and/or a 4-planet resonance may be active and controlling the dynamics of this system. 6. Coplanarity of Planetary Systems The orbits of the planets in our Solar System lie close to the same plane. The distribution of planetary orbits around the invariable plane of the Solar System has a Rayleigh distribution width of σi ≃ 1.5◦ ; if Mercury is excluded, then σi ≃ 1◦ . The strong level of co-planarity has been recognized as an important constraint on models of planet formation for over 250 years (Kant 1755; Laplace 1796) and our current understanding is that growth within a dissipative protoplanetary
  • 14. – 14 – disk generally yields circular orbits and low relative inclinations (e.g., Safronov 1969; Lissauer 1993). Some mechanisms for explaining the higher eccentricity distribution of large exoplanets are likely to increase inclinations as well, such as planet-planet scattering (e.g., Chatterjee et al. 2008; Juri´ & Tremaine 2008). Whatever the mechanism, it is clear that measuring the inclination c distribution of typical planetary systems will create an important test for planet formation theories. The interplay between effects due to the protoplanetary disk, distant perturbers, and planet- planet scattering will leave a unique fingerprint in the coupled eccentricity-inclination distribution of exoplanetary systems and the exoplanet inclination distribution is thus of immense value. Yet neither transit observations, nor the radial velocity technique can directly measure the true mutual inclination between planets observed in multiple systems, except in unusually fortuitous circum- stances. This continues to be true in multi-transiting systems: even though the inclinations to the line of sight of all transiting planets must be small, the orbits could be rotated around the line of sight and mutually inclined to each other. Indirect constraints, however, can be obtained for systems with multiple transiting planets that make them the best probe of mutual inclinations around main sequence stars. For example, the lack of transit duration variations of the planets of Kepler-9 (Holman et al. 2010) and Kepler-11 (Lissauer et al. 2011) already place interesting lim- its on mutual inclinations to be 10◦ , while the inclinations to the line-of-sight of the Kepler-10 planets (Batalha et al. 2011, in prep.) require a mutual inclination 5◦ . Furthermore, given the large number of multi-candidate systems identified in B11, we can approach the question of the inclination distribution of the population statistically. The more mutually inclined a given pair of planets is, the smaller the probability that multiple planets transit (Ragozzine & Holman 2010), and thus the smaller the fraction of multi-planet systems that we would expect to see. Therefore, an investigation of coplanarity requires also addressing the number of planets expected per planetary system. Of course, the question of planetary multiplicity is also inherently interesting to investigate. Here we attempt to provide a self-consistent rigorous estimate of the multiplicity and coplanarity of typical systems using a statistical approach. 6.1. Simulated Population Model 6.1.1. Model Background The number of observed transiting planets per star is clearly affected by multiple observational biases, most notably that the probability of transiting decreases with increasing orbital period. With the current observations, we have little to no insight on the non-transiting planets in these systems, though eventually investigations of transit timing and duration variations (or the lack thereof) will be able to place additional limits on such planets (Steffen et al. 2010, Ford et al. in preparation). If we knew about the presence of every planet, transiting or not, then the problem of determining the typical inclination would be much easier.
  • 15. – 15 – A useful way to address the problem of non-detections is to create a forward model of transiting planet detections based on a minimal number of assumptions and a relatively small number of tunable parameters and to compare the output with the properties of the observed systems. The model described here attempts use Kepler ’s observed frequency of planet multiplicity to determine typical values for the true multiplicity of planetary systems as well as the mutual inclination between planets in the same system. After choosing a distribution for the number of planets assigned to a random Kepler star (characterized by a parameter Np ) and their relative inclinations (characterized by σi ), the model computes how many of these planets would be detected, requiring a geometrical alignment that leads to transit and a signal-to-noise ratio (SNR) large enough to be detected. This process is done for over 100000 planetary systems and the results compared to frequency of observed multi-transiting systems. Though imperfections in the model and observations will only provide an approximate answer, Kepler enables for the first serious attempt at using observations to measure the value of Np and σi for a large number of systems. To isolate the true multiplicity and inclination distributions, other aspects of the simulated systems (such as radius and period distributions) must be consistent with the true underlying distribution. This requires some interpretation of the observed distribution and the identification of potential biases. At the end of Kepler ’s mission, a simulated population that attempts to match all the observations could be employed; here, we hold the period and radius distributions fixed and they are pre-selected in a way that the output of the simulation will closely match the observed distribution. A good example of this process is the choice of the period distribution. The difference between the observed and true underlying period distribution is almost entirely determined by the geometric bias of observing planets in transit that scales as 1/a ∝ P −2/3 . As discussed in B11 and confirmed by our investigations, after correcting for the geometric bias and discarding very short periods (below 3 days) and large planets (above 6 Earth radii), the remaining distribution is very well described by a uniform distribution in log(P ). Hence, our simulator uses a period distribution uniform in log(P ) for a pre-specified range of periods: 3-125 days. An improvement could be placed on this process that takes into account the observed relationship (generally expected from planet formation theories), that larger period planets have larger radii (Figure 10), but this covariance is not included in the simulated planetary systems discussed below. Estimating the true radius distribution of Kepler planets is more complicated. A simple his- togram of planet radii shows that above ∼2 Earth radii, the observed population follows approxi- mately a 1/r 2 distribution (B10, B11). Below ∼1.5-2 Earth radii, the number of planets decreases significantly, which is at least partly due to small planets escaping detection due to low SNR. The 1/r 2 distribution is very sensitive to the minimum allowed radius; to avoid this, we draw radii from the distribution of observed planetary radii around stars with Kepler magnitude greater than 14 and through-Q5 SNR (from Table 1 in B11) greater than 16. A histogram of this distribution, and the full distribution of all detected planets, is shown in Figure 11. In particular, our simulated planets have radii drawn from this distribution between the limits of 1.5 and 6 Earth radii; except
