Full Text
Annual Review of Astronomy and Astrophysics
Vol. 35:
137-177
(Volume publication date September 1997)
(doi:10.1146/annurev.astro.35.1.137)
MODEL ATMOSPHERES OF VERY LOW MASS STARS AND BROWN DWARFS
France Allard,1
Peter H. Hauschildt,2
David R. Alexander,1 and
Sumner Starrfield3
1Department of Physics, Wichita State University, Wichita, Kansas 67260-0032; e-mail: allard@eureka.physics.twsu.edu ; dra@twsuvm.uc.twsu.edu
2Department of Physics and Astronomy, University of Georgia, Athens, Georgia 30602-2451; e-mail: yeti@hal.physast.uga.edu
3Department of Physics and Astronomy, Arizona State University, Tempe, Arizona 85287-1504; e-mail: sumner.starrfield@asu.edu
ABSTRACT
|
Section:
|
|
|
|
Abstract
As progressively cooler stellar and substellar objects are
discovered, the presence first of molecules and then of condensed
particulates greatly complicates the understanding of their physical
properties. Accurate model atmospheres that include these processes are
the key to establishing their atmospheric parameters. They play a
crucial role in determining structural characteristics by setting the
surface conditions of model interiors and providing transformations to
the various observational planes. They can reveal the spectroscopic
properties of brown dwarfs and help establish their detectability. In
this paper, we review the current state-of-the-art theory and modeling
of the atmospheres of very low mass stars, including the coolest known
M dwarfs, M subdwarfs, and brown dwarfs, i.e. Teff 4,000 K and 4.0 [M/H]
+0.0. We discuss ongoing efforts to incorporate molecular and grain
opacities in cool stellar spectra, as well as the latest progress in (a) deriving the effective temperature scale of M dwarfs, (b) reproducing the lower main sequences of metal-poor subdwarfs in the halo and globular clusters, and (c) results of the models related to the search for brown dwarfs.
The crop of extremely cool stars
and substellar objects has been meager until very recently, when marked
improvements in detection ability have finally started to yield a rich
harvest. Stars with masses as low as 0.1M and white dwarfs as cool as 5000 K (Paresce et al 1995, Richer et al 1995, Cool et al 1996, Renzini et al 1996)
have been resolved in nearby globular clusters using the Wide
Field/Planetary Camera 2 on board the Hubble Space Telescope (HST).
Charge-coupled device (CCD) astrometry has unveiled uncharted M dwarfs
in the immediate vicinity of our Solar System (Henry et al 1995, Stone et al 1996, Tinney 1996). Growing numbers of halo carbon dwarfs have been discovered by deep multicolor CCD surveys (Heber et al 1993, Warren et al 1993, Deutsch 1994, Liebert et al 1994). Young deuterium-burning brown dwarfs have been identified in nearby open clusters (Rebolo et al 1995, Basri et al 1996, Rebolo et al 1996) and in the solar neighborhood (Thackrah et al 1997)
using the Keck telescope. Cryogenic coronographic imaging and
astrometric surveys are revealing cool, evolved brown dwarfs hiding in
the solar neighborhood (Nakajima et al 1995, Oppenheimer et al 1995, Mazeh et al 1996, Williams et al 1997),
and high signal/noise radial velocity and astrometric surveys are now
sensitive to massive planets around nearby Sun-like stars (Mayor & Queloz 1995, Butler & Marcy 1996, Marcy & Butler 1996a, b, Butler et al 1997, Marcy et al 1997). These exciting developments, and those of the MACHO, EROS, and OGLE microlensing surveys (Aubourg 1995, Alcock et al 1996),
will soon enable a reconstruction of the faint end of the galactic
initial mass function and the determination of the baryonic fraction of
the missing mass (Chabrier et al 1996b, Flynn et al 1996, Graff & Freese 1996). Very low mass (VLM) stars and brown dwarfs are probably the most numerous objects in the galaxy (Gould et al 1996, Méra et al 1996a).
Nevertheless, until recently, little about their atmospheres,
evolution, or spectral characteristics was clearly understood. The
presence of both a wide variety of molecular absorbers (each with
hundreds of thousands to millions of spectral lines) and numerous
condensates greatly complicates accurate modeling of these cool stellar
atmospheres. The extension of the convection zone to the outermost
photospheric layers means that evolutionary models depend critically on
accurate handling of the surface boundary. Recently, theoretical
calculations of important molecular absorbers such as H2O,
TiO, CN, and CO, as well as grain opacities, have made it possible to
generate greatly improved models of cool stellar and brown dwarf
atmospheres, their high resolution spectra, and their evolution. In
view of the latest progress both on the observational and theoretical
fronts, it seems timely to summarize here our present understanding of
the atmospheres and spectroscopic properties of cool VLM stars and
substellar brown dwarfs. Previous reviews of these topics can be found
in Liebert & Probst (1987), Stevenson (1991), Bessell & Stringfellow (1993), Burrows & Liebert (1993), Gustafsson & Jřrgensen (1994).
GENERAL SPECTROSCOPIC PROPERTIES OF COOL DWARFS
|
Section:
|
|
|
|
A VLM star generally refers to a main
sequence star with a spectral type ranging from mid K to late M and a
mass from about 0.6 M to the hydrogen-burning minimum mass (0.075 0.085 M ,
depending on metallicity). Such stars span a wide range of populations,
from the youngest metal-rich M dwarfs in open clusters such as the
Pleiades (Simons & Becklin 1992, Hambly et al 1993, Williams et al 1996, Zapatero Osorio et al 1997), the Hyades (Leggett & Hawkins 1989, Reid 1993, Bryja et al 1994, Leggett et al 1994), Ophiuchus (Cameron et al 1993), Persei (Zapatero Osorio et al 1996), and the galactic disk (Gliese & Jahreiss 1991), to the several billion year-old metal-poor subdwarfs of the galactic halo (Green et al 1991, Monet et al 1992, Green & Margon 1994) and globular clusters (Paresce et al 1995, Richer et al 1995, Cool et al 1996, Renzini et al 1996).
Even within the solar neighborhood, M dwarfs do not form a homogeneous
sample of unique age and metallicity, but rather they span up to 103 years in age and ±0.5 1.0 dex around the solar value in metallicity (Burrows & Liebert 1993).
In VLM star and brown dwarf atmospheres, most of the hydrogen is locked in H2 and most of the carbon in CO, with excess oxygen bound in molecules such as TiO, VO, and H2O.
The energy distribution of a typical late-type M dwarf is entirely
governed by the absorption of TiO and VO in the optical, and H2O
in the infrared, leaving no window of true continuum in the emergent
spectrum. With their extremely low intrinsic faintness (10
2
10
5 L ), in particular in the V
bandpass, the painstaking spectral classification of the nearby stars
aiming toward a complete census of the luminosity function is still in
progress (Reid 1994, Kirkpatrick & Beichman 1995, Liebert et al 1995, Reid et al 1995a).
Fortunately, the groundwork necessary to construct an effective
temperature scale at the lower end of the main sequence has already
been laid by Boeshaar (1976) in the visual (0.44 0.68 m), with (a) the first identification of CaOH bands at 0.54 0.556 m
in dwarfs later than about M3.5 (these bands are excellent temperature
indicators and good discriminants between field M dwarfs and background
red giant stars); (b) the first report of a saturation of the visual TiO band strengths in M dwarfs later than M5; and (c)
the introduction of the VO to TiO band strength index now being used to
classify M dwarfs and substellar candidates later than M5 (Henry et al 1994, Kirkpatrick et al 1995, Martín et al 1996).
Boeshaar's classifications soon were extended beyond even the limits of
the classical Morgan & Keenan spectral sequence, i.e. to types M9.5 >M10, by Kirkpatrick et al (1995) in the optical to near-infrared regime (0.65 1.5 m) and by Davidge & Boeshaar (1993), Jones et al (1994), Leggett et al (1996) in the near infrared (1.1 2.5 m).
Figure 1
summarizes a typical near-infrared spectral sequence of M dwarfs to
brown dwarfs. The near-infrared water vapor bands become slowly
stronger with the spectral type of M dwarfs. The CO overtones near 2.3 m (and 4.5 m, not shown in Figure 1) are still apparent, although much weaker than in late-type giant stars due to the stronger H2O "continuum" in the dwarfs. At the hydrogen-burning limit, i.e. at a spectral type of M10.5 and a Teff of about 2000 K (Baraffe et al 1995), the peculiar spectral distribution of GD 165B (Zuckerman & Becklin 1992, Kirkpatrick et al 1993a)
suggests that all signs of the TiO bands disappear from the optical
spectral distribution, leaving only atomic lines and perhaps VO bands (Davis 1994, Kirkpatrick et al 1995), CaH, CaOH, and/or FeH bands. As the effective temperature drops into the brown dwarf regime, methane (CH4) features begin to appear (Tsuji et al 1995, Allard et al 1996, Marley et al 1996), and corundum (Al2O3), perovskite (CaTiO3), iron, enstatite (MgSiO3), and forsterite (Mg2SiO4)
clouds may form, enhancing the carbon:oxygen abundance ratio and
profoundly modifying the thermal structure and opacity of the
photosphere (Sharp & Huebner 1990, Fegley & Lodders 1996).
View larger version (59K)
|
Figure 1 A near-infrared spectral sequence of M dwarfs to brown dwarfs. The observed spectral distributions ( full lines) were obtained at UKIRT for the M dwarfs by Jones et al (1994), and for the brown dwarf Gl229B by Geballe et al (1996). A comparison to OS models with, from top to bottom, Teff = 3400, 3000, 2700, 2600, 2000, and 1000 K (F Allard & PH Hauschildt, in preparation) (dotted lines)
reveals a growing overestimation of water vapor band strengths with
decreasing mass. The peculiar optical spectrum of GD 165B forces an
arbitrary choice of the model parameters (here set to those of a star
at the hydrogen-burning limit) for this object.
|
|
The chemistry of cool dwarf
atmospheres is, therefore, a complex nonlinear problem requiring a
detailed knowledge of the concentration of atoms and molecules, which
prevents a straightforward derivation of quantities such as excitation
temperatures and metallicities from line ratios, as is possible for
hotter stars. The most reliable way to estimate effective temperatures
and metallicities of VLM stars and to identify substellar brown dwarfs
is by a direct comparison of observed and model spectra.
A BRIEF HISTORY OF THE MODELS
|
Section:
|
|
|
|
Advances in atmospheric modeling of cool stars have been slowed by the twin bottlenecks of (a) incomplete molecular opacity data bases and (b)
the inability to handle convection rigorously. Once these problems are
addressed reasonably well, we still face other challenges:
incorporating the effects of photospheric grain formation,
chromospheres, magnetic fields, departures from local thermodynamic
equilibrium, spatial variations in atmospheric structure due to
starspots, cloud formation, and eventually weather patterns. Model
atmospheres incorporating such processes have only become possible
within the past two decades with the work of Mould (1975, 1976), Allard (1990), Kui (1991), Brett & Plez (1993), Allard & Hauschildt (1995b), Brett 1995a, b, Tsuji et al (1996a) for M dwarfs; Saumon et al (1994) for zero-metallicity subdwarfs; and Tsuji et al (1995, 1996b), Allard et al (1996), Marley et al (1996) for substellar brown dwarfs.
Mullan & Dermot (1987) have reviewed early efforts in modeling M dwarf atmospheres. Mould (1975, 1976)
was the first to produce an extensive grid of convective M dwarf model
atmospheres between 4750 and 3000 K. The models effectively combined
the ATLAS code (Kurucz 1970), TiO band model opacities and chemical equilibrium by Tsuji (1966, 1973), H2O opacities by Auman (1967), and a mixing-length treatment of convection (Břhm-Vitense 1958, Kippenhan 1962). Mould also incorporated atomic line blanketing in the form of an Opacity Distribution Function (ODF; see Kurucz 1970, Mihalas 1978).
However, the coarseness of his opacity grid kept him from adequately
reproducing the observed spectral characteristics of the coolest M
dwarfs.
It took another 15 years before model calculations finally broke the "3000-K barrier" in Teff, with the work of Allard (1990), Kui (1991). Both adapted their model codes from that of Wehrse (1972), who had treated the more extreme atmospheric conditions of cool white dwarfs (Teff
7000 K). Both authors also handled molecular opacity using band models
and straight mean (SM) techniques that made it possible to include beyond the dominant TiO and H2O opacities a
number of important molecular bands such as those of the hydrides (CaH,
MgH, SiH, OH, CH), which are important in low-metallicity subdwarfs, as
well as the red and infrared bands of VO (Keenan & Schroeder 1952) and CO, respectively, which act as sensitive temperature indicators (Henry et al 1994, Kirkpatrick et al 1995, Martín et al 1996). From the Allard (1990) grid, Kirkpatrick et al (1993b)
derived a revised temperature sequence for M dwarfs that casts new
light on traditional results based on blackbody methods. This new
sequence yielded values of Teff as much as 500-K higher at a
given luminosity and shifted the positions of the late-type dwarfs in
the HR diagram from cooling tracks to the blue side of theoretical
lower main sequences (D'Antona & Mazzitelli 1985, Burrows et al 1989, 1993).
This made it more likely that field late-type M dwarfs were
hydrogen-burning stars rather than young, contracting, substellar brown
dwarfs. Subsequent improvements to these models such as the introduction of (a) laboratory oscillator strengths for the TiO bands (Davis et al 1986) instead of the smaller (by a factor of 2 3) empirically derived astrophysical values of Brett (1989, 1990) and (b) the FeH Wing-Ford bands near 0.98 m (Phillips et al 1987) allowed Allard (1994) to resolve most of the remaining discrepancies in the optical model spectra that had been pointed out by Kirkpatrick et al (1993b), Gustafsson & Jřrgensen (1994), Jones et al (1994).
Despite the initial successes, comparison with
observed near-infrared spectra uncovered another problem: The models
failed to match the infrared spectrum governed by the water vapor
opacity profile (Allard & Hauschildt 1995b, Bessell 1995, Tinney et al 1995). This situation is illustrated in Figure 1,
which shows that the water bands are clearly too strong in the
metal-rich models. The peak of the energy distribution of M dwarfs is
located in the near infrared, at around 1 m. For brown dwarfs, most of the emitted flux emerges between 1 and 10 m.
One difficulty in determining the quality of model spectra is due to
telluric absorption in the Earth's atmosphere. Telluric water bands
filter the light of these faint objects over most of the infrared
range. While some near-infrared spectra from about 0.9 2.5 m can be obtained from ground-based facilities (e.g. the UKIRT spectra of Figure 1),
these are unreliable in intervals where the water bands are strongest.
A proper calibration of the measured fluxes becomes even more delicate
for faint brown dwarfs in close binary systems (for an illustration of
the uncertainties in calibrating the "K" band fluxes in the spectrum of
Gl 229B, see e.g. Oppenheimer et al 1995, Matthews et al 1995, Geballe et al 1996). Beyond 2.5 m,
the Earth's atmosphere is nearly opaque and red dwarfs must be observed
with infrared space-based facilities such as the HST, NICMOS, ISO, and
the planned SIRTF, NGST, and DARWIN missions. But while there remain
uncertainties in the absolute calibration of ground-based
spectrophotometry of faint M dwarfs, these cannot completely account
for the observed flux discrepancy in the infrared spectra of M dwarfs.
For example, Figure 1 indicates that the predicted H2O bands grow in strength more rapidly with decreasing Teff than those of observed M dwarfs (Kirkpatrick et al 1995).
This comparison supports the conclusion that there are shortcomings in
the models. One of those shortcomings is clearly the treatment of
opacity in very cool atmospheres.
MOLECULAR OPACITIES
|
Section:
|
|
|
|
Most of the molecules that play an
important role in cool star atmospheres have been known since the early
1930s from the work of Russell (1934) and later De Jager & Neven (1957). Some of the most extensive studies of cool stellar atmosphere chemistry are by Vardya (1966), Morris & Wyller (1967), Tsuji (1973), Gurvich (1981)
who published equilibrium constants for an extensive list of diatomic
and polyatomic species. More recent studies such as those of Sauval & Tatum 1984, Rossi et al (1985), Irwin (1987, 1988), Cherchneff & Barker (1992), Neale & Tennyson (1995), Sharp & Huebner (1990) provide partition functions for most molecules directly, which allows for more flexible atmospheric calculations.
