You are on page 1of 7

A&A 379, 992998 (2001) Astronomy

DOI: 10.1051/0004-6361:20011373

&
c ESO 2001
Astrophysics

The distribution of exoplanet masses


A. Jorissen1,? , M. Mayor2 , and S. Udry2

1
Institut dAstronomie et dAstrophysique, Universite Libre de Bruxelles, CP 226, Boulevard du Triomphe,
1050 Bruxelles, Belgium
2
Observatoire de Geneve, 1290 Sauverny, Switzerland
e-mail: Michel.Mayor, Stephane.Udry@obs.unige.ch
Received 17 May 2001 / Accepted 24 September 2001
Abstract. The present study derives the distribution of secondary masses M2 for the 67 exoplanets and very
low-mass brown-dwarf companions of solar-type stars, known as of April 4, 2001. This distribution is related to
the distribution of M2 sin i through an integral equation of Abels type. Although a formal solution exists for
this equation, it is known to be ill-conditioned, and is thus very sensitive to the statistical noise present in the
input M2 sin i distribution. To overcome this difficulty, we present two robust, independent approaches: (i) the
formal solution of the integral equation is numerically computed after performing an optimal smoothing of the
input distribution and (ii) the Lucy-Richardson algorithm is used to invert the integral equation. Both approaches
give consistent results. The resulting statistical distribution of exoplanet true masses reveals that there is no reason
to ascribe the transition between giant planets and brown dwarfs to the threshold mass for deuterium ignition
(about 13.6 MJ ). The M2 distribution shows instead that most of the objects have M2 10 MJ , but there is a
small tail with a few heavier candidates around 15 MJ .

Key words. methods: numerical stars: planetary systems

1. Introduction The basic reason why the M2 distribution obtained


in Sect. 4 differs from that of Han et al. (2001) is be-
Han et al. (2001) suggested that most of the exoplanet cause these authors concluded that most of the sys-
candidates discovered so far have masses well above the tems containing exoplanet candidates are seen nearly
lower limit defined by sin i = 1 (where i is the inclination pole-on. This conclusion, based on the analysis of the
of the orbital plane on the sky) and should therefore be Hipparcos Intermediate Astrometric Data (IAD), has how-
considered as brown dwarfs or even stars rather than plan- ever been shown to be incorrect (e.g. Halbwachs et al.
ets. The present paper shows that this claim is not con- 2000; Pourbaix 2001; Pourbaix & Arenou 2001), as sum-
sistent with the distribution of masses extracted from the marized in Sect. 5.
observed M2 sin i distribution (where M2 is the mass of the
companion) under the reasonable assumption that the or- While this paper was being reviewed, Zucker & Mazeh
bits are oriented at random in space. Although the distri- (2001) and Tabachnik & Tremaine (2001) proposed other
butions of M2 and M2 sin i are related through an integral interesting approaches to derive the exoplanet mass dis-
equation of Abels kind (Chandrasekhar & Munch 1950; tribution. Zucker & Mazeh derive the binned true mass
Lucy 1974), its numerical solution is ill-conditioned. Two distribution by using a maximum likelihood approach to
different methods are used here to overcome this difficulty. retrieve the average values of the mass distribution over
In the first method (Sect. 2), the formal solution of Abels the selected bins. Their results are in very good agree-
equation is implemented numerically on an input M2 sin i ment with ours. Tabachnik & Tremaine suppose that the
distribution that has been optimally smoothed with an period and mass distributions follow power laws, and de-
adaptive kernel procedure (Silverman 1986) to remove rive the corresponding power-law indices from a maximum
the high-frequency fluctuations caused by small-number likelihood method. On the contrary, the methods used in
statistics. The other method (Sect. 3) is based on the the present paper (and in Zucker & Mazehs) are non-
Lucy-Richardson inversion algorithm (Richardson 1972; parametric in nature, since they do not require any a priori
Lucy 1974). assumptions on the functional form of the mass distribu-
tion. This is especially important since the comparison of
the shapes of the mass distributions for exoplanets and
Send offprint requests to: A. Jorissen, low-mass stellar companions may provide clues to iden-
e-mail: ajorisse@astro.ulb.ac.be tify the process by which they formed. By imposing a
?
Research Associate, FNRS, Belgium. power-law function like that observed for low-mass stellar

