paper_id
stringlengths 9
16
| version
stringclasses 26
values | yymm
stringclasses 311
values | created
timestamp[s] | title
stringlengths 6
335
| secondary_subfield
sequencelengths 1
8
| abstract
stringlengths 25
3.93k
| primary_subfield
stringclasses 124
values | field
stringclasses 20
values | fulltext
stringlengths 0
2.84M
|
---|---|---|---|---|---|---|---|---|---|
astro-ph/0211326 | 1 | 0211 | 2002-11-14T13:33:24 | The dynamics of the nebula M1-67 around the run-away Wolf-Rayet star WR 124 | [
"astro-ph"
] | A new point of view on the dynamics of the circumstellar nebula M1-67 around the run-away Wolf-Rayet (WR) star WR 124 is presented. We found that it has been interacting with the surrounding ISM and has formed a bow shock due to its high velocity of about 180 km/s relative to the local ISM. The star is about 1.3 parsec away from the front of this bow shock. The outbursts that are responsible for the nebula are assumed to be discrete outbursts that occurred inside this bow shock. The ejecta collide with this bow shock shortly after the outburst. After the collision, they are dragged away by the pressure of the ISM, along the surface of the bow shock. The bow shock is oriented in such way that we are looking from the rear into this paraboloid, almost along the main axis. Evidence for this is given firstly by the fact that the far hemisphere is much brighter than the near hemisphere, secondly by the fact that there is hardly any emission found with radial velocities higher than the star's radial velocity, thirdly by the fact that the star looks to be in the centre of the nebula, as seen from Earth, and finally by the asymmetric overall velocity distribution of the nebula, which indicates higher radial velocities in the centre of the nebula, and lower velocities near the edges. We find evidence for at least two discrete outbursts that occurred inside this bow shock. For these outbursts, we find expansion velocities of about 150 km/s and dynamical timescales of about 8 and 20 kyr, which are typical values for LBV outbursts. We therefore conclude that M1-67 originates from several outbursts that occurred inside the bow shock around WR 124, during an LBV phase that preceded the current WR phase of the star. | astro-ph | astro-ph |
Astronomy&Astrophysicsmanuscript no. h3857
(DOI: will be inserted by hand later)
May 12, 2018
The dynamics of the nebula M1-67 around the run-away
Wolf-Rayet star WR 124
M.V. van der Sluys1 and H.J.G.L.M. Lamers1,2
1 Astronomical
Institute, Princetonplein 5, NL-3584 CC Utrecht,
the Netherlands, ([email protected]) and
([email protected])
2 SRON Laboratory for Space Research, Sorbonnelaan 2, NL-3584 CA Utrecht, the Netherlands
Received July 18, 2002 / Accepted November 7, 2002
Abstract. (The image quality has been reduced to submit this paper to astro-ph. To get a full resolution version, please visit
http://www.astro.uu.nl/sluys/m1-67/ .)
A new point of view on the dynamics of the circumstellar nebula M1-67 around the run-away Wolf-Rayet (WR) star WR 124 is
presented. We simulated the outbursts of nebulae with different morphologies, to compare the results to the observed dynamical
spectra of M1-67. We found that it has been interacting with the surrounding ISM and has formed a bow shock due to its high
velocity of about 180 km s−1 relative to the local ISM. The star is about 1.3 parsec away from the front of this bow shock.
The outbursts that are responsible for the nebula are assumed to be discrete outbursts that occurred inside this bow shock. The
ejecta collide with this bow shock shortly after the outburst. After the collision, they are dragged away by the pressure of the
ISM, along the surface of the bow shock. The bow shock is oriented in such way that we are looking from the rear into this
paraboloid, almost along the main axis. Evidence for this is given firstly by the fact that the far hemisphere is much brighter
than the near hemisphere, secondly by the fact that there is hardly any emission found with radial velocities higher than the
star's radial velocity, thirdly by the fact that the star looks to be in the centre of the nebula, as seen from Earth, and finally by
the asymmetric overall velocity distribution of the nebula, which indicates higher radial velocities in the centre of the nebula,
and lower velocities near the edges. We find evidence for at least two discrete outbursts that occurred inside this bow shock. For
these outbursts, we find expansion velocities of vexp ≈ 150 km s−1 and dynamical timescales of about 0.8 and 2 × 104 yr, which
are typical values for LBV outbursts. We therefore conclude that M1-67 originates from several outbursts that occurred inside
the bow shock around WR 124, during an LBV phase that preceded the current WR phase of the star.
Key words. Stars: circumstellar matter -- Stars: individual: WR 124 -- Stars: mass-loss -- Stars: Wolf-Rayet -- ISM: individual
objects: M1-67 -- ISM: jets and outflows
1. Introduction
In this article, we describe our research on the dynamics of
the Wolf-Rayet ring nebula M1-67. M1-67 is a bright nebula
around the Wolf-Rayet (WR) star WR 124. The star has a high
heliocentric velocity of almost 200 km s−1 and is also known
as Merrill's star (Merrill 1938) and 209 BAC. The star is clas-
sified as a population I WN8 star (Bertola 1964) and is lo-
cated in the constellation Sagittarius. Distance estimates vary
from about 4.5 kpc (Pismis & Recillas-Cruz 1979) to 6.5 kpc
(Nugis & Lamers 2000). The star has a terminal wind velocity
of 710 km s−1 and a mass loss of 2.45 × 10−5 M⊙ yr−1. Its mass
is estimated to be about 20 M⊙ and its luminosity 6 × 105 L⊙
(Nugis & Lamers 2000).