In the absence of detailed lists of transitions, or
sometimes to cope with restricted computational facilities, atmospheric
modelers often resort to band models or to average opacities such as
the Just Overlapping Line Approximation (JOLA), SM, or ODF techniques,
which approximate (by a continuum distribution) the absorption within a
band or a predefined wavelength bin (Kurucz 1970, Mihalas 1978, Tsuji 1994).
While computationally economical, they make the assumption that the
rotational fine structure is smeared out; i.e. the lines overlap
without being saturated. Such conditions are never truly met even for
the strongest bands of TiO and H2O in the densest of the VLM
stellar atmospheres, and these methods tend to overestimate the
resulting molecular blanketing by trapping photons that would have
otherwise escaped from between the lines. A far more accurate account
of molecular and atomic opacities in model atmospheres is achieved by
applying an Opacity Sampling (OS) treatment of transitions lists on a
prespecified fine grid of wavelengths (Peytremann 1974, Sneden et al 1976). This can be done either dynamically within the atmospheric calculations (Kurucz 1992b, Hauschildt et al 1992, Allard & Hauschildt 1995b) or be pretabulated as a function of pressure, temperature, isotopic ratios, and wavelengths (Plez et al 1992, Brett 1995a, Kipper et al 1996).
While the advantage of a dynamical approach lies in the flexibility of
handling depth-dependent mechanisms such as pressure broadening,
departures from local thermodynamic equilibrium, microturbulence, and
abundance variations, the more efficient pretabulation of the OS
opacities gives the modeler freedom to incorporate his or her choice of
complete line lists.
In an attempt to address the too-strong infrared water band problem in M dwarf models, Brett & Plez (1993), Allard et al (1994, 1996), Brett (1995a, b),
and F Allard & PH Hauschildt (in preparation) used the OS treatment
of molecular opacities to compute a new generation of M dwarf model
atmospheres, which brought important breakthroughs in the understanding
of M dwarf atmospheres. In the next sections, we summarize the most
significant improvements in the treatment of opacities due to TiO and H2O.
Optical Bands
The strengths of TiO bands define the optical (0.4 1.2 m)
spectral distribution of late K to M stars. Together with the VO bands
and a few other optical spectral features, they constitute the primary Teff indicators in very cool stars. There currently exist three TiO line lists generated (a) from first principles by Collins & Fa˙ (1974) and more recently extended for isotopic species and the system by Jřrgensen (1994), (b) empirically from molecular levels assigned in laboratory experiments by Kurucz (1993), and (c) by Plez et al (1992). While substantial errors in the Kurucz (1993) line list have been acknowledged by the author, the Plez et al (1992), Jřrgensen (1994) line lists lead to great improvements upon previous models based on SM treatment of opacities (Mould 1975, Kui 1991, Allard & Hauschildt 1995b)
in the modeling of M dwarfs. Each TiO line list applied in an OS
treatment of the opacities leads to better agreement with the observed
optical absolute magnitudes of M dwarfs (see e.g. Brett 1995a, b, Chabrier et al 1996a).
The new models also show excellent agreement with the measured
parameters of the only two known M dwarfs in eclipsing binaries (see
Section 9 below; also see Bessell 1991, 1995, Chabrier & Baraffe (1995).
Unfortunately, the TiO line lists of Plez et al (1992), Jřrgensen (1994) give poor line positions and relative band strengths that prevent accurate high-resolution spectral syntheses of M dwarfs (Piskunov et al 1996, Schweitzer et al 1996). They also fail to reproduce the optical R-I
colors of late-type dwarfs (F Allard & PH Hauschildt, in
preparation), which may reflect either some remaining inaccuracies in
the current estimates of the oscillator strengths (Davis et al 1986, Doverstal & Weijnitz 1992, Hedgecock et al 1995)
or an incomplete account of VO or other opacity in the "R" bandpass.
Indeed, despite the existence of a few spectroscopic studies of VO
systems (Davis 1994, Merer et al 1987, Bauschlicher & Langhoff 1986),
no list of transitions and oscillator strengths adequate for stellar
atmosphere modeling is yet available for this important molecule. The
Berkeley program has generated extensive line lists for FeH (Phillips et al 1987, Phillips & Davis 1993; see also Balfour & Klynning 1994),
which, however, lack matching oscillator strengths. Moreover, the
complexity of the FeH molecule has prevented theoretical models (Langhoff & Bauschlicher 1990, 1991) from reproducing the observed spectrum of FeH (Langhoff & Bauschlicher 1994). A similar situation also prevails for the electronic systems of CaOH (Bernath & Brazier 1985, Ziurys et al 1992, 1996),
despite their importance as one of the strongest visual bands in the
spectra of M-type dwarfs. Modelers have resorted to band models for
most of these molecular systems (Brett 1989, 1990, Brett & Plez 1993, Allard & Hauschildt 1995b),
which overestimate the resulting opacity and compromise both
high-resolution spectral analysis and the determination of accurate
atmospheric parameters. Fortunately, a new ab initio calculation of TiO
is currently under way (SR Langhoff & CW Bauschlicher, Jr, in
preparation), which should soon enable improved modeling of some
aspects of cool M dwarfs.
H2O Bands
In view of the initial success obtained with an
OS treatment of the TiO opacities for the optical spectral distribution
of M dwarfs, Alexander et al (1989) and later Plez et al (1992) developed an OS table of randomly distributed H2O lines derived from line strength and line spacing data measured in the laboratory Ludwig 1971. (Brett & Plez 1993, Brett 1995a, b) then used the Plez et al (1992) table in their models of M dwarfs, but this treatment still failed to reproduce the infrared spectra and colors of M dwarfs (Bessell & Stringfellow 1993, Bessell 1995).
In retrospect, this result was to be expected because H2O lines overlap more than those of TiO, so SM treatment of opacities is more appropriate for H2O. Schryber et al (1995) therefore argued, based on results of their ab initio calculations for H2O, that the H2O laboratory cross sections obtained by Ludwig (1971), used by both groups in the form of either SM (Allard & Hauschildt 1995b) or OS (Plez et al 1992, Brett & Plez 1993, Brett 1995a, b)
may be intrinsically overestimated when applied to gas hotter than
about 1500 K. Theoretical lists of transitions that include "steam" or
"hot" band transitions based on molecular levels assigned in laboratory experiments (semiempirical; e.g. Kurucz 1992a) or on a molecular model from first principles (ab initio; e.g. Miller et al 1994) are
of far greater relevance for atmospheric calculations and are essential
for an adequate account of molecular opacities in cool star and brown
dwarf atmospheres. Over the past decade, efforts have converged in the
development of improved theoretical opacity data for molecules of
astrophysical interest with the creation of the Kurucz (1992a) and SCAN (Jřrgensen 1992) data bases, and with the fruitful work of the University of the College of London (Miller et al 1994) and NASA Ames (Langhoff & Bauschlicher 1994) centers of quantum chemistry calculations.
Theoretical line lists for hot H2O from three independent sources have recently been released by Jřrgensen et al (1994), Miller et al (1994), Partridge & Schwenke (1997). The Miller et al (1994) list (6.2 million lines) uses a laboratory potential surface (Jensen 1989), while the Partridge & Schwenke (1997)
list (300 million lines) uses a purely theoretical potential but the
same computational approach. Both preliminary lists were computed up to
J values of about 30; i.e. they do not include all the necessary hot or steam bands. The Jřrgensen et al (1994) list (20 million transitions) on the other hand, while also based on the Jensen (1989)
potential energy functions, was computed with the goal of completeness
for the atmospheres of cool giants with some compromise on the
treatment of the molecular binding. For example, they use a rigid
rotator approximation with an a posteriori correction to the
Hamiltonian. The three data sets lead to very different opacity
profiles, with the Jřrgensen et al and Partridge & Schwenke lists
reproducing the results obtained previously with the Ludwig opacities.
Only the Miller et al line list led to an improved fit of the infrared
spectral distribution of M dwarfs (Allard et al 1994, Jones et al 1995, 1996, Leggett et al 1996),
as well as to an excellent agreement of early-type M dwarfs with a
whole new generation of evolutionary models that include improved
non-gray surface boundary conditions (Baraffe et al 1995, 1997, Chabrier et al 1996a, Leggett et al 1996). However, none of the current H2O
line lists can explain the apparent saturation of the water vapor bands
observed in the latest-type M dwarfs and illustrated in Figure 1.
The cause of those discrepancies may therefore lie elsewhere, as is
discussed in Section 5 below. A more accurate knowledge of the water
vapor opacity profile is clearly needed and is now being addressed by
the work of Viti et al (1995), Partridge & Schwenke (1997).
The current generation of M dwarf model atmospheres (Brett 1995b, Allard et al 1996)
does not include the condensation of molecules to grains. Condensation
clearly must be included in the calculations as indicated by the work
of Sharp & Huebner (1990), who report the abundance of condensates as a function of the gas conditions. If ZrO2
one of the first condensates to appear at gas temperatures 2000 K is
not an important species in M dwarf atmospheres, the condensation of
corundum at 1800 K and iron, VO, and enstatite at 1600 K most certainly
affects the spectral distribution of late M dwarfs and brown dwarfs
because of the large extinction of solid particles. The importance of
condensation in the atmospheres of late-type M dwarfs and brown dwarfs
has been confirmed by Tsuji et al (1996a, b), Fegley & Lodders (1996), who find large concentrations of such condensates in their model atmospheres.
The impact of condensation on the spectral
distribution and atmosphere of a cool dwarf is to gradually deplete the
gas phase abundance of titanium, iron, vanadium, and oxygen. If we
ignore for the moment the opacity of the grains, the result is a more
transparent optical spectral distribution because the TiO-, VO-, FeH-,
and metal-line opacities decline with decreasing effective temperature
of the star. This should be reflected by an observed saturation of
these molecular bands in the latest-type M dwarfs and brown dwarfs, a
behavior that is presently difficult to ascertain without accurate
model atmospheres that incorporate the effects of condensation. Perhaps
a confirmation can be found in the peculiar optical spectrum of the
coolest known M dwarf, GD 165B, mentioned in Section 2 above. However,
the true nature of GD 165B's atmosphere is uncertain because this
object, the companion of an old pulsating DA white dwarf within an
orbital distance of 128 AU (Becklin & Zuckerman 1988, Bergeron & McGraw 1990, Zuckerman & Becklin 1992, Bergeron et al 1993, Kirkpatrick et al 1993a), may be more metal-poor and/or more carbon-rich than other nearby stars as a result of the white dwarf's prior evolution.
Tsuji et al (1996a)
were the first to calculate model atmospheres for M dwarfs and brown
dwarfs including not only grain formation but also grain opacities, the
so-called dusty models. Their results showed that including corundum,
iron, and enstatite opacities, while assuming arbitrarily spherical
grains with sizes set to 0.1 m,
could heat the photospheric layers and change the overall structure of
the atmosphere. The resulting dusty spectral distributions of late-type
M dwarfs were redder with weaker molecular spectral features than
models without grain opacities, and they were shown to reproduce the
infrared broadband fluxes of the latest-type M dwarfs, including GD
165B. If confirmed, this greenhouse effect, caused by the presence of
photospheric grains, may help explain the observed saturation of the
near-infrared water vapor bands discussed in Section 4 and illustrated
in Figure 1, as well , as well as perhaps the R-I colors (see Section 4.1) of late-type dwarfs, which the grainless models of Allard & Hauschildt (1995b), Brett (1995a, b) fail to reproduce. The calculations presented by Tsuji et al (1996a),
however, are coarse, and a better treatment of both the molecular and
grain opacities, as well as the formal inclusion of dust scattering in
the solution of the radiative transfer equation, can be achieved.
Early attempts to compute the opacity of grains were made by Cameron & Pine (1973), Alexander 1975.
More detailed calculations including the effects of chemical
equilibrium calculations and grain-size distributions were reported by Alexander et al (1983), Pollack et al (1985). Alexander & Ferguson (1994a, b)
have described the computation of the opacity of grains with the
inclusion of equilibrium condensation abundances, the effects of the
distribution of grain sizes, and the effect of grain shape through the
continuous distribution of the ellipsoid model of Bohren & Huffman (1983).
These calculations include the absorption and scattering due to
magnesium silicates, iron, carbon, and silicon carbide grains for a
wide range of chemical compositions down to temperatures of 700 K. The
direct inclusion of the equilibrium calculations of Sharp & Huebner (1990)
in the future will allow for more detailed treatment of the effects of
trace condensates, lower temperature opacity sources, and the effects
of different elemental abundances. The inclusion of high-temperature
condensates such as Al2O3 and CaTiO3
may have significant effects on the opacity in cool star atmospheres,
even though their abundance is quite small because of the high
absorption and scattering efficiency of grains. For lower temperatures,
the optical effects of species such as FeS, Fe3O4, and H2O need to be included. Pollack et al (1994)
have produced opacities for water, ammonia, methane, and other
low-temperature condensates. They assume complete condensation of all
condensible species and extend the temperature range down to 300 K.
These opacities offer an excellent basis for future brown dwarf and
Jovian-type planet atmosphere calculations. However, the extinction
caused by grains in a stellar atmosphere depends critically on the rate
of grain formation and the resulting size distribution.
Moreover, constraints imposed by the lack of detection of cloud layers in Jupiter by the Galileo atmospheric probe (Isbell & Morse 1996, Keane et al 1996), and of any trace of scattering by grains in the evolved brown dwarf Gl 299B (Allard et al 1996, Tsuji et al 1996b),
may imply an inhomogeneous vertical and/or horizontal distribution of
the grains, such as scarce cloud distribution, gravitational settling,
and sedimentation and rains of condensates in substellar dwarf
atmospheres. While grains are likely to be destroyed by the radiative
and convective heat in the inner layers of the atmosphere, the main
effects of the sedimentation and rains of condensates should be a
radial abundance gradient (Muchmore 1987, Guillot et al 1994) and a gradual depletion of the upper photosphere from its condensible elements over time.
The effect of grain formation and of its opacity on
the atmospheric structure of M dwarf atmospheres will, therefore, not
be fully understood until grain formation and time-dependent grain
growth calculations incorporating the effects of sedimentation,
diffusion, coagulation, and coalescence are included. Gail & Sedlmayr (1988), Dominik et al (1989) (see references therein) have developed a formalism to account for the phenomena in the outflows from cool giants (Beck et al 1992), supergiants (Seab & Snow 1989), and nova atmospheres (Beck et al 1995). Grain growth models have also been developed for the atmospheres of cool carbon-rich white dwarf (Zubko 1987) and Jovian planet (Rossow 1978, Dobrijevic et al 1992) atmospheres. However, as yet, no results have been obtained for oxygen-rich dwarf atmospheres.
The contribution of atomic and ionic
line transitions to photospheric opacities is relatively less important
for M dwarfs than for cool giants and hotter stars. This result arises
not only from the fact that molecular absorption bands dominate
opacity, but also because the lower photospheric temperatures cause the
number densities (Ni
) of atoms in higher excitation and ionization levels, such as those of the hydrogen Balmer series, to be quenched (Ni
e
i
/kT
, where T is the gas temperature and
i
is the excitation potential or ionization energy relative to the
ground state). Moreover, the "locking" of elements into molecular
compounds and further condensation of such elements to grains also
reduces the available abundances of atomic species, as is the case for
hydrogen, which is about 70 85% H2 in the photospheres of M dwarfs.