Article published by EDP Sciences and available at http://www.aanda.org or


http://dx.doi.org/10.1051/0004-6361:20011373
A. Jorissen et al.: The distribution of exoplanet masses 993

companions, Tabachnik & Tremaine (2001) implicitly as- While Eq. (8) represents the formal solution of the
sume that these processes must be similar. problem, it is difficult to implement numerically, since it
requires the differentiation of the observed frequency dis-
tribution (Y ). Unless the observations are of high pre-
2. The integral equation of Abels kind relating cision, it is well known that this process can lead to mis-
the distributions of M2 sin i and M2 leading results. To overcome this difficulty, the observed
frequency distribution has been smoothed in an optimal
The M2 sin i values for low-mass companions of main se- way (see Appendix) before being used in Eq. (8). The so-
quence stars may be extracted from the spectroscopic lution (t) is then computed numerically using standard
mass function and from the primary mass as derived differentiation and integration schemes.
through e.g., isochrone fitting. Let (Y ) be the ob-
served distribution of Y M2 sin i, that is easily derived 3. The Lucy-Richardson inversion algorithm
from the observed spectroscopic mass functions provided applied to Abels integral equation
that M2  M1 as is expected to be the case for the sys-
tems under consideration. Then, the distribution (M2 ) The Lucy-Richardson algorithm provides another robust
obeys the relation way to invert Eq. (4) (see also Cerf & Boffin 1994). The
Z method starts from the Bayes theorem on conditional
(Y ) = (M2 ) (Y |M2 ) dM2 . (1) probability in the form
0
(M2 ) (Y |M2 ) = (Y ) R(M2 |Y ), (9)
The kernel (Y |M2 ) corresponds to the conditional prob-
ability of observing the value Y given M2 . Under the as- where R(M2 |Y ) is the reciprocal kernel corresponding to
sumption that the orbits are oriented at random in space, the integral equation inverse to the one that needs to be
the inclination angle i is distributed as sin i, and the fol- solved (Eq. (1)):
lowing expression is obtained for the kernel: Z M2
sin i0 (M2 ) = (Y ) R(M2 |Y ) dY. (10)
(Y |M2 ) = , (2) 0
M2 cos i0
The reciprocal kernel represents the conditional probabil-
where i0 satisfies the relations M2 sin i0 Y = 0 and ity that the binary system has a companion mass M2 when
0 i0 90. Eliminating the inclination i0 in the above the observed M2 sin i value amounts to Y . Thus, one has:
relation yields
(M2 ) (Y |M2 )
Y 1 R(M2 |Y ) = (11)
(Y |M2 ) = with Y M2 . (3) (Y )
M2 (M2 Y 2 )1/2
2
(M2 ) (Y |M2 )
= R , (12)
Equation (1) thus rewrites 0
(M2 ) (Y |M2 ) dM2
Z
1 which
R obviously satisfies the normalization condition
(Y ) = Y (M2 ) dM2 . (4)
Y M 2 (M 2 Y 2 )1/2
0
R(M 2 |Y ) dM2 = 1. The problem in solving Eq. (10)
2
arises because R(M2 |Y ) also depends on (M2 ), so that
Equation (4) is the integral equation to be solved for an iterative procedure must be used. If r (M2 ) is the rth
(M2 ). It can be reduced to Abels integral equation by estimate of (M2 ), it can be used to obtain the (r + 1)th
the substitutions (Chandrasekhar & Munch 1950) estimate in the following way:
Y 2 = 1/ and M22 = 1/t. (5) Z M2
r+1 (M2 ) = (Y ) Rr (M2 |Y ) dY (13)
With these substitutions, Eq. (4) becomes 0
Z with
(t)
() = dt, (6) r (M2 ) (Y |M2 )
0 ( t)1/2 Rr (M2 |Y ) = (14)
r (Y )
where
and
1 1 1 Z
() ( ) and (t) ( ). (7)
2 t t r (Y ) = r (M2 ) (Y |M2 ) dM2 . (15)
0
It is well known that the solution of Abels equa-
tion (Eq. (6)) is given by Thus, r (Y ) represents the corresponding rth estimate of
Z the observed distribution (Y ). Equations (13) and (14)
1 t 1 1 (0) together yield the recurrence relation for the r s,
(t) = d + , (8)
0 (t )1/2 t Z M2
(Y )
where (0) = limY (Y ) = 0. r+1 (M2 ) = r (M2 ) (Y |M2 ) dY, (16)
0 r (Y )
994 A. Jorissen et al.: The distribution of exoplanet masses

with (Y |M2 ) given by Eq. (3) for the problem under


consideration. The conditions for convergence of this re-
currence relation are discussed by Lucy (1974) and Cerf
& Boffin (1994). It needs only be remarked here that (i)
the iterative scheme converges if r (Y ) tends to (Y ),
given the normalization of the probability (Y |M2 )dY ,
and (ii) the full convergence of the method is not nec-
essarily desirable, as the successive estimates r (Y ) will
tend to match (Y ) on increasingly smaller scales, but the
small-scale structure in (Y ) is likely to be dominated by
the noise in the input data. This is well illustrated in Fig. 2
below.
When the number of data points is small (typically
N < 100; Cerf & Boffin 1994), it is advantageous to ex-
press (Y ) as