The nebula M1-67 around WR 124 shows a clumpy struc-
ture, and most of the gas is concentrated in knots and filaments
(Sirianni et al. 1998). An HST image of the nebula is displayed
offprint
Send
[email protected]
requests
to: H.J.G.L.M.
Lamers,
e-mail:
in Fig. 3. The nebula was first classified as an H region. After
the discovery that the nebula has about the same radial velocity
as WR 124, it was suggested that the nebula might be a plan-
etary nebula (Minkowski 1946). However, the presence of a
WR star and the N-enhancement and O-deficiency of the neb-
ula suggest a WR ring nebula (Sirianni et al. 1998). The dis-
tance estimates also point in the direction of an ejected nebula,
so that M1-67 is now generally accepted as a Wolf-Rayet ring
nebula.
Though WR ring nebulae are not necessarily ring-shaped,
they often exhibit a structure of arcs or rings. This suggests
that the nebulae may be created by discrete outburst events.
It is generally thought that WR ring nebulae originate from a
Luminous Blue Variable (LBV) stage of the central star, which
is supposed to precede the WR phase.
More than half of the LBVs have circumstellar nebulae
(Nota & Clampin 1997). The different nebulae are very similar
in terms of physical properties. The expansion velocities are in
the order of 50 to 100 km s−1, their sizes about 1 parsec, and the
2
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
dynamical ages are in the order of 104 yr. The densities of the
nebulae are generally found to be low (500 to 1000 cm−3) and
the temperatures are in the range of 5000 to 10 000 K (Lamers
et al. 2001).
It is still a point of debate whether the LBV outbursts occur
during a Red Supergiant (RSG) or a Blue Supergiant (BSG)
phase. The RSG scenario is proposed by Stothers & Chin
(1993, 1996), who suggest that the ejection of mass occurs only
as a single event during a brief RSG phase. They explain the
enhanced abundances of heavy elements of the ejecta by con-
vective mixing in the RSG envelope (Stothers & Chin 1993,
1996).
On the other hand, Langer et al. (1994) suggest that after
the star has left the main sequence, it moves red-ward in the
Hertzsprung-Russel diagram (HR diagram) and the expanding
envelope becomes unstable, so that the star starts to develop
extreme mass loss. This mass loss may be as high as 5 × 10−3
M⊙ yr−1 and is observed as the LBV ejecta. They explain the
chemically enriched ejecta by rotation induced mixing. These
eruptions therefore take place when the star is a BSG and pre-
vent it from becoming a RSG (Langer et al. 1994; Lamers et al.
2001).
The research of the chemical composition of LBV-ejecta by
Lamers et al. (2001) also indicates that the LBV eruptions oc-
cur during a BSG phase. They suggest that the LBV outbursts
are induced by the rapid, near-critical rotation of the star. In
their scenario, the star is also being prevented from becoming
a RSG by the mass loss. However, they point out that if a mas-
sive, optically thick shell is being expelled from the star, it will
cool as it expands and the physical conditions will temporarily
be similar to that in the outer layers of a RSG, so that the for-
mation of dust can also happen in this case. This mechanism
explains the observed Humphreys-Davidson limit, that depicts
the lack of RSGs with luminosities higher than 6 × 105 L⊙.
The goal of our research is to disentangle the geometry and
dynamics of the nebula M1-67. In order to do so, we create
different numerical models, of which the output is compared
to available observations. Firstly, we present the observational
data we use for this study in Sect. 2. We will then discuss our
models for freely expanding outbursts in Sect. 3. The reason
why we let these models expand freely is that the O-star that
precedes a WR star blows a bubble of typically 30 pc in the ISM
during its lifetime (Lamers & Cassinelli 1999). It will take an
outburst of 100 km s−1 more than 105 yr to cross this distance,
so that it can indeed be considered to expand without any dis-
turbance. We will find that no satisfying fit can be made, so
that the assumption of a freely expanding outburst is wrong.
We show that the cause for this is that the star has a high veloc-
ity relative to the ISM. This causes a paraboloid-like bow shock
instead of a more or less spherical bubble. The star is about 1 pc
away from the front of this bow shock, so that an outburst of
100 km s−1 needs only 104 yr to cross it. Once it has done so,
part of the outburst will collide with the bow shock, and will
possibly be dragged away along the bow shock surface. We
discuss the bow shock models for the case of a constant wind
velocity in Sect. 4. In Sect. 5 we compare the theoretical bow
shock models for a constant stellar wind with the observations
and find a remarkable resemblance. Sect. 6 discusses the results
of impacts of outbursts on the bow shock surface. In Sect. 7 we
summarise the results and present the conclusions of this study.
2. Observations
For this research, we used the following three sets of observa-
tional data.
2.1. Long-slitspectra
The first dataset we used, is velocity information from long slit
spectra we obtained from A. Nota, published in Sirianni et al.
(1998). These data consist of 13 long slit spectra, taken with the
ESO Multi Mode Instrument (EMMI) at the 3.5m NTT in La
Silla. Each slit is positioned over the nebula in the east-west di-
rection, at constant declination. The declinations of the slits lie
between -30.82′′and +24.69′′relative to the star's declination.
In total, 413 good data points (right ascension, declination and
radial velocity) were derived from the spectra, which formed
the input for our study. These data points have been plotted in
Fig. 1.
2.2. Fabry-P´erotimages
The second dataset are Fabry-P´erot (FP) images obtained from
Grosdidier et al. (1999). The FP images are shown in Fig. 2.