As a result, only the strongest
resonance and subordinate lines, with the low excitation energy of
mostly alkali and earth-alkali elements, prevail in the spectra of M
dwarfs. Those lines can be very broad owing to van der Waals (vdW)
pressure broadening, and they often contrast greatly with the narrow
emission and weak absorption lines that originate in the chromospheric
layers of active M stars (such as the Balmer series and the Ca H and K
lines). Only a few of the atomic lines that are created in the
photospheric layers can be detected within the haze of molecular lines
and provide diagnostics of the photospheric parameters. Examples
include the Na I D lines at 5889,5896 Ĺ, as well as other Na I resonance transitions at 8183,8195 Ĺ and 10746,10749,10835 Ĺ and those of K I at 6911,6939, 7665,7699, 9950,9954, and 10480,10482,10487 Ĺ. Lines of Rb I at 7950 Ĺ and Ba I at 7911,7913 Ĺ are also particularly strong (relative to the local continuum) in late-type M dwarfs and brown dwarf candidates.
Despite the relative scarcity of
directly observable atomic lines in their spectra, an accurate modeling
of M dwarf atmospheres nevertheless requires the use of a complete
atomic line list that includes lines of ionized elements for a complete
account of the opacity in the hotter layers (typically about 8000 K in
M dwarfs) of the inner atmosphere. A failure to do so may result in
atmospheric structures that are too cool globally, as the efficient
convection zone assures the transfer of inner atmospheric heat to the
outer photospheric layers. The most complete list of atomic transitions
currently available is that by Kurucz (1994) and its revisions. Several other line lists, such as those generated from first principle model atoms of the Opacity Project (Seaton 1992, Seaton et al 1992) or semiempirically using atomic levels assigned in laboratory experiments (see Verner et al 1996 and revisions), are also available but are still too incomplete for the purpose of model atmosphere calculations.
Line Broadening Mechanisms
The high densities prevailing in VLM star
atmospheres cause strong spectral lines to be significantly broadened.
Because the gas temperatures are not high enough to sustain a
significant amount of ionization, the electron and proton densities are
much smaller than the densities of the most important neutral and
molecular species. Consequently, the contribution of Stark broadening
to the total damping constant is very small, even in stars with very
low metallicities. The total thermal plus microturbulent line widths
are always much smaller than the line width owing to vdW broadening:
which describes the interaction
between two different, unpolarized neutral particles within the impact
or static approximation, with
vdW the full-width half-maximum damping constant of the resulting Lorentz profile, v the relative velocity between perturber and absorber, and Np
the number density of perturbers. While the interaction constant C
6 can be determined exactly for both the ground and excited states of a perturbed atom when the perturber is atomic hydrogen (Michelis 1976),
no exact method has yet been developed for the case of collisions with
the much slower molecular hydrogen perturbers that dominate the
atmospheres of VLM stars, brown dwarfs, and Jovian-type planets (Guillot et al 1994). In those cases, the collisions are not instantaneous and the profiles not strictly Lorentzian (Kunde et al 1982, Goody & Yung 1989), but in the absence of accurate alternatives, modelers often resort to using the hydrogenic approximation formulated by Unsöld (1955) for collisions with neutral hydrogen with some ad hoc modifications:
where Z is the charge of the absorber, E the ionization energy (e.g E
H = 13.6 eV), and El
and Eu
the lower and upper level excitation energies of the absorber. Investigations by Weidemann (1955),
for instance, showed that the values as calculated above are in good
agreement with observed line widths for alkali metals but not for other
elements, such as iron (Kusch 1958). This has led to the introduction of correction factors to the "classical" formula, which can range from 10 C
6 in the Sun (Takeda et al 1996) to 101.8 C
6 in white dwarfs for non alkali-like species (Wehrse & Liebert 1980). No corrections are required for alkali elements. The Unsöld (1955)
approximation, combined with this correction factor for non-alkali
elements and with an explicit account of the different polarizabilities
of each perturber (
p/
H) C
6, where the subscript p refers to the perturber),
leads to improved profiles that appear to describe well the atomic
lines observed in late-type M dwarfs (Schweitzer et al 1996). The most abundant perturbers in M dwarfs and their polarizabilities are given by Weast (1988), Schweitzer et al (1996).
While the situation is poor for atomic line
broadening, it is even worse for molecular lines, for which only a few
sources and techniques exist (Lazarev & Pnomarev 1992, Kurucz 1993, Guillot et al 1994).
Fortunately, individual molecular lines are usually not saturated, so
that broadening is less important for them than for strong atomic
lines. Moreover, molecular lines often overlap so strongly that their
wings are completely masked (Schweitzer et al 1996),
and only the Gaussian line cores of the strongest molecular transitions
are observed. The atmospheres of VLM stars and brown dwarfs are
therefore only weakly sensitive to the adopted value of the vdW damping
constant in the bands of several of the most important molecular
absorbers (e.g TiO and H2O). This may, however, not be the
case for some hydride bands and for the infrared CO overtones that show
larger typical line spacings (Kui 1991, Davis 1994, Tsuji 1994).
The thin radiative skin above the
convective region in an M dwarf determines the surface boundary
conditions for the entire temperature structure of the fully convective
photosphere and interior. This radiative zone is often limited to the
outermost optically thin regions of the photosphere in early-type M
dwarfs (Allard 1990, Kui 1991, Burrows et al 1993, Allard & Hauschildt 1995b): i.e. to optical depths below about 10
3. Figure 2
illustrates how the outer atmosphere of a typical M dwarf is affected
by the atomic and molecular opacities and convection. At such low
optical depths ( = 10
3 corresponds to log P
gas
3.8 in this model), the structure of the atmosphere is sensitive to the strong opacities of TiO and H2O.
Early-type M dwarf atmospheres, spectra, colors, and even their
evolution are, therefore, very dependent upon elemental abundances and
the treatments of molecular opacities and possibly convection (see
Section 9 below; see Baraffe et al 1995
for an illustration of these effects). Early-type M dwarfs should serve
as excellent stellar laboratories in which to study convection.
View larger version (49K)
|
Figure 2 The influence of the molecular and atomic
opacities and convection upon the atmospheric structure of a typical
model atmosphere; here the T
eff
= 2800 K, log g = 5.0, and solar metallicity model of Allard & Hauschildt (1995b). A corresponding gray structure without convection (bold dot-dashed) is also shown for comparison. While the complete neglect of H2O opacities causes a dramatic cooling (by CO) of the atmosphere (long-dashed curve), uncertainties by a factor of two in the H2O opacity cross sections cause only negligible changes in the atmospheric structure (thin dot-dashed relative to dotted curve). A similar drop in the opacity cross section of TiO, however (thin short-dashed relative to dotted curve), causes a much more significant cooling of the atmosphere.
|
|
The standard mixing length theory (Břhm-Vitense 1958, Kippenhan 1962, Mihalas 1978)
used to model convective energy transport in stars is only a crude
approximation. While nonlocal treatments of convection exist (for
review, see Chan et al 1991, Gustafsson & Jřrgensen 1994, Grossman 1996, Kim et al 1996)
that may be better suited to the optically thin medium of cool stellar
atmospheres, they are very computationally prohibitive and have not
been applied to models of M dwarfs. Fortunately or sadly, depending on
your point of view, the large opacities in M dwarfs mean convection is
nearly adiabatic for values of the mixing length ( )
comparable to the atmospheric pressure scale height. The atmospheres
and synthetic spectra of M dwarfs therefore show very little
sensitivity to changes in over the range typical of solar-type atmospheres; i.e.
HP
= 1.2 to 2.2 (Brett 1995a, Baraffe et al 1997).
Moreover, models indicate that the convection zone gradually retreats
with decreasing mass in late-type M dwarfs because of their decreasing
luminosity and with decreasing metallicity due to decreasing
photospheric opacities (Allard 1990).
In cool brown dwarf models such as those computed for Gl 229B, for
example, the convection zone reaches no higher than optical depths of
about unity (although models show signs of a second, separate
convective layer closer to the surface in such cool objects). The
spectroscopic and photometric properties of late-type M dwarfs,
metal-poor subdwarfs, and possibly brown dwarfs are therefore
relatively insensitive to the details of convection (see e.g. Brett 1995a).
This means that standard mixing length approximations are probably
suitable for these stars (which is good news for brown dwarf modelers),
but that these same stars are not very good laboratories to study
convection (which is bad news for convection modelers).
STELLAR ACTIVITY
|
Section:
|
|
|
|
The "solar dynamo" model, also known
as the alpha-omega dynamo, which operates at the radiative convective
boundary layer in the solar interior, predicts a correlation of
activity with rotation that is observed in solar-type stars (Noyes et al 1984, Marilli et al 1986, Rutten 1986).
When coupled with the decrease of the rotation rate as the star ages
(owing to loss of angular momentum during its lifetime), a
rotation-activity-age correlation is expected and has also been
observed (Wilson & Skumanich 1964). The fully convective lowest mass stars, however, like pre main-sequence
solar-type stars, are known to be very active even if no spot-cycle
variability can yet be confirmed in any of them. Several flare
periodically [see e.g. Linsky et al (1995) for a report of recent flare outbursts in VB10], and a large fraction of them (up to 60% in M5 dwarfs) show chromospheric H
emission. The dynamo generation of their fields must therefore
occur from a different, or at least modified, mechanism. And indeed,
the surface activity in the M dwarfs has been observed to exhibit
general characteristics that contrast with those of solar-type stars:
1. The incidence of chromospheric and coronal activity in M dwarfs
grows with decreasing stellar mass (Joy & Abt 1974, Giampapa & Liebert 1986, Reid et al 1995a, b,
Hawley et al 1996). 2. The coronal and chromospheric luminosity and the
luminosity that occurs in flares all decrease with stellar mass (Fleming 1988, Peterson 1989); however, the fraction of the luminosity that appears in these magnetic indicators relative to the total stellar luminosity (LH
/Lbol, LX/Lbol, hereafter "activity level") remains nearly constant (Fleming et al 1995, Reid et al 1995a, Mullan & Fleming 1996).
3. While, as in solar-type stars, M dwarfs show little or no activity
when the rotation rate reaches below a certain threshold ( 5 km/s; cf Marcy & Chen 1992),
stars with rotation rates above this threshold show weak or no
correlation between rotation and the activity level for both
chromospheric and coronal emissions (Stauffer & Hartmann 1986, Rutten et al 1989,
Hawley et al 1996). 4. Coronal and chromospheric activity show a
correlation with scale height from the galactic plane, metallicity, and
probably age of the star (Fleming et al 1995, Reid et al 1995a, Hawley et al 1996).
M dwarfs are therefore more active than solar-type
stars. This clearly indicates that some change in the magnetic field
generation and/or the interaction of the field with the stellar
atmosphere has occurred. Two possible ideas have been suggested to
explain the differing M dwarf behavior: 1. Noyes et al (1984) and later Peterson (1989)
pointed out that the volume of the convection zone is an important
parameter in the generation of the magnetic field, and this volume
begins to decrease in proportion to the mass once the stars are mostly
convective (in M dwarfs). The field strength may become saturated in
the lowest mass stars, and hence no strong rotation-activity connection
would be expected (see e.g. Rosner et al 1985, Stauffer et al 1991).
Moreover, the relative neutrality of the gas in M dwarfs compared with
hotter stars may also alter the behavior of magnetic field lines (P
Ulmschneider, private communication). Direct measurements of the
magnetic field strength and its stellar surface coverage have been
obtained, using Zeeman splitting of highly magneto-sensitive lines, by Robinson (1980), Gray (1984), Saar (1988), Mathys & Solanki (1989), Basri & Marcy (1994) for a number of late-type stars, which confirm the presence of strong magnetic fields in M dwarfs. Saar (1994), Johns-Krull & Valenti (1996), for example, report magnetic field strengths for the dMe dwarfs AD Leo, EV Lac, and AU Mic of 4.0 4.3
kG with covering factors between 55 and 85%. 2. On the other hand, the
propagation of acoustic shocks (which originate at the
convective-radiative surface boundary), and the resulting acoustic
heating of the chromosphere, should become most efficient in the
strongly convective M dwarf photospheres and may therefore play an
important role in the energy budget of their atmospheres and coronae
(Schrijver 1987, Mathioudakis & Doyle 1992, Mullan & Cheng 1993).
Since the extension of the convection zone depends sensitively on the
atmospheric parameters (see Section 7 above), a correlation of the
chromospheric activity with metallicity and age in M dwarfs would
therefore be likely. The first steps in integrating detailed
photosphere models with acoustically heated M dwarf chromospheres were
taken by Buchholz (1995), Mullan & Cheng (1993, 1994). Mullan & Cheng (1994)
report a more effective penetration of acoustic waves into the coronae
of M dwarfs compared with the case of more massive solar-type stars.
They find that acoustic heating can maintain a corona with a
temperature on the order of 0.7 1 106 K and a surface X-ray flux as large as 105 ergs cm
2 s
1,
and they suggest that relatively inactive M dwarfs that display X-ray
emissions below this limit may be candidates for acoustically
maintained coronae.
However, these ideas still leave some unanswered
questions. How can we explain, for example, stars of the same age or
the same mass and rotation rates above the threshold but with widely
different activity levels? In an attempt to explain the dilemma raised
by the two M9.5-type field dwarfs PC 0025+ 0447 (Graham et al 1992) and BRI 0021 012 (Tinney 1993) that have widely different activity levels, Basri & Marcy (1995)
suggested that there could also be a threshold temperature below which
acoustic and Alfvén waves become inefficient and stellar activity
subsides. Below this temperature threshold, the heating of the
chromosphere and corona and the wind generation would be sufficiently
prohibited to prevent the formation of chromospheric emission lines and
to slow the rotation of the star. However, this interpretation still
leaves the case of hotter stars unaddressed. An example is the triple
VLM star system LHS 1070 (Leinert et al 1994;
F Allard et al, in preparation), in which the two faint companions of
the same age and metallicity are also similar both in mass ( 0.085 M ) and Teff (2600 2700
K), but where only the faintest of the three components is inactive.
Several other systems have been observed where only the more massive
primary star is active (Hawley et al 1996). While most cases can be
explained if the primaries are in fact unresolved close-binary stars in
which enhanced rotation (and activity) is maintained by tidal
interactions, this does not seem to explain the lack of activity in the
close-binary star LHS 1070C.
Modeling an M dwarf chromosphere is a complex
problem owing to the complexity of the radiative transfer calculations
in such a cool, dense environment. Cram & Mullan (1985) modeled an M dwarf chromosphere using hydrogen Balmer emission line observations. Giampapa et al (1982) used Ca II observations to model M dwarf chromospheres, with limited success. Houdebine & Panagi (1990)
investigated the effects of changing the model hydrogen atom used in
the calculations, but they did not fit their models to data. Hawley & Fisher (1994)
have developed chromospheric flare models, which incorporate a full
nonlocal thermodynamic equilibrium (NLTE) treatment of the statistical
equilibrium and radiative transfer in the important optically thick
chromospheric lines and a helium ionization equilibrium computed
self-consistently with the downward X-ray flux from the corona. Yet
they were unsuccessful at fitting both the Ca II and hydrogen Balmer
lines in their quiescent and flare observations. More recently, Mauas & Falchi (1996)
were also unable to match both the observed hydrogen Balmer line
strengths in their quiescent model of a well-observed active M dwarf.
To date, no model has successfully predicted all the major
chromospheric lines observed in an active M dwarf atmosphere.
The success of chromospheric modeling may be
limited by these workers' assumptions of monotonically rising
atmospheres or even two-component models. (Hawley et al 1996) found
that active M dwarfs in the field show systematic spectral differences
relative to nonactive stars, which may lead to a better understanding
of the atmospheric heating mechanisms: Early-type dMe appear (a) brighter by 0.5 mag from V-K, (b) 0.1-mag redder in (V-I) and (V-K), and (c)
to show systematic differences in the relative strengths of some
near-infrared TiO subbands compared with those of nonactive dM stars of
the same spectral type. These effects may in part be expected if dMe
stars are systematically younger with larger radii. However, they are
only observed in the most massive M dwarfs. On the other hand, while
direct effects of magnetic fields on the structure of the atmosphere
(e.g. through the magnetic pressure term) and the Zeeman splitting of
atomic lines are negligible for all but the outermost photosphere, the
impinging radiation upon the upper photosphere by a magnetically or
acoustically heated chromosphere can be important. The radiation
temperature of the chromosphere is generally much higher than that of
the photosphere, and even a relatively small irradiation of the
photosphere (0.1% of the total flux of the star) by the chromosphere
can introduce important NLTE effects that may change the temperature
structure of the outermost layers. Since the outermost thin radiative
skin of an M dwarf regulates the entire structure of the convective
photosphere and interior (see Section 7 above), these effects may
couple back to the dynamo-generated magnetic and acoustic heating.