1 X
N
(Y ) = (Y yn ) (17)
N n=1

where the yn (n = 1, . . . N ) are the N individual mea-


sured M2 sin
R i values and (x) is the Dirac function such
that x0 = (x x0 ) dx. Substitution in Eq. (13) then
yields

1 X
N
r+1 (M2 ) = Rr (M2 |yn ) (18)
N n=1

where Rr (M2 |yn ) is defined as in Eq. (14). The sample


size should nevertheless be large enough for the functions
Rr (M2 |yn ) to have sufficient overlap so as to produce a
smooth r+1 (M2 ) function.
In the application of the method described in Sect. 4,
the initial mass distribution 0 (M2 ) was taken as a uni-
form distribution, but it has been verified that this choice
has no influence on the final solution.

4. The frequency distribution (M2 )


The cumulative frequency distribution of the M2 sin i val-
ues smaller than 17 MJ , where MJ is the mass of Jupiter
(=M /1047.35), available in the literature (as of April 4,
2001) is presented in Fig. 1. It appears to be sufficiently Fig. 1. Bottom panel: the cumulative frequency distribu-
well sampled to attempt the inversion procedure. The tion of the observed M2 sin i values for the 67 exoplanets and
corresponding frequency distributions smoothed with two very low-mass brown dwarfs (around 60 stars) known as of
different smoothing lengths, locally self-adapting around April 4, 2001. Top panel: the corresponding frequency distribu-
hopt = 1 MJ and 2hopt (see Appendix) are presented as tion, represented either as an histogram, or smoothed using an
well for comparison. Epanechnikov kernel (see Appendix) with an optimal smooth-
The sample includes 60 main-sequence stars host- ing length of hopt = 1 MJ (dashed curve). The frequency dis-
ing 67 companions with M2 sin i < 17 MJ . Among tribution smoothed with 2hopt is also shown for comparison
those, 6 stars are orbited by more than one compan- (solid curve).
ion, namely HD 74156 and HD 82943 (M2 sin i = 1.50,
7.40 MJ , and M2 sin i = 0.84, 1.57 MJ , respectively; Udry
& Mayor 2001)1 , HD 83433 (0.16,0.38 MJ ; Mayor et al.
Gliese 876 (0.56, 1.88 MJ ; Marcy et al. 2001). The in-
2000), HD 168443 (7.22, 16.2 MJ ; Udry et al. 2000b),
version process is only able to treat these systems under
And (0.71, 2.20 and 4.45 MJ ; Butler et al. 1999) and
the hypothesis that the orbits of the different planets in
1
See also http://www.eso.org/outreach/press-rel/ a given system are not coplanar, since Eq. (4) to hold re-
pr-2001/pr-07-01.html quires random orbital inclinations. The case of coplanar
A. Jorissen et al.: The distribution of exoplanet masses 995

and non-coplanar orbits are discussed separately in the


remainder of this section.