These 30 images were made in August 1996, using CFHT-SIS,
with the ´etalon of the Universit´e Laval, Qu´ebec, Canada. Each
image was taken in Hα, at a slightly different wavelength, so
that it displays the emission at a certain radial velocity. These
observations give a much more detailed view of the velocity
distribution than the long slit spectra. Each image consists of
100 × 100 pixels. From the darkest points, we used the coordi-
nates (x, y, v) in the same way as the data points from the long
slit spectra. A few thousand points were used. Note that the
southern part of the nebula, more than approximately 20′′south
of the star, is missing (compare to Fig. 3 and see Fig. 8(b))
due to deteriorating seeing during the observations (Grosdidier
et al. 1999).
2.3. HSTimage
For reference, we also used the Hubble Space Telescope im-
age taken by Grosdidier et al. (1998), for example to identify
structures found in the Fabry-P´erot images.
The image is a composite image of four WFPC2 images
with a total exposure time of 10 000 seconds, taken in March
1997, using the narrow band F656N Hα filter. In the same way,
four images of the same field were taken through the broad-
band F675W R filter and combined to obtain a 'continuum'
image close to Hα. The continuum image was flux-scaled and
subtracted from the first composite image to obtain the deep
continuum-subtracted Hα image with the field stars removed,
of which a negative version is shown in Fig. 3.
In this image different distinct arcs are visible. However,
although these arcs are clearly seen locally, it is very difficult
to find a global system of rings. The fact that the arcs are so
clear, indicates that they are probably formed during different,
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
3
Fig. 1.Position-
velocity
di-
a-
grams
ob-
tained
from
the
ob-
ser-
va-
tions
by
Sirianni
et al.
(1998).
Each
panel
dis-
plays
the
re-
sult
of
one
slit
po-
si-
tion,
where
the
ra-
dial
ve-
loc-
ity
is
plot-
ted
against
the
po-
si-
tion
along
the
slit.
The
lower
right
panel
shows
the
cov-
er-
age
of
data
points
in
the
sky,
with
the
star
in
the
ori-
4
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
Fig. 2.Fabry-
P´erot
im-
ages
ob-
tained
from
Grosdidier
et al.
(1999).
Each
im-
age
shows
a
pic-
ture
of
the
emis-
sion
of
M1-
67
at
a
spe-
cific
ra-
dial
ve-
loc-
ity,
that
is
dis-
played
in
the
up-
per
right
cor-
ner.
The
star
is
in
the
cen-
tre
of
each
im-
age.
discrete outbursts, whereas the deficit of a global pattern may
tell us that the history of the nebula is less straightforward than
we might think a first.
3. Models of freely expanding outbursts
We have developed models to simulate ejected nebulae numeri-
cally and calculate spectra from them at different slit positions.
The velocity information of the part of the nebula in the slit
is then converted to a position-velocity plot (PV-plot), which
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
5
Fig. 3. Negative image of M1-67 made by the Hubble Space
Telescope (Grosdidier et al. 1998).
can be compared to observational data. The purpose of these
models is to give insight in the relation between features in PV-
plots and the geometry of the nebulae that cause them. Hence
we have modelled the most common and most likely structures
for ejected nebulae: spheres, ellipsoids and cones.
3.1. Thegeometryofthenebulae
We start with creating a geometry in the computer. Our struc-
tures are hollow and made up of single dots that represent the
surface of the nebula. Solid or partially filled structures can be
made by nesting different shells into each other. For each dot,
that represents a small volume, the three rectangular coordi-
nates with respect to the star are calculated.
For spheres we use:
(1)
x = Rneb cos ψ sin θ
y = Rneb sin ψ sin θ
z = Rneb cos θ
where Rneb is the radius of the nebula, ψ ranges from 0 to 2π and
2 , in such way that the dots are evenly spread
θ from − π
over the spherical surface.
2 to + π
Fig. 4. Explanation of the coordinates and angles used in this
section. The x and y axes represent the celestial coordinates
right ascension and declination, the z axis is the radial coordi-
nate, with negative values toward the observer. The thick arrow
represents the symmetry axis of the structure. Left panel: the
inclination is changed by rotation around the x axis about the
angle i. Right panel: the inclined structure is rotated around the
z axis to change the position angle ϕ.
where z varies between −Rneb and +Rneb and ψ between 0 and
2π.
The ellipsoids and cones can be made either elongated to
simulate jets, or flat to create disks. All structures have axial
symmetry and are created with the line of sight as their sym-
metry axis (z axis). They are then rotated around a line perpen-
dicular to the line of sight (x axis) to change the inclination
i and subsequently rotated around the line of sight (which is
then no longer the symmetry axis) to change the position angle
ϕ. This is illustrated in Fig. 4.
3.2. Thevelocitymodels
After choosing the orientation of a structure, we overlay slits of
finite width and transform the radial coordinates of the underly-
ing points to radial velocities. We assume the velocity-distance
relation for the nebula to be
vr = v0 r
vz = vr(cid:18) z
Rneb!α
r(cid:19) = v0 r
Rneb!α
(cid:18) z
r(cid:19)
(4)
(5)
Ellipsoids are made in the same way. We only consider ax-
ially symmetric ellipsoids. Their geometry is then defined by
the oblateness m ≡ Rpol/Req, with Rpol and Req the polar and
equatorial radius respectively. The x and y coordinates are then
similar to that of the sphere in Eq. (1) and the z coordinate is
given by:
z = Rneb cos θ · m
(2)
Bipolar cones are characterised by a semi-opening angle
(SOA). The nebulae with this geometry can be created using
the equations:
x = z cos ψ tan SOA
y = z sin ψ tan SOA )
(3)
ity, r = (cid:16)x2 + y2 + z2(cid:17)1/2
Here, Rneb is the radius of the nebula, v0 is the expansion veloc-
is the distance to the star and z is the
radial distance to the star, measured along the line of sight. For
α we can choose a suitable radius-velocity dependence: α < 0
for braking after the outburst, α = 0 for a continuous outflow
with constant velocity, α = 1 for a single, short burst with no
further interaction and α > 1 for acceleration of the material af-
ter the outburst. For our models of freely expanding outbursts,
a thin nebular shell with α = 1 is used. Since all our structures
are made up of single dots, the transformations can be carried
out point by point.