Systematic development of (magneto) hydrodynamical studies of
chromospheric activity must be tied to realistic photosphere models to
truly understand cool dwarfs.
THE Teff SCALE OF M DWARFS
|
Section:
|
|
|
|
While bolometric luminosities can be
derived from a careful integration of the observed stellar radiation
for single stars within the accuracy of their known parallaxes (Tinney et al 1993, 1995), and stellar masses can be derived for close binaries down to the hydrogen-burning limit (Henry & McCarthy 1993),
the calibration of the observed magnitudes and spectral types as a
function of the physical atmospheric parameters of the stars still
remains difficult. The determinations of M dwarf effective temperatures
have been refined considerably since the work of Veeder (1974), Peterson (1980), Reid & Gilmore (1984),
who fit blackbody curves through broadband colors and points of assumed
observed continuum. But even the current empirical methods (Berriman et al 1992, Jones et al 1994)
still assume that nearly pure thermal radiation escapes from dM
atmospheres at some wavelength(s). Such an assumption is secure only
for optically thick layers of a nonconvective atmosphere, but models
strongly suggest that M dwarf atmospheres are convective out to optical
depths as low as 10
3 (see Section 7 above). The hazards of any type of Planck flux fitting to an M dwarf spectrum are apparent from Figure 3 and have been emphasized by Allard & Hauschildt (1995a), who showed how strong molecular absorption and flux redistribution obliterate all evidence of the original continuum shape.
Fortunately, two double-line
spectroscopic and eclipsing M dwarf binary systems can offer some
guidance in the subsolar mass regime: CM Draconis and YY Geminorum. Lacy (1977) and later Habets & Heintze (1981) determined the Teff of M dwarfs in these systems based on the observed masses and radii. Figure 4 compares the latest OS models of Brett (1995b), F Allard & PH Hauschildt (in preparation), and Kurucz 1992b to these fundamental stellar calibrators. The Teff scales derived from the spectral synthesis of individual stars of Kirkpatrick et al (1993b) [using the SM models of Allard (1990)] and Leggett et al (1996) [using OS synthetic spectra drawn from the Allard & Hauschildt (1995b) model structures] are also shown, which illustrate the tendency of theoretical Teff to become cooler with developments in the treatment of opacities in the models.
View larger version (50K)
|
Figure 4 The M dwarf Teff scale. Current model-dependent effective temperature scales for cool stars down to the hydrogen-burning limit. Open triangles feature results from spectral synthesis of selected stars from the works of Kirkpatrick et al (1993b), Leggett et al 1996 as indicated. The new generation of OS models by Brett (1995b) and F Allard & PH Hauschildt (in preparation), as interpolated onto theoretical isochrones by Chabrier et al (1996a), reproduce closely the independently determined positions of M dwarfs in the eclipsing binaries CM Dra and YY Gem (Habets & Heintze 1981).
Much uncertainty remains, however, in the lowermost portion of the main
sequence where effects of grain formation can become important.
|
|
As an inspection of Figure 2 reveals, M dwarf photospheric structures are much more sensitive to TiO opacities than to the current uncertainties in the H2O
opacity profiles discussed in Section 4. The result of the use of an OS
treatment of the main molecular opacities, in particular for TiO,
appears therefore to be a breakthrough in the agreement of modeled Teff scales with observations of early-type M dwarfs. Note that the OS models of Kurucz (1992b), on the other hand, suffer from an inaccurate TiO absorption profile and a complete lack of H2O opacities, and the models are therefore clearly inadequate in the regime of VLM stars (i.e. below Teff 4500 K and M 0.8 M ), where molecular opacities dominate the stellar spectra and atmospheric structures.
However, the situation still remains uncertain at Teff 3500 K, where the SM models of Allard & Hauschildt (1995b) seem to yield a better agreement, both with the current Teff estimate for CM Draconis and with the VRI colors of the Leggett (1992)
disk stars sample. This could be understood in terms of an
incompleteness of current optical opacities (see Section 4.1 for
details). The SM technique used for the treatment of TiO opacities in
the models of Allard & Hauschildt (1995b)
compensates for missing optical opacities and may represent a better
intermediate solution in this regime. Kinematics indicate that CM
Draconis is most likely an old disk system and may be slightly
metal-depleted compared with young-disk main-sequence stars (Rucinski 1978, Saumon et al 1995a, Chabrier & Baraffe 1995). Recently, new determinations of the masses and radii of CM Draconis have been obtained by Metcalfe et al (1996),
and a revision of the effective temperatures and metallicity of the
system is under way (Viti et al 1996), which may help improve the
temperature estimate for CM Draconis and shed some light on this region
of the lower main sequence.
The models of F Allard & PH Hauschildt (in preparation) and Brett (1995a, 1995b)
presently neglect the effects of both condensation and grain opacities
that may affect stars cooler than 2700 K. The preliminary dusty models
of Tsuji et al (1996a) (also shown in Figure 4) indicate that condensation effects may drive the Teff
scale of these M dwarfs to cooler values at a given color. Tsuji et
al's models also indicate that when grain opacities begin to alter the
thermal equilibrium of the atmosphere, the lowest-mass stars become
gradually hotter and redder, mimicking more closely the behavior of
blackbodies.
Much uncertainty remains, therefore, at the
lowermost portion of the main sequence. The effects of grain formation
and more complete opacities of TiO promise a better understanding of
the stars and brown dwarfs in the vicinity of the hydrogen-burning
limit [the location of which is roughly indicated in Figure 4
by the termination point of the Allard & Hauschildt model
sequence], but they still remain to be ascertained. These questions are
currently being addressed by Tsuji et al (1996b) and F Allard & PH Hauschildt (in preparation).
THE DETECTABILITY OF BROWN DWARFS
|
Section:
|
|
|
|
Brown dwarfs are substellar objects
not massive enough to sustain stable nuclear fusion and therefore
cannot successfully stabilize on the hydrogen-burning main sequence (Burrows et al 1993, Saumon et al 1995a).
While this boundary between stars and brown dwarfs is fairly well
defined, the recently reported massive extrasolar planets, some of
which have orbital eccentricities more reminiscent of stellar binary
systems (e.g. e = 0.4 for 70 Vir; see Marcy & Butler 1996b), blur the distinction between planets and brown dwarfs. The minimum mass for deuterium burning ( 13 M
J
) may represent a more physically meaningful criterion to distinguish Jovian planets from classical brown dwarfs (Saumon et al 1994).
To develop search strategies for substellar objects, we propose three categories for objects in the brown dwarf regime: (a) transitional objects: massive brown dwarfs (M 0.06 M ) younger than about 109 years that are burning deuterium or even hydrogen temporarily; (b) lithium brown dwarfs: low-mass brown dwarfs ( M 0.06 M ) younger than about 107 years that are still burning deuterium but have not depleted their initial reservoir of lithium and beryllium; and (c) evolved brown dwarfs: brown dwarfs older than 107
109
years that have exhausted their nuclear fuel and have cooled and faded
below the parameters of the coolest main-sequence stars [i.e. Teff = 2000 K and L= 10
4 L (Burrows et al 1993, Baraffe et al 1997)].
Many brown dwarfs of all types are expected to be in the solar
neighborhood and nearby clusters, so it was puzzling that concerted
searches did not reveal any (Martín et al 1994, Marcy et al 1994, Henry 1996).
Transitional Objects
If it is massive enough, a brown dwarf can
initiate thermonuclear fusion early in its evolution, before fading to
invisibility as a component of the dark matter (Burrows et al 1993). Such young objects can approach atmospheric conditions similar to those of M dwarf stars (Teff 3000 K) and can easily be confused with them. An extrapolation of the observed mass spectral
type relation of red dwarfs in short-period binary systems seems to
support the existence of a large population of unrecognized brown
dwarfs among red dwarfs with spectral types later than M7 (Henry & McCarthy 1993, Henry et al 1994, Kirkpatrick et al 1994, 1995, Kirkpatrick & McCarthy 1994). However, further extensions of the Henry & McCarthy (1993) mass-spectral type relation and new evolution models (Baraffe & Chabrier 1996) reveal a sharp drop in the mass-luminosity and mass-color relations from about 0.1 M
to the hydrogen-burning limit, which rules out this naive
extrapolation. The possibility that such objects may masquerade as
known field M dwarfs must still be confirmed.
Transitional objects in the field with unknown age
and mass should betray their youth with the following signatures that
distinguish them from more massive M dwarfs: (a) a lower surface gravity (log g < 5.0), since the brown dwarf is still early on its evolutionary track, and (b)
a higher rotation rate accompanied by more chromospheric and coronal
activity. Unfortunately, both these criteria are difficult to apply in
practice: Metal-rich model spectra for the latest-type dwarfs show very
little sensitivity to changes in gravity (Allard & Hauschildt 1995a).
Moreover, the wings of resonance lines that are sensitive to the
pressure stratification of the photosphere are also sensitive to
slightly increased Teff and/or metallicity that can compensate for a lower gravity (Schweitzer et al 1996).
As for rotation, the relation between age, rotation, and activity in
late-type dwarfs is very poorly understood (see Section 8 above) and is
not a reliable discriminant of transitional objects either. This
situation is best illustrated by the striking contrast between the
candidates PC 0025+0447 (classified dM9.5; Graham et al 1992), which shows one of the strongest H
emission lines of the lower main sequence, and the brown dwarf candidate BRI 0021 012 (classified <dM9.5; Tinney et al 1993), which rotates 20 times faster than other field nonemission M stars (Basri & Marcy 1995).
Faced with these ambiguities, nobody has yet been able to confirm any
transitional object in the field. Young stellar clusters, however, have
yielded more positive results (see below).
Lithium Brown Dwarfs
The key to recognizing such substellar objects
is in their spectra. Since an M dwarf is completely convective, the
entire star is mixed and the surface abundances reflect the elemental
abundances in the core. If the central temperatures are low enough or
the star is young enough, then nuclei should survive in the atmosphere
that would otherwise be completely destroyed in hotter, more massive
stars. This is the case for 7Li nuclei, which are destroyed by proton captures at relatively low temperatures of a few ( 2) million degrees in the interior. Therefore, if you can detect and measure the strength of 7Li
lines in a stellar spectrum and you are armed with accurate atmospheric
and evolution models, you should be able to estimate the mass and age
of the star (Rebolo et al 1992, Maggazzú et al 1993).
To date, the detection of Li I lines is the most
decisive spectral indicator of substellarity for young brown dwarfs
with masses below about 0.06 M .
This test is best used for nearby young clusters where the age is
reasonably well known and the 6707-Ĺ region can be studied at
relatively high dispersion to obtain a precise abundance of lithium. If
the cluster is old enough that only the very lowest mass stars should
retain any surface lithium, as is the case for the Pleiades, then this
is a very powerful test of substellarity. The fact that early searches
for Li I in Pleiades brown dwarf candidates yielded no detections
prompted (Pavlenko et al 1995) and later (Allard & Hauschildt 1995a, b)
to investigate a number of concurrent physical processes that could
prevent the Li I 6708 Ĺ doublet from being seen in those objects. They
explored possible molecular bonding involving lithium and departures
from LTE in the Li I lines, all with negative results: The Li I lines
should be observable in young brown dwarfs. Indeed, recent
spectroscopic observations of Teide 1, Calar 3, and PPl 15 (Rebolo et al 1995, Zapatero Osorio et al 1997, Stauffer et al 1994) show that they are cluster members, have low luminosities, and have retained lithium at their surface (Basri et al 1996, Rebolo et al 1996). Teide 1 and Calar 3 are most likely brown dwarfs with estimated masses of 55 ± 15 M
J
Zapatero Osorio et al 1997. One more candidate, the first found in the field, has also been confirmed to have Li in its atmosphere Thackrah et al 1997. Several more Pleiades candidates may soon be confirmed (or rejected) by the Li test Martín et al 1996.
Accurate estimation of brown dwarf masses depends
critically upon an understanding of the depletion of lithium at the
surface as a function of mass Nelson et al 1993.
This in turn requires evolutionary calculations for low-mass stars that
include the most modern equations of state and non-gray atmospheres for
their outer boundary conditions, as have been performed by (Baraffe et al (1995, 1997), Chabrier & Baraffe (1995), Chabrier et al (1996a), Baraffe & Chabrier (1996), Allard et al (1996).
The evolutionary calculations are largely sensitive to the treatment of
the equations of state, which requires the inclusion of the
thermodynamic properties of a strong-coupled Coulomb plasma, electron
degeneracy, and pressure ionization (Dorman et al 1989, Nelson et al 1993, Burrows et al 1993).
Non-gray atmospheres were found to cause the interiors to become
systematically cooler (and the hydrogen-burning minimum mass smaller)
than when using gray models such as those of Burrows et al (1993)
for a given mass. This is because energy is transported in the
interiors of these objects by convection, and their structure can be
approximated by that of a polytrope of index, n = 1.5 [see Burrows & Liebert (1993)
for a discussion of the polytropic nature of low mass stars and brown
dwarfs]. The polytropic interior characteristics are only specified
once the atmospheric structure is determined, and improvements in the
atmospheric properties, therefore, directly impact the determinations
of the ages, masses, and location in the Hertzprung-Russell (HR)
diagram of these stars. A proper treatment of the atmosphere is
therefore essential to the predicted mass Li abundance relation as a function of the age of brown dwarfs (Chabrier et al 1996a).
EVOLVED BROWN DWARFS
A major breakthrough in finding the missing
link between Jovian planets and low mass stars was the discovery of the
first evolved field brown dwarf Gliese (Gl) 229B by Oppenheimer et al (1995), Nakajima et al (1995). Unfortunately, Gl 229B and other candidates like the astrometrically discovered HD 114762 (Mazeh et al 1996, Williams et al 1997), the ZZ Ceti companion GD 165B (Becklin & Zuckerman 1988, Zuckerman & Becklin 1992), and the massive extrasolar giant planet candidate around 70 Vir (Marcy & Butler 1996b) are
all binary companions that are too faint and too close to their stellar
primaries to be tested for an optical lithium signature. Fortunately,
their low temperatures offer an advantage, since unique changes in
molecular chemistry that occur across the temperature transition from
the coolest M dwarfs ( 2000 K) to the Jovian planets ( 150 K) result in distinctive spectral signatures.
There currently exist three sets of model atmospheres and synthetic spectra for the Teff regime of cool brown dwarfs: (a) The Tsuji et al (1996a, b) models cover the range from 4000 K (although only from 2700 K with grains) to 1000 K; (b) the Allard et al (1996) models reach down (from 10,000 K) to 800 K; and (c) the Marley et al (1996) models cover the range from 1000 K to the temperature of Jupiter, i.e. 150 K. Tsuji et al (1995)
were the first to introduce detailed models and synthetic spectra for
cool brown dwarfs. While unsuitable for high-resolution spectral
synthesis, their models (based on a band model treatment of the
molecular opacities) predict the growing intensity of infrared CH4 bands with Teff
cooler than about 1800 K. This signature was later identified in
the near-infrared spectral distribution of the brown dwarf Gl 229B (Matthews et al 1995, Geballe et al 1996)
and helped confirm the substellar nature of the brown dwarf. The study
of Gl 229B also led to the important realization that, although grain
formation must occur in such cool photospheres and may indeed affect
the spectroscopic properties of late M dwarfs (Tsuji et al 1996a), none of the predicted greenhouse effects of grain opacities seem to be present in Gl 229B Allard et al (1996), Tsuji et al (1996b).