  • 16. – 16 – at the smallest radii, this nearly approximates the 1/r 2 distribution observed in B11. Finally, an inspection of the B11 SNR ratios (which are calculated from Quarters 1-5) suggest that completeness effects begin to dominate below a SNR of ∼16, corresponding to a SNR of 9.2 from Quarters 1-2. We require that our simulated planets match this SNR in order to be detectable. We did not consider the FOP vetting flag in our choice of systems as a large fraction of the multiples have not yet been investigated in detail. Applying these same cuts to Kepler ’s observed multi-candidate systems from B11 results in 437, 63, 17, 1, and 1 systems with 1, 2, 3, 4, and 5 transiting planets, respectively. It is possible that some small fraction of the observed planets could be false positives and our simulation does not attempt to account for this. Note that the sixth planet of Kepler-11, planet g, had a transit that fell into the data gap between Quarters 1 and 2, so this system counts as a quintuple for comparison to the simulated population (a minor difference that does not affect our results). 6.1.2. Model Description Under the constraints described above, we want to create simulated planetary populations that will be compared to the Kepler observations. In particular, for each simulated population, we will compared the number of systems with i detectable transiting planets for 1 ≤ i ≤ 5 to the observed distribution from Kepler . The model starts by taking the full stellar population from the 161836 Q2 target stars in the “EX”oplanet program (the list of which is available from MAST), with the assumed masses and radii from the Kepler Input Catalog (Latham et al. 2005, Brown et al. in preparation). After randomly choosing a star, the model then assigns a number of planets, Np . Three distributions were considered for the number of planets: a “uniform” distribution, a Poisson distribution, and an exponential distribution. In the exponential distribution, the probability of having i + 1 planets is a factor of α smaller than the probability of having i planets, with α a free parameter. This model created a very poor fit to the data (particularly the triples to doubles ratio) for any value of α or σi , so we do not consider it further. The “uniform” distribution parameter Np,U is the mean number of planets per system, which is a fixed number of planets per star if Np,U is an integer or a random mix of two fixed values if Np,U is not an integer. For example, if Np,U = 3.25, 75% of systems would be assigned 3 planets and 25% would be assigned 4 planets. The Poisson distribution parameter Np,λ assigns the number of planets per system based on a Poisson distribution with width Np,λ . The frequency of stars with planetary systems is considered separately (see below), so we require the number of planets to be non-zero; for Np,λ 1, the distribution differs somewhat from a true Poisson distribution. Once the each star is hosting a specific non-zero number of planets, the orbital and physical properties of these planets are assigned. As discussed above, the period of each planet is selected
  • 17. – 17 – by drawing from a uniform distribution in log(P ) with a minimum of 3 and a maximum of 125 days. Although the period ratios in multiple systems deviate from a random distribution due to clustering near resonance (Section 5), this is a minor effect compared to the overall geometric bias. Each planetary radius is assigned based on the Kepler observed radii independently of orbital period, as described above. As the observed multiples are almost entirely stable on long timescales (Section 4), there is a desire to impose a proximity constraint that rejects simulated multiples that are too closely packed. This will require assigned a mass to each planet as well. We follow the mass-radius relation used above (eq. 1), i.e., M = R2.06 with masses and radii in units of Earth masses and radii. The mutual Hill separation between each pair of planets, ∆ can then be calculated as described in equation 3. In practice, the systems are built one planet at a time until all the planets assigned to this star (based on Np ) are selected. The first planet is assigned its period, radius, and corresponding mass. Starting with the second planet, a new period, radius, and mass are assigned and ∆ calculated. If any two planets have ∆ < 3.46 or if ∆inner + ∆outer < 18 for any three consecutive planets, then the planet is rejected on stability grounds and the process begins again until all the planets have been assigned. Very rarely, over the wide parameter space we are considering, large numbers of planets ( 10) are assigned which will not fit within the period range and stability criteria; in this case, as many planets as can fit are given to the star. Inclinations with respect to a reference plane are drawn with a Rayleigh distribution (appro- priate for this kind of geometric problem) of width σi and random nodal longitudes are assigned. In general, we assume that eccentricities are zero, and our investigations that considered non-zero eccentricities indicated that, statistically speaking, the observed multiplicities did not change sig- nificantly, though we did not consider the possibility that eccentricities and inclinations may be correlated. At this point, the full orbital and physical characteristics of the planetary systems have been determined and we can begin to assess observability. First, these planetary systems are then rotated by a uniform random rotation matrix in a manner equivalent to choosing a random point on a sphere for the direction of the normal to the reference plane. In the new random orientation, the impact parameter of each planet is calculated. If the planets are transiting, the SNR of a single transit is calculated assuming a box-shaped transit with depth of (Rp /R∗ )2 and duration calculated from the equation in (Kipping 2010). A random epoch is assigned and the number of transits in a 127-day period (corresponding to the duration of Q1-Q2) is calculated. To account for the duty cycle of 92%, each of these transits can be independently and randomly lost with 8% probability. The total signal is then calculated as the signal for a single transit times the square root of the number of observed transits. This is compared to the estimated noise over the course of the duration of the transit, calculated from the TMCDPP values for this star from Quarter 2 (see Kepler Data Release Notes). As described above, we require that our simulated planets have SNR of 9.2 in the simulated Quarters 1 and 2 in order to be detected. We find that approximately 30-50% of