4.1. Non-coplanar orbits


Figure 2 compares the solutions (M2 ) obtained from
the Lucy-Richardson algorithm (after 2 and 20 itera-
tions, denoted 2 and 20 , respectively) and from the
formal solution of Abels integral equation with smooth-
ing lengths hopt and 2 hopt on (Y ) (the corresponding
solutions are denoted hopt and 2hopt ). The solutions
from the two methods basically agree with each other,
although solutions with different degrees of smoothness
may be obtained with each method. On the one hand, 20
and hopt exhibit high-frequency fluctuations that may
be traced back to the statistical fluctuations in the input
data. This can be seen by noting that the peaks present
in hopt correspond in fact to the high-frequency fluctua-
tions already present in hopt (Fig. 1). These fluctuations
should thus not be given much credit. The same expla-
nation holds true for 20 , since it was argued in Sect. 3 Fig. 2. Comparison of (M2 ) distributions obtained from the
that the solutions r resulting from a large number of it- Lucy-Richardson algorithm (after 2 and 20 iterations, rep-
erations tend to match at increasingly small scales (i.e., resented by the solid and dashed histograms, respectively),
higher frequencies) where statistical fluctuations become and from the formal solution of Abels integral equation with
dominant. On the other hand, 2 and 2hopt are much smoothing lengths 2hopt and hopt applied on (Y ) (represented
smoother, and are probably better matches to the actual by solid and dashed curves, respectively). The supposedly best
distribution. The local maximum around M2 1 MJ is representation of the actual (M2 ) distribution is represented
by the thick solid line. Planetary systems have been included in
very likely, however, an artifact of the strong detection
the inversion process with the assumption of non-coplanarity
bias against low-mass companions. of the planetary orbits in a given system.
The most striking feature of the (M2 ) distribution
displayed in Fig. 2 is its decreasing character, reaching
zero for the first time around M2 = 10 MJ , and in any
case well before 13.6 MJ . The latter value, correspond- removed from the original set. Equation (8) is then ap-
ing to the minimum stellar mass for igniting deuterium, plied to these 67 different input distributions. The result-
does not in any way mark the transition between gi- ing distributions are displayed in Fig. 3, which shows that
ant planets and brown dwarfs, as sometimes proposed. the threshold observed at 10 MJ is a robust result not
That transition, which is thus likely to occur at smaller affected by the uncertainty on the solution.
masses, must rely instead on the different mechanisms
governing the formation of planets and brown dwarfs.
4.2. Coplanar orbits in multi-planets systems
Another argument favouring a giant-planet/brown-dwarf
transition mass smaller than 13.6 MJ is provided e.g., by All the results discussed so far are obtained under the as-
the observation of free-floating (and thus most likely stel- sumption that orbits of planets belonging to a planetary
lar) objects with masses probably smaller than 10 MJ in system are not coplanar. To evaluate the impact of this
the Orionis star cluster (Zapatero Osorio et al. 2000). hypothesis, the following procedure has been applied. In
The (M2 ) distribution nevertheless clearly exhibits a a first step, the Lucy-Richardson algorithm is applied on
tail of objects clustering around M2 15 MJ , due to the data set excluding the 13 planets belonging to plane-
HD 114762 (M2 sin i = 11.5 MJ ), HD 162020 (14.3 MJ ), tary systems. That mass distribution obtained after 2 it-
HD 202206 (15.0 MJ ) and HD 168443 c (16.2 MJ ). It would erations is then completed by mass estimates for the re-
be interesting to investigate whether these systems differ maining 13 planets. For each of the 6 different systems, an
from those with smaller masses in some identifiable way inclination i is drawn from a sin i distribution. This is done
(periods, eccentricities, metallicities, ...), so as to assess through the expression i = arccos x, where x is a random
whether or not they form a distinct class (Udry et al., in number with uniform deviation. The same value of i is
preparation). then applied to all planets in a given system to extract M2
The jackknife method (e.g., Lupton 1993) has been from the observed M2 sin i value. The distributions of ex-
used to estimate the uncertainty on the 2hopt solution. oplanet masses obtained with and without the hypothesis
In a first step, 67 input 2hopt distributions are computed, of coplanarity are compared in Fig. 4, and it is seen that
corresponding to all 67 possible sets with one data point planetary systems are not yet numerous enough for the
996 A. Jorissen et al.: The distribution of exoplanet masses

Fig. 3. Comparison of the input 2hopt distribution (thick Fig. 4. Evaluation of the impact of the coplanarity hypothe-
dashed line) and the 67 2hopt distributions (thin dotted lines) sis on the resulting (M2 ) distribution. The dashed histogram
resulting from the application of the jackknife method (see corresponds to the mass distribution obtained assuming copla-
text), which illustrates the uncertainty on the solution 2hopt nar orbits in planetary systems (see text), as compared to
(thick solid line). To guide the eye, a power-law of index 1.6 the 2 (solid histogram) and 2hopt solutions (see Fig. 2).
has been plotted as well (dot-dashed line).