These authors showed that grainless models provide a better description
of the spectral distribution of the brown dwarf when a homogeneous
atmosphere is assumed (see Section 5 above for details; Tsuji et al 1996b). The grainless brown dwarf models of Tsuji et al (1995),
however, are based on band-model opacities that tend to overestimate
the strength of molecular features and cannot be used reliably to
derive accurate atmospheric parameters and absolute fluxes for evolved
brown dwarfs.
The brown dwarf model atmospheres of Allard et al (1996), on the other hand, are based on a direct OS treatment of an ab initio line list for H2O (Miller et al 1994) and lists of transitions observed in planetary atmospheres [e.g. the RADEN, GEISA, HITRAN, and ATMOS projects by Farmer & Norton (1989), Husson et al (1992), Rothman et al (1992), Farmer & Norton (1989), Kuznetsova et al (1993),
respectively] and that reproduce more accurately the spectroscopic
properties of evolved brown dwarfs. This can be appreciated in Figure 1, where the most recent observations of the brown dwarf Gl 229B by Geballe et al (1996) are compared to the best fitting model of Allard et al (1996). Their models, combined with the latest brown dwarf evolution models, led to a Teff of 900 1000 K and a mass of about 30 50 M
J
for Gl 229B. The lack of condensation in their calculations,
however, prevented them from completely covering the range of relevant
brown dwarfs parameters; i.e. brown dwarfs cooler than about 800 K, in
which water vapor begins to condense out of the gas phase, leaving
profoundly transformed photospheres and synthetic spectra. This
limitation was avoided by Marley et al (1996)
by simply neglecting elements that are expected to be condensed at the
effective temperatures of Jovian planets and by including only the H2, He, CH4, NH3, H2O, and H2S
species in their chemical equilibrium calculations. This allowed them
to compute models that cover the regime of brown dwarfs and extrasolar
planets cooler than 1000 K. Despite their approximations in the
treatment of molecular opacities (K-coefficient technique) and
convection (adiabatic mixing only), their model successfully reproduced
most of the observed spectral characteristics of the brown dwarf Gl
229B.
The models of Allard et al (1996), Marley et al (1996) are compared in Figure 5
to summarize the predicted absolute fluxes that brown dwarfs would have
at a distance of 50 parsec (pc). Both the model spectra and blackbody
distributions of the same Teff are shown, which indicate the
range of possible spectroscopic characteristics of brown dwarfs (from a
grainless to a fully dusty atmosphere). As can be seen in that figure,
brown dwarfs are most readily detected at 4.5 m: the peak of their spectral energy distribution. At 5 m,
the hotter (younger or more massive) brown dwarfs and stars show strong
CO bands that cause their flux to drop by nearly 0.5 dex relative to
that at 4.5 m. Between 4.5 and 10 m, opacities of CH4 (and H2O in the hot brown dwarfs) cause the flux to drop by 0.5 1.0 dex. Searches in the 4.5 5 m region and redwards of 10 m
should therefore offer the best possibilities for finding and resolving
brown dwarfs in binaries. The detection limits of current and planned
ground-based and space-based telescopes (from Saumon et al 1994) are also indicated in Figure 5,
which show that brown dwarfs (dusty or not) within 50 pc would be
easily detected by Space Infrared Telescope Facility in the 4.5 5.0 m region. The drop in sensitivity of the various instruments redward of 10 m implies, however, that only brown dwarfs as hot as or hotter than Gl 299B have a good chance of detection at that distance.
View larger version (58K)
|
Figure 5 Predicted absolute fluxes of brown dwarfs at
50 pc as compared with the sensitivity of ground- and space-based
platforms that will be or are currently applied to the search for brown
dwarfs and extrasolar planets. The latter are values reported for the 5
detection of a point source in 1 h of integration, except for the three
NICMOS cameras where the integration is limited to 40 min (see Saumon et al 1994 for details). Models of both Allard et al (1996) (solid line) and Marley et al (1996) (dotted line) are shown, which simulate (a) a brown dwarf near the hydrogen-burning limit (top-most spectrum: Teff = 2000 K), (b) an evolved brown dwarf similar to Gl 229B (central spectra: Teff = 900 K and 960 K), and (c) a brown dwarf closer to the deuterium-burning limit (lower-most spectrum: Teff = 500 K). The corresponding blackbody (dashed line) are also shown for comparison.
|
|
The general spectral distributions
of brown dwarfs hotter than about 800 K are relatively well reproduced
by current models. In contrast, the incompleteness of lists of observed
molecular transitions for several important molecular absorbers may
introduce uncertainty in the predicted absolute fluxes and colors of
cooler brown dwarfs where water is no longer the dominant infrared
opacity. Opacities for CH4, for example, are clearly incomplete because only the strongest lines of CH4
are available from the Gestion et Etude des Informations
Spectroscopiques Atmospheriques and High Resolution Transmision data
bases, while none are available blueward of 1.6 m where systems of CH4 are known to cause strong absorption features in the spectra of planets [e.g. in Jupiter and Titan; see Mickelson & Larson 1992, Bernath et al 1995, Baines et al 1993, Strong et al 1993, Larson & Mickelson 1997]. The complexity of the CH4 compound has so far prevented the accurate modeling of the methane spectrum beyond 2000 cm
1 (see e.g. Tyuterev et al 1994), where a significant fraction of the flux emerges from a brown dwarf or Jovian planet.
THE METAL-DEFICIENT VLM STARS
|
Section:
|
|
|
|
Accurate knowledge of the
compositions of metal-poor VLM subdwarf M (hereafter sdM) stars and
their positions in the Hertzprung-Russell diagram is essential to a
full understanding of the chemical history of our Galaxy. The
reconstruction of the initial mass function in the old disk, halo, and
globular clusters founded on an accurate mass-luminosity relation that is a sensitive function of the stellar chemical composition (Chabrier et al 1996a, D'Antona & Mazzitelli 1996, von Hippel et al 1996) also depends upon it.
Unfortunately the metallicity, atmospheric
parameters, and mass-luminosity relation of sdM stars have long
remained uncertain owing to the lack of VLM stellar model atmospheres
and the resulting lack of bolometric corrections and synthetic
photometry to transform theoretical evolution models into various
empirical planes such as color-magnitude diagrams (see e.g. Greenstein 1989 for details).
This situation began to change in the early 1990s with the Allard Allard (1990)
grid of VLM models that explored a wide range of parameter space with
metallicities ranging from solar values to as low as 1/10,000th solar
(i.e. a logarithmic ratio of metal to hydrogen abundances of [M/H]= 4.0).
This encompassed all relevant age/metallicity populations of cool
dwarfs, including disk, halo, and even unobserved Population III VLM
subdwarfs. These models were later updated by Allard & Hauschildt (1995b),
for an OS treatment of more complete atomic opacities, and most
recently by F Allard & PH Hauschildt (in preparation), for an OS
treatment of the molecular opacities as well. Figure 3
presents three typical model-flux distributions obtained by F Allard
& PH Hauschildt (in preparation) for 3000-K dwarfs: one for
metal-rich conditions of young disk dwarfs, the other two for their
higher-gravity halo and Population III subdwarf counterparts. The
collision-induced absorption (CIA) of H2-H2, H2-H, and H2-He (Borysow et al 1989, Borysow & Frommhold 1989, 1990, Zheng & Borysow 1994) defines the near-infrared continuum in these low-metallicity subdwarfs. Centered at 2 m, H2
CIA depresses the infrared continuum such that most of the flux emerges
only in bluer passbands. Molecular absorption bands of hydrides such as
MgH and CaH are the predominant features in the optical region when
double metals such as TiO and VO are depleted (Mould & Wyckoff 1978, Boeshaar 1976, Bessell 1982).
The optical spectrum of an sdM is therefore much more transparent than
that of a metal-rich M dwarf, where the continuum is defined by H
opacity. The result is a spectral distribution that becomes bluer with decreasing stellar metallicity.
Figure 6 illustrates the behavior of metal-poor VLM stars in the M
V
(V-I) color-magnitude diagram. As can be seen, the work of Monet et al (1992), Dahn et al (1995)
have revealed a number of old disk and halo subdwarfs, some of which
are nearly two magnitudes bluer than previously known field sdM stars. D'Antona (1995) suggested that the extreme subdwarfs of the Monet et al (1992)
sample belong to the galactic halo. But despite the availability of VLM
model atmospheres and synthetic spectra, it has still proven difficult
to untangle the effects of reduced metallicity from those of increased
gravity or reduced effective temperature, which all affect the pressure
structure of the photosphere in similar ways (Allard 1990, Kui 1991, Allard & Hauschildt 1995b).
This difficulty, combined with the lack of accurate line lists for
molecular hydride absorption bands, has hampered efforts to obtain the
atmospheric parameters of observed M subdwarfs (sdM) via spectral
synthesis. Thus far, only a few stars have been analyzed in any detail
(cf Dahn et al 1995).
The M subdwarfs in binary and
multiple systems, for which the metallicity of the hotter primary is
known, and clusters with VLM stars of the same age and metallicity
offer more promise to test Population II evolutionary and atmosphere
models. Two early-type M subdwarfs in binary systems have been recently
reported by Martín et al (1995): the faint, wide proper-motion companion to G116-009 and the fourth and faintest (M
V
= 12.2) companion in the [M/H] = 1.7±0.4 multiple system G176-46 (see also Laird et al 1988, Latham et al 1992). A spectral synthesis of these M subdwarfs with the most recent OS model atmospheres should soon permit calibration of the Teff scale of similar halo subdwarfs (F Allard & PH Hauschildt, in preparation).
One of the unique contributions of the HST to the
understanding of the lower main-sequence star formation in the early
Universe and the contribution of low-mass stars to the universal dark
matter has been the recent detection of large numbers of VLM stars and
white dwarfs in the nearest globular clusters (Paresce et al 1995, Richer et al 1995, Cool et al 1996, Renzini et al 1996). Figure 6 depicts the globular cluster NGC6397 as observed with the Wide Field/Planetary Camera 2 by Cool et al (1996). This cluster displays a genuine Fe underabundance of [M/H] = 1.9 (i.e. 1/80 times solar) and must have formed some 15 10
3
years ago. Its main sequence is narrow and well defined, without the
age, activity, and metallicity dispersion characteristic of samples of
field stars. Fits to that main sequence below about 0.5 M
do not suffer from the same high sensitivity to assumed values of the
helium abundance fraction and mixing length as do fits to more massive
parts of the main sequence (cf Baraffe et al 1997).
Two characteristic kinks also typical of the mass-luminosity relation
shape the lower portion of the main sequence. The first kink at 0.4 M is caused by H2 dissociation (Copenland et al 1970). The second, below 0.1 M , is close to the HST detection limit for several clusters and was first revealed in the Population II models of D'Antona (1987, 1990, 1995).
The first kink is due to the onset of electron degeneracy in the
interior, which is also the cause for the sharp drop in the
mass-luminosity relation below this threshold (Chabrier & Baraffe 1997).
Its exact location depends upon the pressure structure and hence the
opacities of the atmospheres (see also Section 7). The shape of the
lower main sequence of globular clusters represents a nearly
parameter-free test bed for stellar evolution and atmospheric physics
in a metallicity regime where models are less prone to the
incompleteness of H2O and TiO opacities.
D'Antona & Mazzitelli (1985), D'Antona (1987)
laid the ground work for the theoretical modeling of VLM stars down to
the hydrogen-burning limit, and their models long provided among the
best descriptions of the lower main sequence of VLM stars. More
recently, D'Antona & Mazzitelli (1994, 1996) have computed a series of pre main-sequence tracks for Population II stars with masses ranging from 0.015 2.5 M using both standard mixing-length theory and the Canuto & Mazzitelli (1991, 1992)
theory of turbulent convection. However, a critical element to the
handling of convection that describes the stellar structures in these
fully convective objects comes (a) from the equations of state, i.e. the adiabatic gradient (Saumon et al 1995b), and (b)
from the treatment of the surface boundary, i.e. the atmosphere as we
pointed out in Sections 7 and 10. Even with updated versions of the Magni & Mazzitelli (1979) equations of state, the models of D'Antona & Mazzitelli D'Antona & Mazzitelli (1996) are still systematically too hot by 200 K across the VLM range and could not reproduce the observed lower main sequences of globular clusters.
Recently, Alexander et al (1997), Baraffe et al (1997) computed evolution models based on the equation of state of Saumon et al (1995b)
and found a remarkable agreement of their models with the observed main
sequence of the cluster NGC6397 all the way down to the detection
limit, corresponding to masses of 0.13 M . This success is illustrated in Figure 6, where the models of Baraffe et al (1997) are plotted over the photometry. (Note that the dispersion in the diagram at magnitudes fainter than 14.5 mag is likely due to foreground stars.) The two groups, however, disagree on the metallicity of NGC6397: Alexander et al (1997) use bolometric corrections and synthetic colors of Allard & Hauschildt (1995b) and find [M/H] = 2.0. Baraffe et al (1997)
use the more recent, less blanketed (F Allard & PH Hauschildt, in
preparation) models both as surface boundary conditions and for color
transformations, and they find agreement for [M/H] = 1.5, a value that is more consistent with a history of oxygen and other enrichment in old stellar populations (Ryan et al 1991 and references therein). On the other hand, the halo subdwarfs of Monet et al (1992) ( full circles in Figure 6) show a wider dispersion in metallicity. Baraffe et al (1995), Alexander et al (1997) derived isochrones for the most extreme subdwarfs of the Monet et al (1992) sample and obtained a metallicity of [M/H] = 1.5, while the analysis of Baraffe et al (1997) (shown in Figure 6) led to consistant values of [M/H] = 1.3 to 1.5
for the same subdwarfs. Clearly, these results illustrate the progress
brought about by more accurate stellar equations of states, model
atmospheres, and synthetic photometry to the understanding of the lower
main sequence of halo and globular cluster stars.
These successes increase our confidence that the
present Population II stellar models are now sufficiently accurate to
derive reliable mass-luminosity relations and the stellar mass
functions that rely upon them. The mass function in the stellar halo
has been derived from the Dahn et al (1995) luminosity function by Méra et al (1996b, c), Chabrier et al (1996b), using the theoretical mass-luminosity relations drawn from the Baraffe et al (1995, 1997)
evolution models. These authors found a halo mass function which is
rising all the way down to the hydrogen-burning limit, suggesting a
large population of substellar objects in the halo. But to determine
exactly the amount of dark mass in the form of metal-poor substellar
brown dwarfs, we need to extrapolate the mass function into the
substellar domain. Indeed, despite all the advances in cool star models
and detection techniques, we still have identified only a handful of
halo M subdwarfs and none below the hydrogen-burning limit. A hint of
the nature of the missing mass in the halo is nevertheless already
provided by the frequency of events reported by the EROS and MACHO
microlensing surveys: The average time of recorded events indicates the
existence of a rich population of halo objects with masses of at least
0.3 0.5 M (Aubourg 1995, Alcock et al 1996). Metal-poor brown dwarfs seem therefore eliminated as a strong contender for the missing mass (measured to be 1012 M ,
i.e. 10 times the visible mass) in the halo. Rather, the average masses
of the microlenses suggest at least two possibilities, either main
sequence VLM subdwarf stars or white dwarf stellar remnants. However,
although the results of astrometric surveys like 2MASS, DENIS, and more
accurate parallaxes for nearby stars from, for example, the USNO, CCD,
and HIPPARCOS surveys may give us larger halo samples in the near
future, HST pencil surveys and the Dahn et al (1995) luminosity function of the halo seem for now to eliminate the possibility of a large population of VLM stars in the halo (Bahcall et al 1994, Elson et al 1996, Flynn et al 1996, Gould et al 1996, Graff & Freese 1996).
Realistic model atmospheres of VLM
stars and brown dwarfs are essential if we are ever to fully understand
the population of the lower main sequence, the mass-luminosity
relation, and the initial mass function and its dependance on the
chemical history of the Galaxy. The last few years have brought
significant improvements in the models, as well as the first convincing
detections of substellar objects that can be used to test the models.