For each slit, our models produce one PV-plot, in which the
radial velocity information under the slit is plotted against the
6
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
position along the slit. We took all slits parallel to the x axis, so
that they have a constant y coordinate. The PV-plots thus show
the radial velocity vz plotted against the position x. All plots
are scaled to arbitrary units in position and velocity, because
we are only interested in the shape of a structure. In actual ob-
servations, the structures may therefore appear stretched out in
either direction (position or velocity). The result of this whole
exercise is a 'reference guide' that we used to determine the
origin of certain structures in observed long slit spectra.1
3.3. FreelyexpandingoutburstsforWR124
We tried to fit the different geometries described above and
combinations of them to the long slit spectra data of M1-67,
shown in Fig. 1. The clumpy and chaotic structure of the neb-
ula made it hard to fit. We then also tried to fit parts of the
nebula. However, we never found a convincing solution. We
therefore questioned our assumption that the outburst around
WR 124 is expanding freely and concluded that this may not
be the case. The Fabry-P´erot data of Fig. 2, which we obtained
afterwards, supported this conclusion. The following features
in the velocity data of M1-67 particularly convinced us:
-- There is hardly any nebular emission seen at radial veloc-
ities higher than +208 km s−1, though the star has a ra-
dial velocity of about +199 km s−1. In case of a continu-
ous outflow by a wind or discrete LBV-like outbursts, one
would expect to find nebular emission with radial velocities
of tens, or even hundreds of kilometres per second higher
than that of the star.
-- Emission is seen at radial velocities from about +78
to +208 km s−1, of which the average is +143 km s−1.
However, the bulk of this emission is found at velocities
higher than about +140 km s−1 (See Fig. 2). This means
that the high-velocity part of the nebula shows much more
emission than the low-velocity part. This was already men-
tioned by Solf & Carsenty (1982).
-- There is an asymmetry in the emission distribution. When
looking at the different panels in Fig. 2, starting at the high-
est velocity and skipping down, it is clearly seen that the
emission appears close to the star and becomes increasingly
broader. When skipping from the lowest velocity panel up-
ward, it is seen that (except for some small patches at lower
velocities) the emission 'starts' at about 120 -- 130 km s−1 as
a very broad pattern. The far hemisphere is thus narrower
than the near hemisphere and the bright emission region
between +140 and +208 km s−1 looks like a triangle that is
pointing away from us.
The explanation that can be found in literature is that the
nebula as a whole is braked by the ISM, so that the star is dis-
placed from the centre of expansion toward the leading edge
(Chu & Treffers 1981; Solf & Carsenty 1982). This means that
the far and near hemisphere are braked equally, which is in
contradiction with the asymmetry in the emission distribution.
Instead, we think that the reason that the model of a freely ex-
1 The simulated long-slit spectra for various geometries and veloc-
ities are available from the first author via e-mail.
panding nebula does not hold for M1-67, is that there is indeed
a strong interaction between the stellar wind and the ISM and
that this interaction produces a bow shock around the stellar
surface. As we will point out in Sect. 6, the bow shock model
is also qualitatively able to explain the chaotic structure of the
nebula, which is much more difficult to do with a freely ex-
panding nebula.
4. The bow shock models
The failure of fitting a freely expanding shell to the velocity
data of M1-67, the high radial velocity of WR 124 and the
global picture from the more detailed Fabry-P´erot data show
that our assumption of the freely expanding shell is incorrect.
Instead, the high velocity of the star is responsible for the for-
mation of a paraboloid-shaped bow shock. The possibility of a
bow shock was already mentioned by Grosdidier et al. (1999).
The bow shock model allows the star to be much closer to the
ISM, which makes it easier for outbursts to collide with it. We
will discuss the models for such a bow shock in this section. We
start with a two-dimensional model and convert it to a rotation-
ally symmetric three-dimensional model, which can be tilted
to any wanted orientation, before obtaining the radial velocity
information.
4.1. 2Dmodelsforcontinuousmassloss
The models we use here are published by Canto et al. (1996).
They derive analytical expressions to describe wind-wind in-
teractions in general. They also discuss the special case where
one wind is plane-parallel and the other spherical, as is the
case when a star with spherical wind is ploughing through a
homogeneous ISM. In our models we assume the interstellar
medium to be homogeneous. We also assume that the stellar
wind is constant over a long period (∼ 104 yr) compared to the
timescale of LBV-outbursts (usually on the order of 102 yr),
except during these outbursts. The wind-ISM interaction is re-
sponsible for the geometry and dynamics of the bow shock,
while all emission comes from the material that is ejected dur-
ing the outbursts and its collision into the bow shock surface.
This implies the assumption that the momentum that is carried
along with the outbursts is lower than the momentum in the
wind-ISM interaction.