  • 18. – 18 – simulated transiting planets are “missed” due to insufficient SNR. The output of the model is the number of stars with N detectable transiting planets, for all N . The next step is a statistical comparison between a variety of simulated populations (with different choices for Np and σi ) with the observed distribution from Kepler . An assessment of whether a particular simulation is rejectable is done using a G-test (similar to the categorical χ2 test) by calculating the G-statistic: G= 2Oi ln(Oi /Ei ) (8) i where Oi is the number of systems with i planets observed by Kepler and Ei is the expected number of such systems from the simulated population, which is scaled so that the total values of O and E are the same. The values Oi and Ei always refer to the number of systems with i observed and expected detectable transiting planets. Note that this statistic gives the most weight to categories with large numbers of objects and no weight to the categories where there are no observed or expected systems (i.e. i > 5). The G-statistic grows as the two populations become more different. Though the value of the G-statistic follows the χ2 distribution when O and E are large enough, it is more robust to compare the G value to a suite of G-statistics from a Monte Carlo simulation. Using the expected frequencies Ei , 100 random Monte Carlo populations are created with the true underlying distribution and their G-statistics are calculated. The significance of the observed value of G is calculated by comparing it to the Monte Carlo generated G-statistics. For example, if the observed value of G is 10 and this is larger than 95% of the Monte Carlo values of G, then the hypothesis that the population observed by Kepler is the same as the simulated population can be rejected with 95% confidence. 6.2. Results and Discussion The simulated population assumes that all 161836 Q2 target stars have planetary systems. B11 report that the expected number of planets per star based on the Kepler observations is about 0.4, though this describes planets of all types (unlike the limited size range we are considering) and converting this to the fraction of stars with planetary systems requires dividing by the typical multiplicity. Our investigation could address the frequency of planetary systems by adding a new free parameter fp and fitting to Oi with 0 ≤ i ≤ 5. However, this fit would be equivalent to: fitting a scaled population of systems that all contained planets to Oi , i > 0; calculating the scaled number of systems that had planets but did not transit, E0 ; and setting E0 = (1 − fp )O0 . That is, we can return to the question of the frequency of planetary systems after we find simulated populations that match the observed frequencies of systems where at least one transiting planet is detected. The first result is that there is no simulated population that adequately fits the observed distribution Oi for 1 ≤ i ≤ 5. The only population that adequately describes the singly-transiting systems has a uniform distribution of two planets per star (Np,U = 2) and an inclination width of σi = 4◦ . While the G-statistic shows that this solution has high confidence, clearly any population
  • 19. – 19 – that has only 2 planets per star cannot explain the large number of triples and higher-multiplicities seen in the Kepler observations. The “good fit” here is biased by the larger weight that the G- statistic gives to categories with more members and that categories with no observations have no weight. However, this model is useful to point out as one that correctly reproduces the ratio of doubles to singles by requiring a typical inclination dispersion of σi = 4◦ , corresponding to a Rayleigh distribution with a mean of ¯ ≡ σi π/2 ≃ 5◦ . i All other simulated populations produce relatively fewer singly-transiting systems than are observed. The difficulty can be seen qualitatively by comparing the ratio of doubly-transiting systems to singly-transiting systems (O2 /O1 = 14.5%) to the ratio of triply-transiting systems to doubly-transiting systems (O3 /O2 = 25.8%). These two ratios are quite different, yet the proba- bility of an additional planet transiting depends only on the semi-major axis ratio and the mutual inclination (Ragozzine & Holman 2010). We have also found that the probability that a planet passed our detection threshold is more-or-less independent of multiplicity, i.e., the detectability can be estimated on a planet-by-planet basis. (In detail, systems with high multiplicities tend to have planets with longer periods which lowers the SNR somewhat.) The simulated populations have both the geometric and detectability probabilities decreasing at an approximately constant rate as additional planets are added, leading to ratios Ei+1 /Ei that are nearly constant. Unlike in the single homogeneous simulated populations, the observed ratios are not similar and Kepler seems to be seeing an “excess” of singly-transiting systems. Note that this cannot be attributed to the fact that “hot Jupiters” appear to be singletons (Latham et al. 2011, in prep), since we are considering objects that are Neptune size or smaller. While it is possible that the inability to create a simulated population that matches the observations of multiples by Kepler is due to a shortcoming in the model, a theoretically motivated alternative is that the true planet population that has a bimodal distribution of characteristic inclination, with some have small inclinations and typically more planets than the other group. This is strongly suggested for distant planets (e.g., Levison et al. 1998) and may be present in the close-in population we are modeling here due to a distinction between systems that have had large dynamical scatterings in the past versus those systems which have remained relatively calm. As we are only attempting to match 5 multiplicity frequencies Oi , we cannot justify adding a full second population to our fits since this would require 4 highly degenerate parameters. Instead, we suppose that there may be a excess population of singly-transiting systems, and the simulated populations are compared again to the observations, but are now fit only to the observed population with two or more transiting planets (i.e., 2 ≤ i ≤ 5) using the same G-statistic method described above. Most of the constraint on this population is from the ratio of O3 /O2 . Note that a second planetary population with a large inclination dispersion would also contribute slightly to the number of doubles and possibly triples, but that we are ignoring this contribution. The best fit (not rejectable at the 93% level) in this case is from a population of systems with Np,U = 3.25 (i.e., 75% of systems with 3 planets and 25% of systems with 4 planets) and with a