This progress on both theoretical and observational fronts has led to
several noteworthy advances in our knowledge of the lower main
sequence, including the following:
The progress outlined above and the
good agreement among brown dwarf model spectra generated by various
independent model codes is reassuring, but we cannot afford to be
complacent at this stage. While the effective temperature scale of low
mass stars is now reasonably well determined for Teff
3500 K, it is still poorly defined for M dwarfs and young brown dwarfs
with spectral types later than M6, and it will remain so until more
complete opacities become available. These opacities must include
better treatments of TiO, H2O, and CH4 molecules
(including hot bands), grain size distributions, and grain growth time
scales appropriate to the high pressures and oxygen-rich conditions
found in cool dwarf atmospheres.
While we labor to understand the
outer layers of M dwarfs, we must also remember that the interior
models that define such sequences suffer their own uncertainties, and
that interior and atmospheric properties are intimately coupled in
these objects. With fully convective interiors, their radiation fields
play the role of an energy valve that regulates both the internal
structure and the hydrodynamical (either magnetic or acoustic) heating
of their chromospheres. Systematic development of classical atmosphere
models (in which molecular opacity calculations and laboratory
molecular data are tested) must be tied to (magneto) hydrodynamical
studies of chromospheric activity and applied to interior models to
truly understand cool dwarfs. The fact that researchers are taking
these first steps is an encouraging sign that our field is finally
"coming of age" after a long but fruitful adolescence.
We thank Drs. Hugh RA Jones and Tom R
Geballe for providing UKIRT near-infrared spectra of M dwarfs and Gl
299B and Sandy K Leggett, Mike S Bessell, John M Brett, T Tsuji, MS
Marley, Didier Saumon, DC Monet, Conard C Dahn, and Adrienne M Cool for
providing data in electronic form. We would also like to express our
gratitude to Suzanne Hawley and Peter Ulmschneider for instructive
discussions on activity in M dwarf stars and to Francesca D'Antona,
Isabelle Baraffe, and Gilles Chabrier for discussions about VLM star
evolution and equations of state, as well as for providing their
theoretical isochrones in a numerical form. We are also particularly
indebted to Jaymie Matthews for generously proofreading the manuscript
and to the Cornell Theory Center (CTC) and the San Diego Supercomputer
Center (SDSC) for their allocation of computer time, which made
possible some of the calculations and conclusions presented in this
review.
This work is funded by grants from
the National Science Foundation (NSF) (AST-9217946) to Indiana
University, NASA LTSA (NAG5-3435) to Wichita State University, NASA
LTSA and ATP to the University of Georgia in Athens, and NASA LTSA
(NAGW2628), ATP (NAG53068), and NSF (AST94-17057) to the Arizona State
University.
Literature Cited
|
Section:
|
|
|
|
1.
|
Alcock C,
Allsman RA,
Axelrod TS,
Bennett DP,
Cook KH, et al. 1996. Ap. J. 461:84 103
[CrossRef]
[ISI]
[ADS]
|
2.
|
Alexander DR.
1975. Ap. J. Suppl. 29:363
[CrossRef]
[ADS]
|
3.
|
Alexander DR,
Augason GC,
Johnson HR.
1989. Ap. J. 345:1014 21
[CrossRef]
[ISI]
[ADS]
|
4.
|
Alexander DR,
Brocato E,
Cassisi S,
Castellani V,
Ciacio F,
DeglíInnocenti S.
1997. Astron. Astrophys. 317:90
[ADS]
|
5.
|
Alexander DR,
Ferguson JW.
1994a. In Molecules in the Stellar Environment, ed. UG Jřrgensen, Lect. Notes Phys., pp. 149 62. Berlin/Heidelberg: Springer-Verlag
|
6.
|
Alexander DR,
Ferguson JW.
1994b. Ap. J. 437:879 91
[CrossRef]
[ISI]
[ADS]
|
7.
|
Alexander DR,
Rypma RL,
Johnson HR.
1983. Ap. J. 272:773 80
[CrossRef]
[ISI]
[ADS]
|
8.
|
Allard F.
1990. Model atmospheres for M dwarfs. PhD thesis. Ruprecht Karls University, Heidelberg, Germany
|
9.
|
Allard F.
1994. In Molecular Opacities in the Stellar Environment, ed. P Thejll, U Jřrgensen, IAU Colloq. 146, pp. 1 5. Copenhagen: Niels Bohr Inst.; Nordita
|
10.
|
Allard F,
Hauschildt PH.
1995a. See Tinney 1995, pp. 32 44
|
11.
|
Allard F,
Hauschildt PH.
1995b. Ap. J. 445:433 50
[CrossRef]
[ISI]
[ADS]
|
12.
|
Allard F,
Hauschildt PH,
Baraffe I,
Chabrier G.
1996. Ap. J. Lett. 465:L123 27
[CrossRef]
[ADS]
|
13.
|
Allard F,
Hauschildt PH,
Miller S,
Tennyson J.
1994. Ap. J. Lett. 426:L39 L41
[CrossRef]
[ADS]
|
14.
|
Aubourg E.
1995. Astron. Astrophys. 301:1 5
[ADS]
|
15.
|
Auman JJ.
1967. Ap. J. Suppl. 14:171
[CrossRef]
[ADS]
|
16.
|
Bahcall JN,
Flynn C,
Gould A,
Kirkhakos S.
1994. Ap. J. 435:L51 L54
[CrossRef]
[ISI]
[ADS]
|
17.
|
Baines KH,
West RA,
Giver LP,
Moreno F.
1993. J. Geophys. Res. 98:5517 29
[ISI]
[ADS]
|
18.
|
Balfour WJ,
Klynning L.
1994. Ap. J. 424:1049 53
[CrossRef]
[ISI]
[ADS]
|
19.
|
Baraffe I,
Chabrier G.
1996. Ap. J. 461:L51 L54
[CrossRef]
[ISI]
[ADS]
|
20.
|
Baraffe I,
Chabrier G,
Allard F,
Hauschildt PH.
1995. Ap. J. Lett. 446:L35 L38
[CrossRef]
[ADS]
|
21.
|
Baraffe I,
Chabrier G,
Allard F,
Hauschildt PH.
1997. Ap. J. In press
|
22.
|
Basri G,
Marcy GW.
1994. Ap. J. 431:844 49
[CrossRef]
[ISI]
[ADS]
|
23.
|
Basri G,
Marcy GW.
1995. Astron. J. 109:762 73
[CrossRef]
[ISI]
[ADS]
|
24.
|
Basri G,
Marcy GW,
Graham JR.
1996. Ap. J. 458:600 9
[CrossRef]
[ISI]
[ADS]
|
25.
|
Bauschlicher CWJ,
Langhoff SR.
1986. J. Chem. Phys. 85:5936
[CrossRef]
[ISI]
[ADS]
|
26.
|
Beck HKB,
Gail H-P,
Henkel R,
Sedlmayr E.
1992. Astron. Astrophys. 265:626 42
[ADS]
|
27.
|
Beck HKB,
Hauschildt PH,
Gail H-P,
Sedlmayr E.
1995. Astron. Astrophys. 294:195 205
[ADS]
|
28.
|
Becklin EE,
Zuckerman B.
1988. Nature 336:656 58
[ISI]
[ADS]
|
29.
|
Bergeron P,
Fontaine G,
Brassard P,
Lamontagne R,
Wesemael F, et al. 1993. Ap. J. 106:1987 99
[CrossRef]
[ADS]
|
30.
|
Bergeron P,
McGraw JT.
1990. Ap. J. Lett. 352:L45 L48
[CrossRef]
[ADS]
|
31.
|
Bernath PF,
Black JH,
Brault JW.
1995. Preprint.
|
32.
|
Bernath PF,
Brazier CR.
1985. Ap. J. 288:373 76
[CrossRef]
[ISI]
[ADS]
|
33.
|
Berriman G,
Reid IN,
Leggett SK.
1992. Ap. J. Lett. 392:L31 L33
[CrossRef]
[ADS]
|
34.
|
Bessell MS.
1982. Publ. Astron. Soc. Austr. 4:417 19
[ADS]
|
35.
|
Bessell MS.
1991. Astron. J. 101:662
[CrossRef]
[ISI]
[ADS]
|
36.
|
Bessell MS.
1995. See Tinney 1995, pp. 123 31
|
37.
|
Bessell MS,
Stringfellow GS.
1993. Astron. Astrophys. Rev. 31:433 71
[Abstract]
|
38.
|
Boeshaar PC.
1976. The spectral classification of M dwarf stars. PhD thesis. Ohio State Univ., Columbus, OH.
|
39.
|
Břhm-Vitense E.
1958. Zh. Astrophys. 46:108
[ADS]
|
40.
|
Bohren CF,
Huffman DR.
1983. Absorption and Scattering of Light by Small Particles. New York: Wiley & Sons. 541 pp.
|
41.
|
Borysow A,
Frommhold L.
1989. Ap. J. 341:549 55
[CrossRef]
[ISI]
[ADS]
|
42.
|
Borysow A,
Frommhold L.
1990. Ap. J. Lett. 348:L41 L43
[CrossRef]
[ADS]
|
43.
|
Borysow A,
Frommhold L,
Moraldi M.
1989. Ap. J. 336:495 503
[CrossRef]
[ISI]
[ADS]
|
44.
|
Brett JM.
1989. MNRAS 241:247 57
[ISI]
[ADS]
|
45.
|
Brett JM.
1990. Astron. Astrophys. 231:440 52
[ADS]
|
46.
|
Brett JM.
1995a. Astron. Astrophys. 295:736 54
[ADS]
|
47.
|
Brett JM.
1995b. Astron. Astrophys. Suppl. 109:263 64
[ADS]
|
48.
|
Brett JM,
Plez B.
1993. Publ. Astron. Soc. Austr. 10:250 53
[ADS]
|
49.
|
Bryja C,
Humphreys RM,
Jones TJ.
1994. Astron. J. 107:246 53
[CrossRef]
[ISI]
[ADS]
|
50.
|
Buchholz B.
1995. Acoustically heated chromospheres of late type stars. PhD thesis. Univ. Heidelberg. 135 pp.
|
51.
|
Burrows A,
Hubbard WB,
Lunine JI.
1989. Ap. J. 345:939 58
[CrossRef]
[ISI]
[ADS]
|
52.
|
Burrows A,
Hubbard WB,
Saumon W,
Lunine JI.
1993. Ap. J. 406:158 71
[CrossRef]
[ISI]
[ADS]
|
53.
|
Burrows A,
Liebert J.
1993. Mod. Phys. Rev. 65:301
[CrossRef]
[ISI]
[ADS]
|
54.
|
Butler RP,
Marcy GW.
1996. Ap. J. Lett. 464:L153 56
[CrossRef]
[ADS]
|
55.
|
Butler RP,
Marcy GW,
Williams E,
Hauser H,
Shirts P.
1997. Ap. J. Lett. 474:115 18
[CrossRef]
[ADS]
|
56.
|
Cameron AGW,
Pine MR.
1973. Icarus 18:377
[ISI]
[ADS]
|
57.
|
Canuto VM,
Mazzitelli I.
1991. Ap. J. 370:295 311
[CrossRef]
[ISI]
[ADS]
|
58.
|
Canuto VM,
Mazzitelli I.
1992. Ap. J. 389:724 30
[CrossRef]
[ISI]
[ADS]
|
59.
|
Chabrier G,
Baraffe I.
1995. Ap. J. Lett. 451:L29 L32
[ADS]
|
60.
|
Chabrier G,
Baraffe I.
1997. Astron. Astrophys. In press
|
61.
|
Chabrier G,
Baraffe I,
Plez B.
1996a. Ap. J. Lett. 459:L91 L94
[ADS]
|
62.
|
Chabrier G,
Segretain L,
Méra D.
1996b. Ap. J. 468:L21 L24
[CrossRef]
[ISI]
[ADS]
|
63.
|
Chan KL,
Nordlun A,
Steffen M,
Stein RF.
1991. Solar Interior and Atmosphere, pp. 223 74. Tucson, AZ: Univ. Ariz. Press
|
64.
|
Cherchneff I,
Barker JR.
1992. Ap. J. 394:703 16
[CrossRef]
[ISI]
[ADS]
|
65.
|
Collins JG,
Fa˙ TDJ.
1974. J. Quant. Spectrosc. Radiat. Transfer 14:1259
[CrossRef]
[ISI]
|
66.
|
Comeron F,
Rieke GH,
Burrows A,
Reike MJ.
1993. Ap. J. 416:185 203
[CrossRef]
[ISI]
[ADS]
|
67.
|
Cool AM,
Piotto G,
King IR.
1996. Ap. J. 468:655 62
[CrossRef]
[ISI]
[ADS]
|
68.
|
Copenland H,
Jensen JO,
Jřrgensen HE.
1970. Astron. Astrophys. 5:12
[ADS]
|
69.
|
Cram LE,
Mullan DJ.
1985. Ap. J. 294:626 33
[CrossRef]
[ISI]
[ADS]
|
70.
|
Dahn CC,
Liebert J,
Harris HC,
Guetter HH.
1995. See Tinney 1995, pp. 239 48
|
71.
|
D'Antona F.
1987. Ap. J. 320:653 62
[CrossRef]
|
72.
|
D'Antona F.
1990. In Physical Processes in Fragmentation of Star Formation, ed. RC Dolcetta, C Choisi, A Di Fazio, pp. 367 79. Dordrecht: Kluwer
|
73.
|
D'Antona F.
1995. See Tinney 1995, pp. 13 23
|
74.
|
D'Antona F,
Mazzitelli I.
1985. Ap. J. 296:502 13
[CrossRef]
|
75.
|
D'Antona F,
Mazzitelli I.
1994. Ap. J. Suppl. 90:467 500
[CrossRef]
|
76.
|
D'Antona F,
Mazzitelli I.
1996. Ap. J. 456:329 36
[CrossRef]
|
77.
|
Davidge TJ,
Boeshaar PC.
1993. Ap. J. Lett. 403:L47 L50
[CrossRef]
[ADS]
|
78.
|
Davis SP.
1994. In Molecules in the Stellar Environment, ed. UG Jřrgensen, Lect. Notes Phys., pp. 397 411. Berlin/Heidelberg: Springer-Verlag
|
79.
|
Davis SP,
Littleton JE,
Phillips JG.
1986. Ap. J. 309:449 54
[CrossRef]
[ISI]
[ADS]
|
80.
|
De Jager C,
Neven L.
1957. Ličge R. Soc. Mem. 18:257
|
81.
|
Deutsch RW.
1994. PASP106:1134 37
|
82.
|
Dobrijevic M,
Goutoulli L,
Toublanc D,
Parisot JP,
Brillet J.
1992. In Titan, ESA Symp. Ser., pp. 329 32, II Univ. Bordeaux, France
|
83.
|
Dominik C,
Sedlmayr E,
Gail H-P.
1989. Astron. Astrophys. 223:227 36
[ADS]
|
84.
|
Dorman B,
Nelson LA,
Chau WJ.
1989. Ap. J. 342:1003 818
[CrossRef]
[ISI]
[ADS]
|
85.
|
Doverstal M,
Weijnitz P.
1992. Mol. Phys. 75:1357
[ISI]
|
86.
|
Elson RAW,
Santiago BX,
Gilmore GF.
1996. New Astron. 1:1
[CrossRef]
[ISI]
[ADS]
|
87.
|
Farmer CB,
Norton RH.
1989. A High-Resolution Atlas of the Infrared Spectrum of the Sun and the Earth Atmosphere from Space, Vol. 1: The Sun, NASA Ref. Publ. 1224. Washington, DC: NASA
|
88.
|
Fegley BJ,
Lodders K.
1996. Ap. J. Lett. 472:L37 L39
[ADS]
|
89.
|
Fleming TA.
1988. Ap. J. 331:958 73
[CrossRef]
[ISI]
[ADS]
|
90.
|
Fleming TA,
Schmitt JHMM,
Giampapa MS.