coplanarity hypothesis to alter significantly the resulting


(M2 ) distribution. the true value, in fact of the order of the residuals. But
In any case, the main result of the present paper is since a sin i is constrained by the spectroscopic orbital el-
that the statistical properties of the observed M2 sin i dis- ements, the too-large astrometric a value will force i to be
tribution coupled with the hypothesis of randomly oriented close to 0 to match the spectroscopic value of the product
orbital planes confine the vast majority of planetary com- a sin i, as convincingly shown by Pourbaix (2001). Hence,
panion masses below about 10 MJ . Zucker & Mazeh (2001) this approach gives the impression that all orbits are seen
reach the same conclusion. nearly face-on. As an illustrative example, Pourbaix &
It should be remarked that the above conclusion can- Arenou (2001) have shown that such an approach leads
not be due to detection biases, since the high-mass tail to a stellar mass for the companion of HD 209458 that, on
of the M2 distribution is not affected by the difficulty of another hand, has been proven to be a 0.69 MJ planet by
finding low-amplitude, long-period orbits. the photometric observation of the planet transit in front
of the star (Charbonneau et al. 2000).
The Han et al. (2001) result is moreover statistically
5. Discussion
very unlikely if the orbital planes are oriented at random
Although the assumption of random orbital inclinations in space (Pourbaix & Arenou 2001). Han et al. (2001)
seems reasonable, it is at variance with the conclusion of have tried to justify this unlikely statistical occurrence by
Han et al. (2001) that most of the systems containing ex- invoking biases against high-amplitude orbits in the se-
oplanet candidates are seen nearly pole-on. These authors lection process of the radial-velocity-monitoring samples.
reached this conclusion by trying to extract the astromet- To the contrary, the planet-search surveys were specifi-
ric orbit, hence the orbital inclination, from the Hipparcos cally devised to avoid such biases, as they aim at find-
IAD. Halbwachs et al. (2000) had already cautioned that ing not only giant planets but also brown dwarfs so as to
this approach is doomed to fail for systems with apparent constrain the substellar secondary mass function of solar-
separations on the sky that are below the Hipparcos sensi- type stars. Furthermore, the Han et al. (2001) argument is
tivity (i.e. <
1 mas). In those cases, the solution retrieved totally invalid in the case of volume-limited, statistically
from the fit of the IAD residuals is spurious, since the true well-defined samples like that of the CORALIE planet-
angular semi-major axis a is simply too small to be seen search programme in the southern hemisphere (Udry et al.
by Hipparcos. Since Halbwachs et al. (2000) have shown 2000a). This sample has been specifically designed to
that a actually follows a Rayleigh probability distribution, detect companions of solar-type stars all the way from
the fit of the IAD residuals will yield a solution larger than q = M2 /M1 = 1 down to q 0.001.
A. Jorissen et al.: The distribution of exoplanet masses 997

Appendix: Non-parametric treatment of the data In the adaptive kernel version, a local bandwidth hn =
h(Xn , f ) is defined and used in Eq. (19). In order to follow
To decrease the noise and allow a tractable use of the in-
the true underlying function in the best possible way,
formation present in small data samples, heavy smoothing
the amount of smoothing should be small when f is large
techniques are often required. A common practice consists
whereas more smoothing is needed where f takes lower
converting a set of discrete positions into binned counts.
values. A convenient method to do so consists in deriving
Binning is a crude sort of smoothing and many studies in
first a pilot estimate f of f , e.g. by using an histogram or
statistical analysis have shown that, unless the smooth-
a kernel with fixed bandwidth hopt , and then by defining
ing is done in an optimum way, some, or even most, of
the local bandwidths
the information content of the data could be lost. This is
especially true when a large amount of smoothing is nec- hn = h(Xn ) = hopt [f(Xn )/s] , (21)
essary, which then changes the shape of the resulting
function. In statistical terms, the smoothing process not where
1 X
only decreases the noise (i.e., the functions variance), but N

at the same time introduces a bias in the estimate of the log s = n ).