  • 20. – 20 – Rayleigh inclination width of σi = 2◦ , implying a mean inclination of ¯ ≃ 2.5◦ . This population i produced Ei = (62.1, 17.5, 1.4, 0) for i from 2-5, which is an excellent match to the observed values Oi = (63, 17, 1, 1). (Recall that the expected numbers are scaled so that the total number of planets is the same as the observations.) When scaled to the observed values, this population produced only E1 = 130.1 singly-transiting systems, suggesting that over half of the observed singles cannot be explained in a simple model that matches the higher multiplicities. Note that these fits have only 1 degree of freedom (4 categories - 1 from scaling - 2 parameters), so it is not a surprise that a wide variety of models provide moderately good fits to the data. Figure 12 gives a contour plot of the acceptability of simulated populations as a function of Np,U and σi . As expected, increasing the number of planets per system requires the inclination dispersion to increase in order to provide a good match to the Kepler data, which can be demonstrated quantitatively. These fits also cannot explain the large number of singly-transiting systems: the better the fit to Oi , 2 ≤ i ≤ 5, the lower the expected number of singly-transiting systems. Some of the simulated populations where the number of planets is taken from a Poisson dis- tribution also give moderately good fits. The best is rejectable at the 70% level and has Np,λ = 2.5 with σi = 4◦ . This simulation can (rarely) produce systems with more than 4 detectable transiting planets, Ei = (65.3, 14.3, 1.9, 0.4) for i from 2-5, but underproduces the number of triply-transiting systems. A contour plot showing these the acceptability levels as a function of Np,λ and σi is shown in Figure 13. This figure also demonstrates the trend that more planets require a larger inclination dispersion to match the Kepler observations. It is expected that not every star has planets in this size and period range that would have been detected by Kepler ; the Sun is an example of such a star. The population of doubles that best matched the observations for 1 ≤ i ≤ 5 required E0 =9699 stars, i.e., ∼6% of the Kepler sample. The best-fit population for 2 ≤ i ≤ 5 described above (Np , U = 3.25, σi = 2◦ ) required only E0 = 3225 stars (∼2%) to have, on average, 3.25 planets (within the period and radius limits) to reproduce the observations seen by Kepler . The Poisson distributed simulated population (Np , λ = 2.5, σi = 4◦ required 4925 stars (∼3%). (Other simulated populations that were adequate fits as shown in Figures 12 and 13 had similar values of E0 .) In these latter cases, an additional set of stars with planets are needed to account for the additional singly-transiting systems. Nevertheless, these numbers appear quite small and fall about a factor of 2 short of the completeness calculations reported in B11, even after accounting for the fact that we are considering a different size range and that the number of stars with planetary systems is the average number of planets with stars (calculated by B11) divided by the typical multiplicity. The source of this discrepancy is unclear. In multiple systems with significant width in the inclination distribution, a substantial fraction of observers will see at least one planet in transit, even if the periods are somewhat large. Our simulations suggest that only a small fraction (∼5%) of the ∼162000 stars observed by Kepler have large numbers of planets with periods less than 125 days and radii between 1.5 and 6 Earth radii. As our completeness SNR cutoff is just an approximation to the actual fraction of planetary signals
  • 21. – 21 – that Kepler will detect among all actual planetary transits, the true fraction of stars with planets will be different from this estimate, with a more detailed assessment of completeness to be done by the Kepler Mission in the future. For example, the large errors and, in some cases, biases in the KIC will propagate into our simulations to some degree, though the ratio of observed multiplicities which drive the fits above are not as strongly affected as the estimate of the number of stars with planetary systems. With these caveats, Kepler is providing incredibly powerful insight into the formation of plan- etary systems. A single population model does not match the observed distribution, suggesting a superposition of two populations. A small fraction of stars have multiple (3 or more) super-Earth or Neptune size planets with periods less than 125 days. It appears that many of the doubly-transiting systems and less than half of the singly-transiting systems are actually systems with 3-4 planets, with non-transiting planets potentially observable through TTVs (Ford et al. 2011, in preparation). Systems with large values of ∆ in Tables 1 and 2 are good candidates for systems where additional non-transiting planets may be missing in between. We have also provided the first estimate of the inclination dispersion that has been given for a large sample of observed planetary systems. To match the numbers of higher multiples observed by Kepler requires a population with relatively low inclination dispersion; mean inclinations greater than 5◦ are poor fits. It does not appear that all of the multiples observed by Kepler come from a population of purely coplanar systems. Systems with multiple super-Earth and Neptune size planets with periods less than 125 days at low relative inclinations suggest interactions with a protoplanetary disk that induce migration while damping inclinations. There are good prospects for improving our understanding of the planetary populations ob- served by Kepler . Besides continued observations which will increase the detectability of planets and find new candidates in multiple systems, a more detailed population simulator could be de- veloped. For example, though the errors in the impact parameters of Kepler planets are usually relatively large, they do contain information on relative inclinations. This can be used both in a statistical population sense as well as for individual systems. In some systems, the detection of or upper limits to Transit Duration Variations (TDVs) will constrain the unobserved component of the relative inclinations of planets. Potential mutual events in Kepler multi-planet systems will be discussed and studied in more detail by Ragozzine et al. (2011, in preparation). Measurements of the Rossiter-McLaughlin effect for planets within multi- planet systems to measure true mutual inclination would constrain orbital inclinations to star’s equator, though this is observationally challenging for most Kepler stars (Ragozzine & Holman 2010). Each measurement of the true mutual inclination in individual systems helps fill in the picture for the typical inclination distribution of planetary systems.