1995. Ap. J. 450:401 10
[CrossRef]
[ISI]
[ADS]
|
91.
|
Flynn C,
Gould A,
Bahcall J.
1996. Ap. J. Lett. 466:L55 L58
[CrossRef]
[ADS]
|
92.
|
Gail H-P,
Sedlmayr E.
1988. Astron. Astrophys. 206:153 68
[ADS]
|
93.
|
Geballe TR,
Kulkarni SR,
Woodward CE,
Sloan GC.
1996. Ap. J. Lett. 467:L101 4
[CrossRef]
[ADS]
|
94.
|
Giampapa MS,
Liebert J.
1986. Ap. J. 305:784 94
[CrossRef]
[ISI]
[ADS]
|
95.
|
Giampapa MS,
Worden SP,
Linsky JL.
1982. Ap. J. 258:740 60
[CrossRef]
[ISI]
[ADS]
|
96.
|
Gliese W,
Jahreiss H.
1991. Astron. Rechen-Inst. Heidelberg Mitt. Ser. A 224:161 64
|
97.
|
Goody RM,
Yung YL.
1989. Atmospheric Radiation. London/New York: Oxford Univ. Press. 2nd ed.
|
98.
|
Gould A,
Bahcall JN,
Flynn C.
1996. Ap. J. 465:759 68
[CrossRef]
[ISI]
[ADS]
|
99.
|
Graff DS,
Freese K.
1996. Ap. J. Lett. 467:L65 L68
[CrossRef]
[ADS]
|
100.
|
Graham JR,
Matthews K,
Greenstein JL,
Neugebauer G,
Tinney CG,
Persson SE.
1992. Astron. J. 104:2016 21
[CrossRef]
[ISI]
[ADS]
|
101.
|
Gray DF.
1984. Ap. J. 277:640 47
[CrossRef]
[ISI]
[ADS]
|
102.
|
Green PJ,
Margon B.
1994. Ap. J. 423:723 32
[CrossRef]
[ISI]
[ADS]
|
103.
|
Green PJ,
Margon B,
MacConnell DJ.
1991. Ap. J. Lett. 380:L31 L34
[CrossRef]
[ADS]
|
104.
|
Greenstein JL.
1989. Pac. Astron. Soc. Publ. 101:787 810
[CrossRef]
[ISI]
[ADS]
|
105.
|
Grossman SA.
1996. MNRAS 279:305 36
[ISI]
[ADS]
|
106.
|
Guillot T,
Gautier D,
Chabrier G,
Mosser B.
1994. Icarus 112:337 53
[CrossRef]
[ISI]
[ADS]
|
107.
|
Gurvich LV.
1978 1981. Thermodinamicheskie Svoistva Individuak'nikh Vershev, Vol. 1 3. Moscow: Sovit Acad. Sci.
|
108.
|
Gustafsson B,
Jřrgensen UG.
1994. Astron. Astrophys. Rev. 6:19 65
[ISI]
[ADS]
|
109.
|
Habets GMHJ,
Heintze JRW.
1981. Astron. Astrophys. Suppl. 46:193 237
[ADS]
|
110.
|
Hambly NC,
Hawkins MRS,
Jameson RF.
1993. Astron. Astrophys. Suppl. 100:607 40
[ADS]
|
111.
|
Han C,
Gould A.
1996. Ap. J. 467:540 45
[CrossRef]
[ISI]
[ADS]
|
112.
|
Hauschildt PH,
Wehrse R,
Starrfield S,
Shaviv G.
1992. Ap. J. 393:307 28
[CrossRef]
[ISI]
[ADS]
|
113.
|
Hawley SL,
Fisher GH.
1994. Ap. J. 426:387 403
[CrossRef]
[ISI]
[ADS]
|
114.
|
Hawley SL,
Reid IN,
Gizis JE.
1997. Astron. J. In press
|
115.
|
Heber U,
Bade N,
Jordan S,
Voges W.
1993. Astron. Astrophys. 267:L31 L34
[ADS]
|
116.
|
Hedgecock LM,
Naulin C,
Costes M.
1995. Astron. Astrophys. 304:667 77
[ADS]
|
117.
|
Henry TJ.
1996. Sky Telescope 91:24
[ADS]
|
118.
|
Henry TJ,
Kirkpatrick JD,
Ianna PA,
Jahreiss H.
1995. Bull. Am. Astron. Soc. 187:7022
[ADS]
|
119.
|
Henry TJ,
Kirkpatrick JD,
Simons DA.
1994. Astron. J. 108:1437 44
[CrossRef]
[ISI]
[ADS]
|
120.
|
Henry TJ,
McCarthy DW.
1993. Astron. J. 106:773 89
[CrossRef]
[ISI]
[ADS]
|
121.
|
Houdebine ER,
Panagi PM.
1990. Astron. Astrophys. 231:459 65
[ADS]
|
122.
|
Husson N,
Bonnet B,
Scott NA,
Chedin A.
1992. J. Quant. Spectrosc. Radiat. Transfer 48:509
[CrossRef]
[ISI]
[ADS]
|
123.
|
Irwin AW.
1987. Astron. Astrophys. 182:348 58
[ADS]
|
124.
|
Irwin AW.
1988. Astron. Astrophys. Suppl. 74:145 60
[ADS]
|
125.
|
Isbell D,
Morse D.
1996. NASA press release 96-10
|
126.
|
Jensen P.
1989. J. Mol. Spectrosc. 133:438
[ISI]
|
127.
|
Johns-Krull C,
Valenti J.
1996. Ap. J. Lett. 459:L95 L98
[ADS]
|
128.
|
Jones HRA,
Longmore AJ,
Allard F,
Hauschildt PH.
1996. MNRAS 280:77 94
[ISI]
[ADS]
|
129.
|
Jones HRA,
Longmore AJ,
Allard F,
Hauschildt PH,
Miller S,
Tennyson J.
1995. MNRAS 277:767 76
[ISI]
[ADS]
|
130.
|
Jones HRA,
Longmore AJ,
Jameson RF,
Mountain CM.
1994. MNRAS 267:413 23
[ISI]
[ADS]
|
131.
|
Jřrgensen UG.
1992. Rev. Mex. Astron. Astrofis. 23:49 62
[ISI]
[ADS]
|
132.
|
Jřrgensen UG.
1994. Astron. Astrophys. 284:179 86
[ADS]
|
133.
|
Jřrgensen UG,
Jensen P,
Sorensen GO.
1994. In Molecular Opacities in the Stellar Environment, ed. P Thejll, U Jřrgensen, IAU Colloq. 146, pp. 51 54. Copenhagen: Bohr Inst., Nordita
|
134.
|
Joy AH,
Abt HA.
1974. Ap. J. Suppl. 28:1 18
[CrossRef]
[ADS]
|
135.
|
Keane TC,
Yuan F,
Ferris JP.
1996. Icarus 122:205
[CrossRef]
[ISI]
[ADS]
|
136.
|
Keenan PC,
Schroeder LW.
1952. Ap. J. 115:82
[CrossRef]
[ISI]
[ADS]
|
137.
|
Kim YC,
Fox PA,
Demarque P,
Sofia S.
1996. Ap. J. 461:499 506
[CrossRef]
[ISI]
[ADS]
|
138.
|
Kippenhan R.
1962. Proc. XXVIII Course, Tech. Rep., Int. School Enrico Fermi
|
139.
|
Kipper T,
Jřrgensen UG,
Klochkova VG,
Panchuk VE.
1996. Astron. Astrophys. 306:489
[ADS]
|
140.
|
Kirkpatrick JD,
Beichman CA.
1995. Bull. Am. Astron. Soc. 187:7512
[ADS]
|
141.
|
Kirkpatrick JD,
Henry TJ,
Liebert J.
1993a. Ap. J. 406:701 7
[CrossRef]
[ISI]
[ADS]
|
142.
|
Kirkpatrick JD,
Henry TJ,
Simons DA.
1995. Astron. J. 109:797 807
[CrossRef]
[ISI]
[ADS]
|
143.
|
Kirkpatrick JD,
Kelly DM,
Rieke GH,
Liebert J,
Allard F,
Wehrse R.
1993b. Ap. J. 402:643 54
[CrossRef]
[ISI]
[ADS]
|
144.
|
Kirkpatrick JD,
McCarthy DWJ.
1994. Astron. J. 107:333 49
[CrossRef]
[ISI]
[ADS]
|
145.
|
Kirkpatrick JD,
McGraw JT,
Hess TR,
Liebert J,
McCarthy DW J.
1994. Ap. J. Suppl. 94:749 88
[CrossRef]
[ADS]
|
146.
|
Kui R.
1991. Model atmospheres for M-dwarfs. PhD. thesis. Natl. Univ. Aust.
|
147.
|
Kunde V,
Hanel R,
Maguire W,
Gautier D,
Baluteau JP, et al. 1982. Ap. J. 263:443 67
[CrossRef]
[ISI]
[ADS]
|
148.
|
Kurucz RL.
1970. Smithsonian Astrophys. Obs. Spec. Rep. 309. Washington, DC: Smithsonian Astrophys. Obs.
|
149.
|
Kurucz RL.
1992a. Rev. Mex. Astron. Astrofis. 23:45
[ISI]
[ADS]
|
150.
|
Kurucz RL.
1992b. Rev. Mex. Astron. Astrofis. 23:187
[ISI]
[ADS]
|
151.
|
Kurucz RL.
1993. Molecular data for opacity calculations, Kurucz CD-ROM No. 15
|
152.
|
Kurucz RL.
1994. Atomic data for opacity calculations, Kurucz CD-ROM No. 1
|
153.
|
Kusch HJ.
1958. Z. Astrophys. 45:1
[ADS]
|
154.
|
Kuznetsova LA,
Pazyk EA,
Stolyarov AV.
1993. Russ. J. Phys. Chem. 67:2046
|
155.
|
Lacy CH.
1977. Ap. J. 218:444
[CrossRef]
[ISI]
[ADS]
|
156.
|
Laird JB,
Carney BW,
Latham DW.
1988. Astron. J. 95:1843
[CrossRef]
[ISI]
[ADS]
|
157.
|
Langhoff SR,
Bauschlicher CW.
1990. J. Mol. Spectrosc. 141:243
[ISI]
|
158.
|
Langhoff SR,
Bauschlicher CWJr .
1991. Ap. J. 375:843
[CrossRef]
[ISI]
[ADS]
|
159.
|
Langhoff SR,
Bauschlicher CWJr .
1994. In Molecules in the Stellar Environment, ed. UG Jřrgensen, Lect. Notes Phys., pp. 310 25. Berlin/Heidelberg: Springer-Verlag
|
160.
|
Larson LE,
Mickelson ME.
1997. Preprint
|
161.
|
Latham DW,
Mazeh T,
Stefanik RP,
Davis RJ,
Carney BW, et al. 1992. Astron. J. 104:774 95
[CrossRef]
[ISI]
[ADS]
|
162.
|
Lazarev VV,
Pnomarev YN.
1992. Optics Lett. 17:1283
[ISI]
[ADS]
|
163.
|
Leggett SK.
1992. Ap. J. Suppl. 82:351 94
[CrossRef]
[ADS]
|
164.
|
Leggett SK,
Allard F,
Berriman G,
Dahn CC,
Hauschildt PH.
1996. Ap. J. Suppl. 104:117
[CrossRef]
[ADS]
|
165.
|
Leggett SK,
Harris HC,
Dahn CC.
1994. Astron. J. 108:944
[CrossRef]
[ISI]
[ADS]
|
166.
|
Leggett SK,
Hawkins MRS.
1989. MNRAS 238:145
[ISI]
[ADS]
|
167.
|
Leinert C,
Weitzel N,
Richichi A,
Eckart A,
Taccono-Garman LE.
1994. Astron. Astrophys. 291:L47
[ADS]
|
168.
|
Liebert J,
Kirkpatrick JD,
Beichman C,
Reid IN,
Monet DC,
Dahn CC.
1995. Bull. Am. Astron. Soc. 187:7502
[ADS]
|
169.
|
Liebert J,
Probst RG.
1987. Astron. Astrophys. Rev. 25:473
[Abstract]
|
170.
|
Liebert J,
Schmidt G,
Lesser M,
Stephanian JA,
Lipovetsky VA,
Chaffe FH,
Foltz CB,
Bergeron P.
1994. Ap. J. 421:733
[CrossRef]
[ISI]
[ADS]
|
171.
|
Linsky JL,
Wood BE,
Brown A,
Giampapa MS,
Ambruster C.
1995. Ap. J. 455:670
[CrossRef]
[ISI]
[ADS]
|
172.
|
Ludwig CB.
1971. Appl. Opt. 10(5):1057
|
173.
|
Maggazzú A,
Martín EL,
Rebolo R.
1993. Ap. J. 404:L17
[CrossRef]
[ADS]
|
174.
|
Magni G,
Mazzitelli I.
1979. Astron. Astrophys. 72:134
[ADS]
|
175.
|
Marcy G,
Basri G,
Graham JR.
1994. Ap. J. Lett. 428:L57
[CrossRef]
[ADS]
|
176.
|
Marcy GW,
Butler RP.
1996a. Ap. J. Lett. 464:L147
[CrossRef]
[ADS]
|
177.
|
Marcy GW,
Butler RP.
1996b. Ap. J. Lett. 464:L151
[ADS]
|
178.
|
Marcy GW,
Butler RP,
Williams E,
Bildsten L,
Graham J.
1997. Ap. J. In press
|
179.
|
Marcy GW,
Chen GH.
1992. Ap. J. 390:550
[CrossRef]
[ISI]
[ADS]
|
180.
|
Marilli E,
Catalano S,
Trigilio C.
1986. Astron. Astrophys. 167:297
[ADS]
|
181.
|
Marley MS,
Saumon D,
Guillot T,
Freedman R,
Hubbard WB, et al. 1996. Science 272:1919
[Medline]
[ISI]
[ADS]
|
182.
|
Martín EL,
Rebolo R,
Magazzú A.
1994. Ap. J. 436:262
[CrossRef]
[ISI]
[ADS]
|
183.
|
Martín EL,
Rebolo R,
Zapatero Osorio MR.
1995. See Tinney 1995, pp. 253 56
|
184.
|
Martín EL,
Rebolo R,
Zapatero Osorio MR.
1996. Ap. J. 469:706
[CrossRef]
[ISI]
[ADS]
|
185.
|
Mathioudakis M,
Doyle JG.
1992. Astron. Astrophys. 262:523
[ADS]
|
186.
|
Mathys G,
Solanki SK.
1989. Astron. Astrophys. 208:189
[ADS]
|
187.
|
Matthews K,
Nakajima SR,
Oppenheiner BR.
1995. IAU Circ. 6280:1
[ADS]
|
188.
|
Mauas PJD,
Falchi A.
1996. Astron. Astrophys. 310:245
[ADS]
|
189.
|
Mayor M,
Queloz D.
1995. Nature 378:355
[ISI]
[ADS]
|
190.
|
Mazeh T,
Latham DW,
Stefanik RP.
1996. Ap. J. 466:415
[CrossRef]
[ISI]
[ADS]
|
191.
|
Méra D,
Chabrier G,
Baraffe I.
1996a. Ap. J. 459:L87
[ISI]
[ADS]
|
192.
|
Méra D,
Chabrier G,
Schaeffer R.
1996b. Europhys. Lett. 33:327
[ISI]
|
193.
|
Méra D,
Chabrier G,
Schaeffer R.
1996c. Astron. Astrophys. In press
|
194.
|
Merer AJ,
Huang G,
Cheung AS-C,
Taylor AW.
1987. J. Mol. Spectrosc. 125:465
[ISI]
|
195.
|
Metcalfe TS,
Mathieu RD,
Latham DW,
Torres G.
1996. Ap. J. 456:356
[CrossRef]
[ISI]
[ADS]
|
196.
|
Michelis CHP.