The distance between the star and the front of the bow
shock is called the stagnation point distance and can simply
be derived by momentum equilibrium:
1/2
r0 =
Mv∞
4πρismv2
ism
(6)
where vism is the velocity of the ISM relative to the star (we will
observe the situation from the rest frame of the star for simplic-
ity). The geometry of the bow shock is expressed in terms of r,
the distance between the star and a point on the bow shock, as
a function of the angle θ that is defined in Fig. 5(a):
r = r0 csc θp3(1 − θ cot θ)
(7)
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
7
and by Eq. (8).
The three-dimensional bow shock model is now complete
and oriented in such a way that the observer is looking into the
hollow shock from the rear. We can now rotate the shock to
any orientation we like, about the angles i and ϕ as indicated
in Fig. 4. The velocity components are rotated the same way,
and we can use the celestial coordinates x and y and the radial
velocity vz to compare the predicted radial velocity maps to the
observations.
Fig. 6 displays the Right Ascension against the radial ve-
locity vz of a number of bow shocks with different orienta-
tions. From these plots, two effects of the orientation of the
bow shock can easily be seen. Firstly, when looking at the dif-
ferent columns from the left to the right, the inclination of the
bow shock increases. The most important result of this is that
the maximum radial velocity observed shifts from 0 km s−1
(which is the radial velocity of the star) for i = 0◦ to more than
30 km s−1 for i = 50◦. This means that the velocity difference
between the maximum radial velocity where nebular emission
is seen and the radial velocity of the star puts an upper limit
to the inclination. As we have seen in Sect. 3.3, in the case of
M1-67 this difference is approximately 10 km s−1, which gives
a strong indication for a small inclination of the main axis of
the bow shock. Secondly, when one looks through the different
rows from the top to the bottom, the position angle changes.
The result is that the axial symmetry around RA = 0 disap-
pears. We will later see that for the case of M1-67, only about
the upper 100 km s−1 in radial velocity are observed, so that
this effect manifests itself as a steep drop in radial velocity at
one side of the maximum and a slower drop at the other side.
The side of the steep drop indicates the direction where the top
of the bow shock points at. For example, at the lower right plot
of Fig. 6, the steep side is at the right, which means that the
top of the bow shock is pointing to this side, so that the star
is moving, with respect to its local ISM, toward higher Right
Ascension.
4.3. Orientationofthebowshock
It is possible to roughly say something about the orientation of
a bow shock of a given star with respect to the line of sight
(LOS). This way, we can limit the total number of possible
orientations of the bow shock in M1-67 appreciably. In order
to do so, we need to know the velocities of both the star and
the ISM surrounding the star. The orientation of the main axis
of the bow shock is simply the orientation of the vector that
is the difference of the spatial velocity of the star and that of
the local ISM. The spatial velocity of the star can be derived
from the measured radial velocity and proper motion. For the
ISM we will assume that it follows the laws of galactic rotation
and derive its velocity relative to the sun. After subtracting the
ISM velocity from that of the star, we can calculate the total
velocity difference and the inclination and position angle of the
main axis of the bow shock. We will derive these quantities
here for the general case of a star with galactic longitude and
latitude l and b respectively and distance d from the sun, and
then apply this to M1-67 in Sect. 5.
Fig. 5. Graphical representation of the bow shock model by
Canto et al. (1996), with the physical parameters for WR 124.
Upper panel (a): geometry (z versus x coordinate), lower panel
(b): velocity along the shock surface (v). In both panels the
horizontal axis represents the symmetry (z) axis of the shock in
parsecs.
(Canto et al. 1996). The velocity of the material along the sur-
face of the shock is split in a component parallel to the symme-
try axis (z axis, vz) and a component perpendicular to it (vxy).
They are given by:
vz =
vxy =
Mv∞ sin2 θ − 4πρismv2
ismr2 sin2 θ
2 M(1 − cos θ) + 4πρismvismr2 sin2 θ
Mv∞ (θ − sin θ cos θ)
2 M(1 − cos θ) + 4πρismvismr2 sin2 θ
v = qv2
z + v2
xy
(8)
(9)
(10)
(Canto et al. 1996), where v is the total velocity along the bow
shock surface. In Fig. 5 the geometry and the total velocity
along the surface are displayed.
4.2. 3Dmodelsforcontinuousmassloss
The model of Canto et al. (1996) is two-dimensional and can
easily be converted into a 3D-model by
x = r sin θ cos ψ
y = r sin θ sin ψ
z = r cos θ
where θ is defined in Fig. 5(a) and ψ takes evenly spread val-
ues between 0 and 2π around the symmetry axis. Thus, the 3D
surface is obtained by rotating Fig. 5(a) around the horizontal
axis. The velocity components are given by
vx = vxy cos ψ
(11)
(12)
vy = vxy sin ψ )
8
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
Fig. 6. Model output for the 3D bow shock models. For all models vism = 200 km s−1 and the star is placed at a Right Ascension
of 0′′ and a radial velocity of 0 km s−1. The different rows have position angles (ϕ) of 0, 45 and 90◦, the columns have inclinations
(i) of 0, 10, 20, 30, 40, and 50◦. For each panel, the horizontal axis shows the Right Ascension (RA) in arcseconds, the vertical
axis displays the radial velocity in km s−1. See the text in Sect. 4.2 for more explanation and interpretation.
In the observations of the radial velocity and proper motion
of the star the motion of the sun is already taken into account,
so that we indeed need to correct for the motion of the sun when
calculating the velocity of the ISM. First, we will correct for the
motion of the local standard of rest (LSR), that moves with a
velocity of v0 = 220 km s−1 around the galactic centre (GC)
and subsequently the peculiar motion of the sun with respect
to the LSR will be taken into account. This peculiar velocity is
calculated from Hipparcos observations for different classes of
stars in the galactic plane by Mignard (2000). We will use the
averages of the values given in that article:
u⊙ = 10.0 ± 1.3 km s−1
v⊙ = 13.9 ± 3.7 km s−1
w⊙ = 7.4 ± 2.6 km s−1
(13)
for the direction toward the GC, the direction of galactic rota-
tion and toward the north galactic pole (NGP) respectively.