  • 22. – 22 – 7. How Rare are Planetary Systems Similar to Kepler-11 The Kepler-11 (= KOI-157) planetary system, with six transiting planets, is clearly an outlier to the number/frequency distribution for detected transiting planets. Moreover, the period ratio of the two inner planets is only 1.264, which is the 4th lowest ratio among the 238 neighboring pairs of transiting Kepler exoplanets (Figure 3). The smallest period ratio is 1.001, between the candidate co-orbital planets in KOI-730 (Section 5.3). The second smallest period ratio, 1.038, is between two weak (vetting flag 3) candidates in the KOI-284 system; this system would be unstable for any reasonable planetary masses if indeed both of these candidates are actually planets in orbit about the same star (section 4). The third smallest period ratio, 1.258, is for a pair of candidates in the KOI-191 system that stability considerations strongly suggest are locked in a 5:4 mean motion resonance that prevents close approaches (section 5.4). Thus, Kepler-11b and Kepler-11c may well have the smallest period ratio of any non-resonant pair of planets in the entire candidate list. Additionally, the five inner planets of Kepler-11 travel on orbits that lie quite close to one another, both in terms of period ratio and absolute distance. Kepler-11 clearly does not possess a “typical” planetary system, but are systems of this type quite rare, or simply somewhat scarce? The two candidate most analogous to Kepler-11 are KOI-500 and KOI-707. With 5 planet candidates, KOI-500 is second in abundance to Kepler-11 and four of these candidates are close to one another in terms of period ratio (Table 4). The target star is significantly smaller than the Sun (∼0.74 R⊙ ; B11), and the orbital periods of the candidates are shorter, so if confirmed the planets in this system would be most closely spaced in terms of physical distance between orbits of any several planetary systems known. Moreover, four of the planets orbiting KOI-500 appear to be locked in three-planet resonances (Section 5.5), whereas no analogous situation is observed in Kepler-11. Unfortunately, KOI-500 is quite faint, so it will be more difficult to study than is Kepler-11. The four candidates in KOI-707 have periods within the range of the periods of the 5 inner planets observed in the Kepler-11 system. Periods ratios of neighboring candidates in the four planet KOI-117 are all less than two, but nonetheless significantly larger than in Kepler-11 and KOI-707. None of the six other four candidate systems is nearly as closely spaced. Furthermore, the better-fitting population simulations described in 6 rarely produced 5-planet systems, suggesting that Kepler-11 is not just a natural extension of Kepler multiple systems. Although its solitary nature makes determining a lower bound on planetary systems like Kepler-11 an ill-posed question (Kepler could have just gotten lucky), we can estimate the number of Kepler- 11 systems that would not be seen in transit probabilitistically. If all the Kepler-11 planets are coplanar (which is slightly inconsistent with the observed impact parameters), then the probability of seeing all six planets transit is the same as seeing Kepler-11g transit, which is 1.2%, suggesting that there are dozens of systems similar to Kepler-11, unless we were observing from a particularly favorable position.
  • 23. – 23 – 8. Conclusions Analysis of the first four months of Kepler data reveals 170 targets with more than one tran- siting planet candidate (B11). While the vast majority of these candidates have yet to be validated as true planets, we expect that the fidelity of this subsample of Kepler planet candidates to be high (Section ), and the small fraction of false positives not to affect the robustness of our statistical results. Our major conclusions are: 1. The large number of multiple planet systems observed by Kepler show that flat multi-planet systems are common in short-period orbits around other stars. This result holds for planets in the size range of ∼ 1.5 – 6 R⊕ , but not for giant planets. Not enough data are yet available to assess its viability for Earth-size and smaller worlds, nor for planets with orbital periods longer than a few months. The abundance and distribution of multiply transiting systems implies that the mean number of planets per star with at least one planet of this type is ∼3-4. Thus, only ∼3% of stars have planets within these in size-period limits. 2. Almost all candidate systems survived long-term dynamical integrations that assume circular, planar orbits and a mass-radius relationship derived from planets within our Solar System 1. This bolsters the evidence for the fidelity of the sample. 3. Mean-motion orbital resonances and near resonances are favored over random statistical prob- ability, but only a small fraction of multiple candidates are in or very near such resonances. In particular, resonances seem to be less common among nearly coplanar planets with short periods and radii a few times that of Earth than among the typically more massive and dis- tant planets detected by RV (Wright et al. 2011). Nonetheless, we note that a few systems are likely to have particularly interesting resonances configurations (Section 5). 4. Taken together, the properties of the observed systems suggest that the majority are free from false positives. It is unlikely that all of the candidates in multiple system are false positives, but there may be systems where one of the candidates is not a transiting planet around the same star as the other candidate(s). Those very near mean-motion resonances are most likely to be true planetary systems. 5. No single population of multiple planets with a single inclination dispersion can describe all the Kepler transiting systems. It appears that Kepler systems come from two populations which produce an excess in singly-transiting systems. Among systems with two or more transiting planets, the inclination dispersion appears to have a mean between 1-5◦ , suggesting relatively low mutual inclinations similar to the Solar System. Many more Kepler targets are multi- planet systems where additional non-transiting planets are present and these may eventually show TTVs.