log f(X (22)
N n=1
function.
The variance-bias trade off is unavoidable but, for a It may be shown (Abramson 1982) that = 1/2 leads to
given level of variance, the bias increase can be minimized. a bias of smaller order than for the fixed h estimate.
The correct manner of achieving that task is provided by The optimum kernel K may be taken as the one min-
the so-called non-parametric density estimate methods for imizing the integrated mean square error beween f and f
the determination of the frequency function of a given (MISE), where
parameter or by the non-parametric regression methods Z h i2
for the determination of a smooth function g inferred from MISE(f) = E f(x) f (x) dx (23)
direct measurements of g itself. Moreover, adaptive non-
parametric methods are designed to filter the data in some is usually taken as an indicator of efficiency. In the above
local, objective way minimizing the bias, in order to get expression, E denotes the statistical expectation. This
the smooth desired function without constraining its form optimum kernel is the so-called Epanechnikov (Ke ) or
in any way. The theory and algorithms related to those quadratic kernel:
methods, originally built to handle ill-conditioned prob-
3
lems (either under-determined or extremely sensitive to Ke (u) = (1 u2 ), |u| < 1. (24)
errors or incompleteness in the data), are widely discussed 4
in the specialized literature and summarized in easy-to- Other choices of K differ only slightly in their asymptotic
read textbooks (e.g., Silverman 1986; Hardle 1990; Scott efficiencies and could be more adapted to particular pur-
1992). poses as e.g., for bi-dimensional data.
The simplest of the available algorithms is provided by The pilot smoothing length (hopt ) is the only subjec-
the kernel estimator leading to the following form of the tive parameter required by the method. It relates to the
normalized frequency function quality of the sampling of the variable under considera-
  tion. There are several ways for automatically estimating
1 X
N
x Xn an optimum value of hopt (see e.g. Silverman 1986 for an
fK (x) = K (19)
N h n=1 h extensive review). A simple approach based on the data
variance gives in our case hopt ' 1.0 MJ . As the deriva-
with the normalization tive of the frequency function rather than the function
Z itself is actually used in Eq. (6), a larger pilot smoothing
K(u) du = 1, (20) length (2 hopt ) was also considered in order to remove spu-

rious small-statistics fluctuations of the density estimate.
where Xn (n = 1, ... N ) are the n available data points.
Each data point is thus simply replaced by a bump.
References
The kernel function K determines the shape of the bumps
while the bandwidth h (also called smoothing parameter) Abramson, I. S. 1982, Ann. Statist., 10, 1217
Butler, R. P., Marcy, G. W., Fischer, D. A., et al. 1999, ApJ,
determines their width2 .
526, 916
2
In the regression problem, the Nadaraya-Watson estimator Cerf, N., & Boffin, H. M. J. 1994, Inverse Problems, 10, 533
(Nadaraya 1964; Watson 1964) is commonly used: Chandrasekhar, S., & Munch, G. 1950, ApJ, 111, 142
Charbonneau, D., Brown, T. M., Latham, D. W., & Mayor,
X
N
x Xn M. 2000, ApJ, 529, L45
g(Xn )K( ) Halbwachs, J. L., Arenou, F., Mayor, M., Udry, S., & Queloz,
h
n=1
g(x) = D. 2000, A&A, 355, 581
X
N
x Xn Han, I., Black, D., & Gatewood, G. 2001, ApJ, 548, L57
K( ) Hardle, W. 1990, Applied nonparametric regression
h
n=1 (Cambridge University Press, Cambridge)
998 A. Jorissen et al.: The distribution of exoplanet masses

Lucy, L. B. 1974, AJ, 79, 745 Silverman, B. W. 1986, Density estimation for statistics
Lupton, R. 1993, Statistics in theory and practice (Princeton and data analysis, Monographs on Statistics and Applied
University Press, Princeton) Probability (London: Chapman and Hall)
Marcy, G. W., Butler, R. P., Fischer, D. A., Vogt, S., & Tabachnik, S., & Tremaine, S. 2001, AJ, in press
Lissauer, J. 2001, ApJ, 556, 296 Udry, S., Mayor, M. 2001, in Astrobiology, Lect. Notes Phys.
Mayor, M., Naef, D., Pepe, F., et al. 2000, in Planetary (Springer Verlag)
Systems in the Universe: Observations, Formation and Udry, S., Mayor, M., Naef, D., et al. 2000a, A&A, 356, 590
Evolution, ed. A. Penny, P. Artymowicz, A.-M. Lagrange, Udry, S., Mayor, M., & Queloz, D. 2000b, in Planetary Systems
& S. Russell, ASP Conf. Ser., in press in the Universe: Observations, Formation and Evolution,
Nadaraya, E. A. 1964, Theor. Probab. Appl., 10, 186 ed. A. Penny, P. Artymowicz, A.-M. Lagrange, & S.
Pourbaix, D. 2001, A&A, 369, L22 Russell, ASP Conf. Ser., in press
Pourbaix, D., & Arenou, F. 2001, A&A, 372, 935 Watson, G. S. 1964, Sankhya, Ser. A, 26, 359
Richardson, W. H. 1972, J. Opt. Soc. Am., 62, 55 Zapatero Osorio, M. R., Bejar, V. J. S., Martin, E. L., et al.
Scott, D. W. 1992, Multivariate Density Estimation 2000, Science, 290, 103
(Wiley, New York) Zucker, S., & Mazeh, T. 2001, ApJ, in press

You might also like