Fig. 7. The Milky Way as seen from the NGP, with the defini-
tions of the quantities used in this section.
When looking along the line of sight (LOS) defined by the
galactic coordinates l and b, we can calculate for a point at any
distance from the sun d, the distance from that point to the GC,
which we will call R (see Fig. 7):
R = (cid:16)d2 sin2 l + (R0 − d cos l)2(cid:17)1/2
(14)
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
9
where R0 is the standard value for the distance from the sun
to the GC: R0 = 8.5 kpc. We can then calculate the angle γ,
defined in Fig. 7, by
γ = arctan R0 − d cos l
d sin l
!
(15)
The circular velocity for a distance R with 3 kpc < R < R0 is
given by Burton (1988):
1.0074 R
R0!0.0382
vc = v0
+ 0.00698
The components of the ISM velocity in the radial direction and
the direction of galactic longitude and latitude, are then given
by:
vism,r = vc cos(γ − l) cos b
vism,l = vc sin(γ − l) cos b
vism,b = vc sin b
In these expressions, we have neglected the velocity component
of the ISM perpendicular to the plane of the galaxy.
The proper motion of the star can be expressed in terms of
galactic coordinates (µl and µb). The proper motion gives rise to
spatial velocity components that are dependent of the distance.
We thus have:
(16)
(17)
(18)
v∗,r
v∗,l = µl · d
v∗,b = µb · d
We can now calculate the velocity difference between the
star and the local ISM in the three components, where we also
correct for the motion of the sun:
vr = v∗,r −(cid:0)vism,r − v0 sin l − u⊙ cos l − v⊙ sin l(cid:1)
vl = v∗,l −(cid:0)vism,l − v0 cos l + u⊙ sin l − v⊙ cos l(cid:1)
vb = v∗,b −(cid:0)vism,b + w⊙(cid:1)
The total velocity of the star with respect to the local ISM
(19)
is what we called vism in Sect. 4.1:
vism = (cid:16)v2
r + v2
l + v2
b(cid:17)1/2
Furthermore, we can also find an expression for the inclination
i and position angle ϕ:
i = arccos vr
vism!
ϕ = arctan vl
vb! + ǫ
(21)
(22)
where ǫ denotes the angle between the line of constant declina-
tion and the line of constant galactic latitude at the position of
the star.
The uncertainties in the distance and in the proper motion
of the star introduce a range of possible values for vism, i and ϕ,
which is a constraint that we can use to reduce the total num-
ber of possible orientations drastically. This makes it easier to
compare our models to the observations, as we will do in the
next section.
Fig. 8. Negative images of the emission distribution as inferred
from the Fabry-P´erot images in Fig. 2. Upper panel (a): radial
velocity against Right Ascension, Lower Panel (b): radial ve-
locity against Declination. The process of creating this image
is described in the text. The disturbances in the edges of the im-
age are artifacts caused by manipulating the image. The sharp
cutoff in the lower image at about -20′′is caused by the seeing
during the observations (see Sect. 2.2).
(20)
5. Comparison of bow shock models for
continuous mass loss with observations
In this section we will use the bow shock model discussed in
Sect. 4 to explain the distribution of emission as observed in
the Fabry-P´erot images. In order to do this properly, we trans-
formed the observational data in Fig. 2 to an image that dis-
plays the emission distribution in the Right Ascension -- radial
velocity plane. This was done by stacking the 30 different FP
images on top of each other to create a 3D body, with two spa-
tial dimensions and one radial velocity dimension. Then we
summed all layers with constant declination to get the image
that is shown in Fig. 8(a). Fig. 8(b) shows the same image, but
here the Right Ascension is replaced by the declination. These
are the images we will compare our bow shock models with
different parameters to.
We apply the properties of WR 124 to the results that we
derived in Sect. 4.3. The star has galactic coordinates l = 50.2◦,
b = +3.31◦. At this position, the angle ǫ has the value of 62.7◦.
10
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
The resemblance between the models and the data is re-
markable. It is clear that the best model can be found some-
where in the first row and the third or fourth column of Fig. 10.
From that, it can be inferred that the best values for the param-
eters are vism ≈ 180 km s−1, i ≈ 20◦ and ϕ ≈ −185◦, so that
the star is moving south, under a very small inclination with
respect to the line of sight. The resemblance between the data
and the model indicates that the values for these three param-
eters fit well, not that this distance is better than other values.
The distance was only used to obtain likely values for these pa-
rameters. As can be seen from Fig. 9, there are more distances
that can explain this combination of the three parameters. We
conclude that the bow shock model with the parameters given
above is successful in explaining the overall features of the two
velocity versus position (α or δ) plots of Fig. 8 and Fig. 10.
There are of course also differences between the model out-
put and the observations. Part of these differences can be at-
tributed to the fact that we assume a continuous mass loss from
the star, which is in reality unlikely to be the case. In fact, in
Sect. 6 we will argue that the mass loss occurred in outbursts.
Also, we assume the ISM to be homogeneous. Inhomogeneities
in the ISM will cause a less smooth bow shock surface, as in-
deed seen in the observations. The largest disagreement be-
tween model and observations is of course seen in the lower
panel of Fig. 10. This is due to the lack of observational data
for declinations less than about -20′′(see Sect. 2.2 and Fig. 8).