  • 24. – 24 – The rich population of multi-transiting systems discovered by Kepler and reported by B11 have immense value both as individuals and as a population for improving our understanding of the formation and evolution of planetary systems. The Kepler spacecraft is scheduled to continue to return data on these multi-planet system for the remainder of its mission, and the longer temporal baseline afforded by these data will allow for the discovery of more planets and more accurate measurements of the planets and their interactions. Kepler was competitively selected as the tenth Discovery mission. Funding for this mission is provided by NASA’s Science Mission Directorate. The authors thank the many people who gave so generously of their time to make the Kepler mission a success. D. C. F. acknowledge NASA support through Hubble Fellowship grants #HF-51272.01-A and #HF-51267.01-A, respectively, awarded by STScI, operated by AURA under contract NAS 5-26555. We thank Avi Loeb, Josh Carter, and others for valuable discussions. REFERENCES Agol, E., Steffen, J., Sari, R., & Clarkson, W. 2005, MNRAS, 359, 567 Anglada-Escud´, G., L´pez-Morales, M., & Chambers, J. E. 2010, ApJ, 709, 168 e o Borucki, W. J., Kepler, & Team. 2011a, ApJ —. 2011b, ApJ, in press Borucki, W. J., Koch, D., Basri, G., Batalha, N., Brown, T., Caldwell, D., Caldwell, J., Christensen- Dalsgaard, J., Cochran, W. D., DeVore, E., Dunham, E. W., Dupree, A. K., Gautier, T. N., Geary, J. C., Gilliland, R., Gould, A., Howell, S. B., Jenkins, J. M., Kondo, Y., Latham, D. W., Marcy, G. W., Meibom, S., Kjeldsen, H., Lissauer, J. J., Monet, D. G., Morrison, D., Sasselov, D., Tarter, J., Boss, A., Brownlee, D., Owen, T., Buzasi, D., Charbonneau, D., Doyle, L., Fortney, J., Ford, E. B., Holman, M. J., Seager, S., Steffen, J. H., Welsh, W. F., Rowe, J., Anderson, H., Buchhave, L., Ciardi, D., Walkowicz, L., Sherry, W., Horch, E., Isaacson, H., Everett, M. E., Fischer, D., Torres, G., Johnson, J. A., Endl, M., MacQueen, P., Bryson, S. T., Dotson, J., Haas, M., Kolodziejczak, J., Van Cleve, J., Chandrasekaran, H., Twicken, J. D., Quintana, E. V., Clarke, B. D., Allen, C., Li, J., Wu, H., Tenenbaum, P., Verner, E., Bruhweiler, F., Barnes, J., & Prsa, A. 2010, Science, 327, 977 Bryson, S. T., Tenenbaum, P., Jenkins, J. M., Chandrasekaran, H., Klaus, T., Caldwell, D. A., Gilliland, R. L., Haas, M. R., Dotson, J. L., Koch, D. G., & Borucki, W. J. 2010, ApJ, 713, L97 Caldwell, D. A., Kolodziejczak, J. J., Van Cleve, J. E., Jenkins, J. M., Gazis, P. R., Argabright, V. S., Bachtell, E. E., Dunham, E. W., Geary, J. C., Gilliland, R. L., Chandrasekaran, H.,
  • 25. – 25 – Li, J., Tenenbaum, P., Wu, H., Borucki, W. J., Bryson, S. T., Dotson, J. L., Haas, M. R., & Koch, D. G. 2010, ApJ, 713, L92 Chambers, J. E. 1999, MNRAS, 304, 793 Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008, ApJ, 686, 580 Dawson, R. I., & Fabrycky, D. C. 2010, ApJ, 722, 937 Fabrycky, D. C. 2009, in IAU Symposium, Vol. 253, IAU Symposium, 173–179 Ford, E. B., Quinn, S. N., & Veras, D. 2008, ApJ, 678, 1407 Giuppone, C. A., Tadeu dos Santos, M., Beaug´, C., Ferraz-Mello, S., & Michtchenko, T. A. 2009, e ApJ, 699, 1321 Gladman, B. 1993, Icarus, 106, 247 Goldreich, P. 1965, MNRAS, 130, 159 Go´dziewski, K., & Konacki, M. 2006, ApJ, 647, 573 z Holman, M. J. 2010, in EAS Publications Series, Vol. 42, EAS Publications Series, ed. K. Go˙ dziewski, A. Niedzielski, & J. Schneider, 39–54 z Holman, M. J., Fabrycky, D. C., Ragozzine, D., Ford, E. B., Steffen, J. H., Welsh, W. F., Lissauer, J. J., Latham, D. W., Marcy, G. W., Walkowicz, L. M., Batalha, N. M., Jenkins, J. M., Rowe, J. F., Cochran, W. D., Fressin, F., Torres, G., Buchhave, L. A., Sasselov, D. D., Borucki, W. J., Koch, D. G., Basri, G., Brown, T. M., Caldwell, D. A., Charbonneau, D., Dunham, E. W., Gautier, T. N., Geary, J. C., Gilliland, R. L., Haas, M. R., Howell, S. B., Ciardi, D. R., Endl, M., Fischer, D., F¨r´sz, G., Hartman, J. D., Isaacson, H., Johnson, ue J. A., MacQueen, P. J., Moorhead, A. V., Morehead, R. C., & Orosz, J. A. 2010, Science, 330, 51 