1976. A study of van der Waals broadening for lines of astrophysical interest. PhD thesis. Texas Univ., Austin
|
197.
|
Mickelson ME,
Larson LE.
1992. Bull. Am. Astron. Soc. 24:990
[ADS]
|
198.
|
Mihalas D.
1978. Stellar Atmospheres. San Francisco: Freeman. 2nd ed.
|
199.
|
Miller S,
Tennyson J,
Jones HRA,
Longmore AJ.
1994. In Molecules in the Stellar Environment, ed. UG Jřrgensen, Lect. Notes Phys., pp. 296 309. Berlin/Heidelberg: Springer-Verlag
|
200.
|
Monet DG,
Dahn CC,
Vrba FJ,
Harris HC,
Pier JR, et al. 1992. Astron. J. 103:638
[CrossRef]
[ISI]
[ADS]
|
201.
|
Morris S,
Wyller A.
1967. Ap. J. 150:877
[CrossRef]
[ISI]
[ADS]
|
202.
|
Mould JR.
1975. Astron. Astrophys. 38:283
[ADS]
|
203.
|
Mould JR.
1976. Ap. J. 210:402
[CrossRef]
[ISI]
[ADS]
|
204.
|
Mould JR,
Wyckoff S.
1978. MNRAS 182:63
[ISI]
[ADS]
|
205.
|
Muchmore D.
1987. Ap. J. 278:769
[CrossRef]
[ADS]
|
206.
|
Mullan DJ,
Cheng QQ.
1993. Ap. J. 412:312
[CrossRef]
[ISI]
[ADS]
|
207.
|
Mullan DJ,
Cheng QQ.
1994. Ap. J. 420:392
[CrossRef]
[ISI]
[ADS]
|
208.
|
Mullan DJ,
Dermot J.
1987. In NASA Wash. M-Type Stars, pp. 455 79
|
209.
|
Mullan DJ,
Fleming TA.
1996. Ap. J. 464:890
[CrossRef]
[ISI]
[ADS]
|
210.
|
Nakajima T,
Oppenheimer BR,
Kulkarni SR,
Golimowski DA,
Matthews K,
Durrance ST.
1995. Nature 378:463
[ISI]
[ADS]
|
211.
|
Neale L,
Tennyson J.
1995. Ap. J. Lett. 454:L169
[ADS]
|
212.
|
Nelson LA,
Rappaport S,
Chiang E.
1993. Ap. J. 413:364
[CrossRef]
[ISI]
[ADS]
|
213.
|
Noyes RW,
Hartmann LW,
Balinunas SL,
Duncan DK,
Vaughan AH.
1984. Ap. J. 279:763
[CrossRef]
[ISI]
[ADS]
|
214.
|
Oppenheimer BR,
Kulkarni SR,
Nakajima T,
Matthews K.
1995. Science 270:1478
[Medline]
[ISI]
[ADS]
|
215.
|
Paresce F,
De Marchi G,
Romaniello M.
1995. Ap. J. 468:655
|
216.
|
Partridge H,
Schwenke DW.
1997. J. Chem. Phys. In press
|
217.
|
Pavlenko YV,
Rebolo R,
Martín EL,
Garcia-Lopez RJ.
1995. Astron. Astrophys. 303:807
[ADS]
|
218.
|
Peterson BR.
1980. Astron. Astrophys. 82:53
[ADS]
|
219.
|
Peterson BR.
1989. Sol. Phys. 121:299
[ADS]
|
220.
|
Peytremann E.
1974. Astron. Astrophys. 33:203
[ADS]
|
221.
|
Phillips JG,
Davis SP.
1993. Ap. J. 409:860
[CrossRef]
[ISI]
[ADS]
|
222.
|
Phillips JG,
Davis SP,
Lindgren B,
Balfour WJ.
1987. Ap. J. Suppl. 65:721
[CrossRef]
[ADS]
|
223.
|
Piskunov N,
Valenti J,
Johns-Krull C.
1996. Bull. Am. Astron. Soc. 188:4407
|
224.
|
Plez B,
Brett JM,
Nordlund A.
1992. Astron. Astrophys. 256:551
[ADS]
|
225.
|
Pollack JB,
Hollenbach D,
Beckwith S,
Simonelli DP,
Roush T,
Fong W.
1994. Ap. J. 421:615
[CrossRef]
[ISI]
[ADS]
|
226.
|
Pollack JB,
McKay CP,
Christofferson BM.
1985. Icarus 64:471
[CrossRef]
[ISI]
[ADS]
|
227.
|
Rebolo R,
Martín EL,
Basri G,
Marcy GW,
Zapatero Osorio MR.
1996. Ap. J. 469:L53
[CrossRef]
[ISI]
[ADS]
|
228.
|
Rebolo R,
Martín EL,
Maggazzú A.
1992. Ap. J. 389:L83
[CrossRef]
[ISI]
[ADS]
|
229.
|
Rebolo R,
Zapatero Osorio MR,
Martín EL.
1995. Nature 377:129
[ISI]
[ADS]
|
230.
|
Reid IN.
1994. Astrophys. Space Sci. 217:57
[ISI]
[ADS]
|
231.
|
Reid IN,
Gilmore RF.
1984. MNRAS 206:19
[ISI]
[ADS]
|
232.
|
Reid IN,
Hawley SL,
Gizis JE.
1995a. Astron. J. 110:1838
[CrossRef]
[ISI]
[ADS]
|
233.
|
Reid IN,
Hawley SL,
Mateo M.
1995b. MNRAS 272:828
[ISI]
[ADS]
|
234.
|
Reid N.
1993. MNRAS 265:785
[ISI]
[ADS]
|
235.
|
Renzini A,
Bragaglia A,
Ferraro FR,
Gilmozzi R,
Ortolani S, et al. 1996. Ap. J. 465:23
[CrossRef]
[ISI]
[ADS]
|
236.
|
Richer HB,
Fahlman GG,
lbata RA,
Stetson PB,
Bell RA, et al. 1995. Ap. J. 451:17
[ADS]
|
237.
|
Robinson RD.
1980. Ap. J. 239:961
[CrossRef]
[ISI]
[ADS]
|
238.
|
Rosner R,
Golub L,
Vaiana G.
1985. Astron. Astrophys. Rev. 23:413
[Abstract]
|
239.
|
Rossi SCF,
Marciel WJ,
Benevides-Soares P.
1985. Astron. Astrophys. 148:93
[ADS]
|
240.
|
Rossow WB.
1978. Icarus 36:1
[CrossRef]
[ISI]
[ADS]
|
241.
|
Rothman LS,
Gamache RR,
Tipping RH,
Rinsland CP,
Smith MAH, et al. 1992. J. Quant. Spectrosc. Radiat. Transfer 48:469
[CrossRef]
[ISI]
[ADS]
|
242.
|
Rucinski SM.
1978. Acta Astron. 28:167
[ADS]
|
243.
|
Russell HM.
1934. Ap. J. 79:317
[CrossRef]
[ADS]
|
244.
|
Rutten RGM.
1986. Astron. Astrophys. 159:291
[ADS]
|
245.
|
Rutten RGM,
Zwaan C,
Schrijver CJ,
Duncan DK,
Mewe R.
1989. Astron. Astrophys. 219:239
[ADS]
|
246.
|
Ryan SG,
Norris JE,
Bessell MS.
1991. Astron. J. 102:303
[CrossRef]
[ISI]
[ADS]
|
247.
|
Saar SH.
1988. Ap. J. 324:441
[CrossRef]
[ISI]
[ADS]
|
248.
|
Saar SH.
1994. In Infrared Solar Physics, ed. DM Rabin et al, IAU Colloq. 154, p. 493. Dordrecht: Kluwer
|
249.
|
Saumon D,
Bergeron P,
Lunine JI,
Hubbard WB,
Burrows A.
1994. Ap. J. 424:333
[CrossRef]
[ISI]
[ADS]
|
250.
|
Saumon D,
Burrows A,
Hubbard WB.
1995a. See Tinney 1995, pp. 3 12
|
251.
|
Saumon D,
Chabrier G,
Van Horn HM.
1995b. Ap. J. Suppl. 99:713
[CrossRef]
[ISI]
[ADS]
|
252.
|
Saumon D,
Hubbard WB,
Burrows A,
Guillot T,
Lunine JI,
Chabrier G.
1996. Ap. J. 460:993
[CrossRef]
[ISI]
[ADS]
|
253.
|
Sauval AJ,
Tatum JB.
1984. Ap. J. Suppl. 56:193
[CrossRef]
[ADS]
|
254.
|
Schryber HS,
Miller S,
Tennyson J.
1995. J. Quant. Spectrosc. Radiat. Transfer 53:373
[CrossRef]
[ISI]
|
255.
|
Schweitzer A,
Hauschildt PH,
Allard F,
Basri G.
1996. MNRAS 283:821 29
[ISI]
[ADS]
|
256.
|
Seab GC,
Snow TP.
1989. Ap. J. 347:479
[CrossRef]
[ISI]
[ADS]
|
257.
|
Seaton JM.
1992. Rev. Mex. Astron. Astrofis. 23:180
[ISI]
[ADS]
|
258.
|
Seaton JM,
Zeippen CJ,
Tully JA,
Pradhan AK,
Mendoza C, et al. 1992. Rev. Mex. Astron. Astrofis. 23:19
[ISI]
[ADS]
|
259.
|
Sharp CM,
Huebner WF.
1990. Ap. J. Suppl. 72:417
[CrossRef]
[ADS]
|
260.
|
Simons DA,
Becklin EE.
1992. Ap. J. 390:431
[CrossRef]
[ISI]
[ADS]
|
261.
|
Sneden L,
Johnson HR,
Krupp BM.
1976. Ap. J. 204:281
[CrossRef]
[ISI]
[ADS]
|
262.
|
Stauffer J,
Hamilton D,
Probst R.
1994. Astron. J. 108:155
[CrossRef]
[ISI]
[ADS]
|
263.
|
Stauffer JR,
Giampapa MS,
Herbst W,
Vincent JM,
Hartmann LW,
Stern RA.
1991. Ap. J. 374:142
[CrossRef]
[ISI]
[ADS]
|
264.
|
Stauffer JR,
Hartmann LW.
1986. Ap. J. Suppl. 61:531
[CrossRef]
[ADS]
|
265.
|
Stevenson DJ.
1991. Astron. Astrophys. Rev. 29:163
[Abstract]
|
266.
|
Stone RC,
Monet DG,
Monet AKB,
Walker RL,
Ables HD.
1996. Astron. J. 111:1721
[CrossRef]
[ISI]
[ADS]
|
267.
|
Strong K,
Taylor FW,
Calcutt SB,
Remedios JJ,
Ballard J.
1993. J. Quant. Spectrosc. Radiat. Transfer 50:363
[CrossRef]
[ISI]
[ADS]
|
268.
|
Takeda Y,
Kato K-I,
Watanabe Y,
Sadakane K.
1996. Publ. Astron. Soc. Jpn. 48:511
[ISI]
[ADS]
|
269.
|
Thackrah A,
Jones HRA,
Hawkins M.
1997. MNRAS. In press
|
270.
|
Tinney CG.
1993. The faintest stars. PhD thesis. Calif. Inst. Technol., Pasadena
|
271.
|
Tinney C, ed. 1995. The Bottom of the Main Sequence and Beyond. ESO Astrophys. Symp. Berlin/Heidelberg: Springer-Verlag
|
272.
|
Tinney CG.
1996. MNRAS 281:644
[ISI]
[ADS]
|
273.
|
Tinney CG,
Mould JR,
Reid IN.
1993. Astron. J. 105:1045
[CrossRef]
[ISI]
[ADS]
|
274.
|
Tinney CG,
Reid IN,
Gizis J,
Mould JR.
1995. Astron. J. 110:3014
[CrossRef]
[ISI]
[ADS]
|
275.
|
Tsuji T.
1966. Publ. Astron. Soc. Jpn. 18:127
[ISI]
[ADS]
|
276.
|
Tsuji T.
1973. Astron. Astrophys. 23:411
[ADS]
|
277.
|
Tsuji T.
1994. In Molecules in the Stellar Environment, ed. UG Jřrgensen, Lect. Notes Phys., pp. 79 97. Berlin/Heidelberg: Springer-Verlag
|
278.
|
Tsuji T,
Ohnaka K,
Aoki W.
1995. See Tinney 1995, pp. 45 52
|
279.
|
Tsuji T,
Ohnaka K,
Aoki W.
1996a. Astron. Astrophys. 305:L1
[ADS]
|
280.
|
Tsuji T,
Ohnaka K,
Aoki W,
Nakajima T.
1996b. Astron. Astrophys. 308:L29
[ADS]
|
281.
|
Tyuterev VG,
Babikov YL,
Tashkun SA,
Perevalov VI,
Nikitin A, et al. 1994. J. Quant. Spectrosc. Radiat. Transfer 52:459
[CrossRef]
[ISI]
[ADS]
|
282.
|
Unsöld A.
1955. Physik der Sternatmosphären. Berlin/Göttingen/Heidelberg: Springer-Verlag. 2nd ed.
|
283.
|
Vardya MS.
1966. MNRAS 134:347
[ISI]
[ADS]
|
284.
|
Veeder GJ.
1974. Astron. J. 79:1056
[CrossRef]
[ISI]
[ADS]
|
285.
|
Verner DA,
Verner EM,
Ferland GJ.
1996. Bull. Am. Astron. Soc. 188:5418
[ADS]
|
286.
|
Viti S,
Jones HRA,
Tennyson J,
Polyansky OL,
Miller S, et al. 1995. Bull. Am. Astron. Soc. 187:10314
[ADS]
|
287.
|
von Hippel T,
Gilmore G,
Tanvir N,
Robinson D,
Jones D.
1996. Astron. J. 112:192
[CrossRef]
[ADS]
|
288.
|
Warren SJ,
Irwin MJ,
Evans DW,
Liebert J,
Osmer PS,
Hewett PC.
1993. MNRAS 261:185
[ISI]
[ADS]
|
289.
|
Weast RC, ed. 1988. Handbook of Chemistry and Physics. Boca Raton, FL: CRC Press. 2nd ed.
|
290.
|
Wehrse R.
1972. Astron. Astrophys. 19:453
[ADS]
|
291.
|
Wehrse R,
Liebert J.
1980. Astron. Astrophys. 86:139
[ADS]
|
292.
|
Weidemann V.
1955. Z. Astrophys. 36:101
[ADS]
|
293.
|
Williams DM,
Boyle RP,
Morgan WT,
Rieke GH,
Stauffer JR,
Rieke MJ.
1996. Ap. J. 464:238
[CrossRef]
[ISI]
[ADS]
|
294.
|
Williams E,
Butler R,
Marcy G.
1997. Ap. J. In press
|
295.
|
Wilson OC,
Skumanich A.
1964. Ap. J. 140:1401
[CrossRef]
[ISI]
[ADS]
|
296.
|
Zapatero Osorio MR,
Rebolo R,
Martín EL.
1997. Astron. Astrophys. 317:164
[ADS]
|
297.
|
Zapatero Osorio MR,
Rebolo R,
Martín EL,
Garcia Lopez RJ.
1996. Astron. Astrophys. 305:519
[ADS]
|
298.
|
Zheng C,
Borysow A.
1994. Ap. J. 441:960
[CrossRef]
[ISI]
[ADS]
|
299.
|
Ziurys LM,
Barclay WLJ,
Anderson MA.
1992. Ap. J. 384:63
[CrossRef]
[ADS]
|
300.
|
Ziurys LM,
Fletcher DA,
Anderson MA,
Barclay WLJ.
1996. Ap. J. Suppl. 102:425
[CrossRef]
[ADS]
|
301.
|
Zubko VG.
1987. On the Microstructure Theory of a Condensate Layer in a White Dwarf Atmosphere. Kiev: Acad. Sci. Ukrainian SSR
|
302.
|
Zuckerman B,
Becklin EE.
1992. Ap. J. 386:260
[CrossRef]
[ISI]
[ADS]
|
|
|
|