6. The effect of outbursts on the bow shock
So far, we have looked at a bow shock that is formed due to a
continuous wind. However, as we have already noted in Sect. 1
and 2.3, the distinct arcs that can be seen in the HST image are
likely to originate from discrete outbursts. Since these kinds
of outbursts only take place after the star has moved from the
main sequence and since the O-star that preceded WR 124 must
already have had a strong wind, these outbursts must have oc-
curred inside the bow shock. In this section we will describe
what the effects of a short outburst on the bow shock is.
In the first phase the outburst will expand freely and both
the geometrical and velocity structure are undisturbed, apart
from the extra radial velocity due to the velocity of the star.
Typical timescales for this phase are derived from division of
the distance r0 of a few parsec by a typical LBV outburst ve-
locity of 100 km s−1, which gives a few times 104 yr. After
this time, the outburst will impact onto the bow shock surface,
which will affect the observed nebula in different ways. Firstly,
the outburst will be braked or possibly even be halted by the
surface of the bow shock. This means that the radial velocity
will change drastically and that the nebula will brighten. Even
in the case of a spherical outburst, the different parts of the out-
burst will reach the bow shock at different times, so that these
effects of braking and brightening will propagate through the
nebula, starting at its top. Secondly, the gas of the outburst will
be dragged along the surface of the bow shock and eventually
adopt its velocity. This means again a difference in the veloc-
ity pattern and a distortion in the (thus far possibly symmetric)
geometry of the outburst.
Fig. 9. Theoretical properties of the orientation of the bow
shock around WR 124 as a function of the distance from the
sun: Upper panel: Total velocity of WR 124 relative to the ISM.
Middle panel: Inclination (i) of the main axis of the bow shock.
Lower panel: Position Angle (ϕ) of the main axis of the bow
shock. The solid line gives the most probable value, the dashed
lines display the uncertainties that are a result of the uncertainty
in the proper motion of the star.
The proper motion of WR 124 was measured by Hipparcos and
can be converted to galactic coordinates:
µl = −6.1 ± 2.0 mas yr−1
µb = −3.0 ± 2.0 mas yr−1 )
(23)
We can then use Eqs. (20), (21) and (22) to calculate the
velocity difference between the star and the ISM (vism), as well
as the inclination (i) and position angle (ϕ) of the main axis
of the bow shock (as defined in Fig. 4), as a function of the
distance (d) of WR 124 from the sun. The results are plotted
in Fig. 9, where the most probable value, the upper and lower
limit of the proper motion of M1-67 are depicted in the solid
and dashed lines respectively. Because of the great distance to
the star, the uncertainties in the proper motion are quite large,
so that the resulting uncertainties in the velocity components
perpendicular to the LOS are also large and increase with the
distance.
Next, we apply the results from Sect. 4.1. For d = 6.5 kpc,
Eq. (20) gives vism ≈ 174 km s−1 and we find from Eq. (6) that
r0 = 1.3 pc, where we used M = 2.45 × 10−5 M⊙ yr−1, v∞ =
710 km s−1 (Nugis & Lamers 2000) and ρism = 1 mH cm−3.
We can now build the 3D models as described in Sect. 4.2
and rotate them according to the angles found in Fig. 9.
Because of the large uncertainty in the proper motion, we cal-
culate three different models for each distance d, with the three
different values found at each line in Fig. 9 for vism, i and ϕ
and plot a contour of the model output over Fig. 8. The result
is displayed in Fig. 10, for distances of 6.0, 6.5, 7.0, 7.5, 8.0,
and 8.5 kpc.
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
11
Fig. 10. Model output for the 3D bow shock models for WR 124. Upper panel (a): Right Ascension (′′) versus radial velocity
(km s−1), lower panel (b): Declination (′′) versus radial velocity (km s−1). The grey-scaled images in the background are smoothed
versions of Fig. 8. The thin lines are contours of the data. The model output is plotted as the thick curve, that encloses the actual
bow shock in this projection and that would look like the output in Fig. 6 if plotted completely. The different columns have values
for the total velocity, inclination and the position angle for the distances of 6.0, 6.5, 7.0, 7.5, 8.0 and 8.5 kpc. The rows are the
three different values for these quantities for the lower limit, the best value and the upper limit for the proper motion respectively,
so that each row displays one line of each graph in Fig. 9. The dashed lines in each plot are at 0′′ and 199 km s−1, the position
and radial velocity of WR 124. Note the lack of observational data at declinations lower than -20′′(See Sect. 2.2).
Because the magnitude of the effect described above will
depend strongly on the relative momenta of the matter in both
the bow shock and the outburst, we divide the outbursts in three
simple categories:
1. Spherical outbursts where the momentum of the outburst is
smaller than that of the wind-ISM shock
2. Spherical outbursts where the momentum of the outburst is
larger than that of the wind-ISM shock
3. Non-spherical outbursts that may be a combination of the
two cases above: in some directions poutburst < pbowshock, in
others poutburst > pbowshock.
12
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
In the first case, the geometry and dynamics of the bow
shock will not be changed too much by the impact of the out-
burst. Those of the outburst, however, will change apprecia-
bly. The freely expanding gas will impact onto the bow shock,
brighten and be dragged along the surface layer of the bow
shock. The outburst will adopt the geometry and dynamics of
the bow shock quickly, so that it is eventually only recognis-
able as a bright, distorted ellipse in the diffuse, weak, if at all
visible, background of the gas that resides in the bow shock.