Holman, M. J., & Murray, N. W. 2005, Science, 307, 1288 Jenkins, J. M., Caldwell, D. A., Chandrasekaran, H., Twicken, J. D., Bryson, S. T., Quintana, E. V., Clarke, B. D., Li, J., Allen, C., Tenenbaum, P., Wu, H., Klaus, T. C., Middour, C. K., Cote, M. T., McCauliff, S., Girouard, F. R., Gunter, J. P., Wohler, B., Sommers, J., Hall, J. R., Uddin, A. K., Wu, M. S., Bhavsar, P. A., Van Cleve, J., Pletcher, D. L., Dotson, J. A., Haas, M. R., Gilliland, R. L., Koch, D. G., & Borucki, W. J. 2010, ApJ, 713, L87 Juri´, M., & Tremaine, S. 2008, ApJ, 686, 603 c Kant, I. 1755, Allgemeine Naturgeschichte und Theorie des Himmels Kipping, D. M. 2010, MNRAS, 407, 301
  • 26. – 26 – Koch, D., Borucki, W., Cullers, K., Dunham, E., Webster, L., Miers, T., & Reitsema, H. 1996, J. Geophys. Res., 101, 9297 Koch, D. G., Borucki, W. J., Basri, G., Batalha, N. M., Brown, T. M., Caldwell, D., Christensen- Dalsgaard, J., Cochran, W. D., DeVore, E., Dunham, E. W., Gautier, T. N., Geary, J. C., Gilliland, R. L., Gould, A., Jenkins, J., Kondo, Y., Latham, D. W., Lissauer, J. J., Marcy, G., Monet, D., Sasselov, D., Boss, A., Brownlee, D., Caldwell, J., Dupree, A. K., Howell, S. B., Kjeldsen, H., Meibom, S., Morrison, D., Owen, T., Reitsema, H., Tarter, J., Bryson, S. T., Dotson, J. L., Gazis, P., Haas, M. R., Kolodziejczak, J., Rowe, J. F., Van Cleve, J. E., Allen, C., Chandrasekaran, H., Clarke, B. D., Li, J., Quintana, E. V., Tenenbaum, P., Twicken, J. D., & Wu, H. 2010, ApJ, 713, L79 Laplace, P. 1796, Exposition du Systeme du Monde Latham, D. W., Brown, T. M., Monet, D. G., Everett, M., Esquerdo, G. A., & Hergenrother, C. W. 2005, in Bulletin of the American Astronomical Society, Vol. 37, Bulletin of the American Astronomical Society, 1340 Laughlin, G., & Chambers, J. E. 2002, AJ, 124, 592 Levison, H. F., Lissauer, J. J., & Duncan, M. J. 1998, AJ, 116, 1998 Lissauer, J. J. 1993, ARA&A, 31, 129 Lissauer, J. J., Fabrycky, D. C., Ford, E. B., & et al. 2011, Nature, 470, 53 Lissauer, J. J., & Slartibartfast. 2008, in Astronomical Society of the Pacific Conference Series, Vol. 398, Astronomical Society of the Pacific Conference Series, ed. D. Fischer, F. A. Rasio, S. E. Thorsett, & A. Wolszczan, 491 Malhotra, R. 1998, in Astronomical Society of the Pacific Conference Series, Vol. 149, Solar System Formation and Evolution, ed. D. Lazzaro, R. Vieira Martins, S. Ferraz-Mello, & J. Fernandez , 37 Nobili, A., & Roxburgh, I. W. 1986, in IAU Symposium, Vol. 114, Relativity in Celestial Mechanics and Astrometry. High Precision Dynamical Theories and Observational Verifications, ed. J. Kovalevsky & V. A. Brumberg, 105–110 Ragozzine, D., & Holman, M. J. 2010, preprint (arxiv:1006.3727) Ragozzine, D., & Wolf, A. S. 2009, ApJ, 698, 1778 Safronov, V. S. 1969, Evolution of the protoplanetary cloud and formation of the Earth and planets. Saha, P., & Tremaine, S. 1994, AJ, 108, 1962 Smith, A. W., & Lissauer, J. J. 2009, Icarus, 201, 381
  • 27. – 27 – —. 2010, Celestial Mechanics and Dynamical Astronomy, 107, 487 Steffen, J. H., Batalha, N. M., Borucki, W. J., Buchhave, L. A., Caldwell, D. A., Cochran, W. D., Endl, M., Fabrycky, D. C., Fressin, F., Ford, E. B., Fortney, J. J., Haas, M. J., Holman, M. J., Howell, S. B., Isaacson, H., Jenkins, J. M., Koch, D., Latham, D. W., Lissauer, J. J., Moorhead, A. V., Morehead, R. C., Marcy, G., MacQueen, P. J., Quinn, S. N., Ragozzine, D., Rowe, J. F., Sasselov, D. D., Seager, S., Torres, G., & Welsh, W. F. 2010, ApJ, 725, 1226 Terquem, C., & Papaloizou, J. C. B. 2007, ApJ, 654, 1110 Torres, G., Fressin, F., Batalha, N. M., Borucki, W. J., Brown, T. M., Bryson, S. T., Buchhave, L. A., Charbonneau, D., Ciardi, D. R., Dunham, E. W., Fabrycky, D. C., Ford, E. B., Gautier, III, T. N., Gilliland, R. L., Holman, M. J., Howell, S. B., Isaacson, H., Jenkins, J. M., Koch, D. G., Latham, D. W., Lissauer, J. J., Marcy, G. W., Monet, D. G., Prsa, A., Quinn, S. N., Ragozzine, D., Rowe, J. F., Sasselov, D. D., Steffen, J. H., & Welsh, W. F. 2011, ApJ, 727, 24 Veras, D., & Armitage, P. J. 2004, Icarus, 172, 349 Veras, D., & Ford, E. B. 2010, ApJ, 715, 803 Wisdom, J. 1980, AJ, 85, 1122 Wisdom, J., & Holman, M. 1991, AJ, 102, 1528 Wright, J. T., Fakhouri, O., Marcy, G. W., Han, E., Feng, Y., Johnson, J. A., Howard, A. W., Valenti, J. A., Anderson, J., & Piskunov, N. 2010, ArXiv e-prints Wright, J. T., Upadhyay, S., Marcy, G. W., Fischer, D. A., Ford, E. B., & Johnson, J. A. 2009, ApJ, 693, 1084 Wright, J. T., Veras, D., Ford, E. B., Johnson, J. A., Marcy, G. W., Howard, A. W., Isaacson, H., Fischer, D. A., Spronck, J., Anderson, J., & Valenti, J. 2011, preprint (arxiv:1101.1097) This preprint was prepared with the AAS L TEX macros v5.2. A
  • 28. – 28 – Table 1. Characteristics of Systems with Two Transiting Planets KOI # Rp,1 (R⊕ ) Rp,2 (R⊕ ) P2 /P1 ∆ 72 1.30 2.29 54.083733 90.8 82 6.84 3.70 1.565732 7.6 89 4.36 5.46 1.269541 4.8 112 3.68 1.71 13.770977 55.7 115 3.37 2.20 1.316598 7.4 116 4.73 4.81 3.230779 21.6 123 2.26 2.50 3.274199 34.1 124 2.33 2.83 2.499341 25.5 139 5.66 1.18 67.267271 54.4 150 3.41 3.69 3.398061 26.3 153 3.17 3.07 1.877382 14.2 209 7.55 4.86 2.702205 14.8 220 2.62 0.67 1.703136 17.3 222 2.06 1.68 2.026806 20.5 223 2.75 2.40 12.906148 55.8 232 3.58 1.55 2.161937 20.1 244 4.53 2.65 2.038993 15.5 260 1.19 2.19 9.554264 70.3 270 0.90 1.03 2.676454 49.1 271 1.99 1.82 1.654454 17.5 279 4.90 2.09 1.846201 13.4 282 2.77 0.86 3.252652 36.0 291 1.45 1.05 3.876548 57.0 313 3.10 2.17 2.220841 21.0 314 1.95 1.57 1.675518 14.8 339 1.47 1.08 3.240242 50.6 341 3.32 2.26 1.525757 11.0 343 2.19 1.64 2.352438 28.6 386 3.40 2.90 2.462733 21.7 401 6.24 6.61 5.480100 23.3 416 2.95 2.82 4.847001 35.8 431 3.59 3.50 2.485534 19.6 433 5.78 13.40 81.440719 28.7 440 2.80 2.24 3.198296 30.0 442 1.86 1.43 7.815905 67.9 446 2.31 1.69 1.708833 15.5 448 2.33 3.79 4.301984 28.2 456 3.12 1.66 3.179075 31.6 459 3.69 1.28 2.810266 26.6 464 7.08 2.66 10.908382 33.2 474 2.34 2.32 2.648401 28.7 475 2.36 2.63 1.871903 17.0 490 2.29 2.25 1.685884 15.1 497 2.49 1.72 2.981091 33.8 508 3.76 3.52 2.101382 16.3 509 2.67 2.86 2.750973 25.6
  • 29. – 29 – Table 1—Continued KOI # Rp,1 (R⊕ ) Rp,2 (R⊕ ) P2 /P1 ∆ 510 2.67 2.67 2.172874 20.6 518 2.37 1.91 3.146856 32.3 523 7.31 2.72 1.340780 4.9 534 2.05 1.38 2.339334 28.7 543 1.89 1.53 1.371064 11.1 551 2.12 1.79 2.045843 23.5 555 1.51 2.27 23.366026 77.3 564 2.35 4.99 6.073030 34.0 573 3.15 2.10 2.908328 28.3 584 1.58 1.51 2.138058 28.5 590 2.10 2.22 4.451198 44.3 597 2.60 1.40 8.272799 59.7 612 3.51 3.62 2.286744 18.1 638 4.76 4.15 2.838628 19.5 645 2.64 2.48 2.796998 28.9 657 1.59 1.90 4.001270 42.5 658 1.53 2.22 1.698143 18.1 663 1.88 1.75 7.369380 52.0 672 3.98 4.48 2.595003 18.8 676 4.48 2.94 3.249810 20.4 691 2.88 1.29 1.828544 18.8 693 1.80 1.72 1.837696 22.2 700 3.05 1.93 3.297027 32.2 708 2.19 1.71 2.262657 27.3 736 2.64 1.96 2.789104 24.7 738 3.30 2.86 1.285871 6.2 749 1.98 1.35 1.357411 11.0 752 2.69 3.39 5.734874 39.3 775 2.11 2.47 2.079976 17.9 787 2.87 2.19 1.284008 7.0 800 2.73 2.52 2.659942 26.3 837 1.83 1.45 1.919049 22.3 841 4.00 4.91 2.042803 13.0 842 2.78 3.14 2.835637 22.9 853 2.87 2.12 1.767103 15.2 869 3.19 4.33 4.845276 30.2 870 3.62 3.45 1.519922 8.9 877 2.50 2.32 2.021803 17.0 881 2.54 3.92 10.793324 43.8 896 3.85 2.81 2.574418 20.6 904 2.11 2.96 12.635882 50.2 936 3.46 2.04 10.601824 40.6 938 2.89 1.28 9.512348 57.5 945 2.04 2.64 1.575035 13.6 954 2.35 2.45 4.550137 40.9 1015 2.36 1.63 2.305817 27.5
  • 30. – 30 – Table 1—Continued KOI # Rp,1 (R⊕ ) Rp,2 (R⊕ ) P2 /P1 ∆ 1060 1.54 1.23 2.545144 39.7 1089 9.64 5.69 7.094116 22.6 1102 3.22 0.94 1.513919 12.3 1113 2.65 2.81 3.217321 30.7 1151 1.17 1.20 1.420286 16.6 1163 1.90 1.77 2.729492 33.8 1198 2.04 1.53 1.561840 16.0 1203 2.51 2.44 2.256556 23.0 1215 2.21 2.12 1.905318 20.8 1221 5.03 5.31 1.694236 10.0 1236 2.78 1.67 5.807534 49.5 1241 10.43 6.99 2.039957 9.3 1278 1.57 2.39 3.575571 40.0 1301 1.86 2.32 2.954666 32.3 1307 2.96 2.81 2.204782 20.1 1360 2.73 2.29 2.520266 24.0 1364 2.86 2.74 2.952861 27.1 1396 2.53 1.86 1.790301 17.8 1475 1.79 2.18 5.910858 43.3 1486 8.45 2.38 8.433622 27.9 1589 2.23 2.28 1.476360 12.0 1590 2.75 1.90 10.943201 55.8 1596 2.26 3.45 17.785655 54.0