In the second case, the steady bow shock is distorted by the
impact. The outburst will be brightened and slowed down at
the moment of impact, but will continue its motion for a while.
Eventually, the outburst will be halted by the ongoing ram-
pressure of the ISM, but this will happen at a larger distance
from the star than where it first hit the bow shock. This situa-
tion can be described in the same way as the situation before
the outburst, only with a stronger wind (with lower velocity but
a much higher mass loss). This will result in a larger stagnation
point distance r0, as given by Eq. (6). The outburst will change
the shape of the bow shock temporarily, because the different
parts of the outburst will impact at different times. The distor-
tion will thus propagate through the bow shock, starting at its
top. Eventually, this results in a bow shock with the same ge-
ometry, but a different size and different dynamics. The size
simply scales with r0, but the velocities in Eq. (8) and Eq. (9)
depend on the mass loss and terminal wind velocity in a more
complex way. After the outburst, the momentum of the quiet
wind will drop back to its old value, because the outburst itself
lasts only very short compared to the evolutionary timescale of
the star in this phase. Because of this, the old value of r0 may
be reinstalled after some time.
In the third case a non-spherical outburst occurs. Many
LBV outbursts are known to be bipolar, the best known exam-
ple being η Carinae. Bipolarity of an LBV outburst can result
from binarity, or from fast rotation. Since an interacting spheri-
cal outburst as described above can at best result in an elliptical
ring, the structures in this case are already more complex. This
complexity will increase when the impact occurs. In the case
of a non-spherical outburst, it is likely that the outburst will not
hit the bow shock at its top first, but somewhere on the side,
where the surface velocity of the shock is much higher. Thus
the resulting drag will be stronger and will decrease symmetry
even further.
In all cases, part of the outburst will move toward the rear
end of the bow shock. Because this end is open, it will never
interact with the bow shock surface and keep expanding freely.
This part of the outburst can therefore be used to derive a dy-
namical timescale directly. In the velocity data of M1-67, we
find a freely expanding structure that is labelled as A in Fig. 11.
From comparison with Fig. 10 it is clear that this structure is
not located on the bow shock surface. We fitted the structure A
to an ellipse, assuming that WR 124 is at the centre of expan-
sion. From the fit, we found vexp ≈ 150 ±15 km s−1 and r ≈ 40′′
(∝ 1.3 pc at d = 6.5 kpc), which gives a dynamical timescale
of 1250 ± 125 yr kpc−1 or about 8.2 ± 0.8 × 103 yr for for the
estimated distance of 6.5 kpc for WR 124. This timescale is
typical for LBV outbursts and therefore a strong hint that the
nebula M1-67 may indeed be the result of LBV-like eruptions.
Fig. 11. The two data sets used for this research. Upper left
panel (a): HST image, upper right panel (b): Fabry-P´erot data
projected in the declination-radial velocity plane, lower left
panel (c): Fabry-P´erot data projected in the right ascension-
radial velocity plane. In the images, parts of the nebula that are
discussed in the text are encircled and labelled, for easier refer-
ence. The reader can also easily compare the positions of each
structure in the different images. Note that (roughly) the part of
the upper right image where no data is available is left blank.
It is very unlikely that the structure A is the only result of the
outburst that caused it, which we will refer to as outburst A.
The structure is clearly expelled toward us, and one would ex-
pect to find a counterpart of this structure that was caused by
the same outburst, but directed away from us. This counterpart
should be clearly visible, as long as it hasn't reached the bow
shock surface yet, because it should have a much higher radial
velocity than the star (up to +345 km s−1). Since we do not
see this counterpart, it must have collided with the bow shock
surface already. Given the expansion velocity derived above
(150 km s−1) and the stagnation point distance of r0 ≈ 1.3 pc
(Sect. 5), it would take the part of outburst A that is directed
away from us about 8.5 × 103 yr to reach the bow shock sur-
face. Within the accuracy of the fit, this time agrees with the
fitted dynamical timescale of 8.2 ± 0.8 × 103 yr. This means
that the counterpart of the freely expanding structure A might
have collided with the bow shock just a short time ago. It is
therefore very likely that the structure A originates from the
most recent outburst that occurred on WR 124.
The interactions between parts of an outburst and the steady
bow shock can cause a very chaotically looking nebula, just as
is the case for M1-67. In particular, it will generate elliptical
structures of which many can be seen in Fig. 3. Another result
of the impact of an outburst onto the bow shock, is that it is no
longer possible to simply fit an undisturbed, freely expanding
nebula to this part and derive its dynamical timescale. However,
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
13
we may still be able to estimate the dynamical timescales of
parts of the outburst that have been interacting.
The majority of the emission of M1-67 seems to originate
from the surface of the bow shock in the position-velocity di-
agram of Fig. 10. Consider the structure that we labelled B in
Fig. 11. From the Fabry-P´erot images, we derive its velocity,
as shown in Fig. 11(c). This figure, compared to the first row,
fourth column image of Fig. 10(a), shows that the structure lies
on the bow shock surface. We do not know the exact path that
it travelled to get to its current position, but we do know that
this path must be in the plane that is defined by the main axis
of the bow shock and the current position of this structure B.
We can therefore characterise the path by a single angle θ, that
is the angle between the line that joins the front of the bow
shock and the star and the direction in which the structure was
expelled, as depicted in Fig. 12(a). After the outburst B, that
caused the structure B, occurred, the ejecta first crossed the dis-
tance between the star and the bow shock surface in a straight
line and at a constant velocity. After having reached the bow
shock surface, the ejecta collided with it and were eventually
dragged away along the bow shock surface toward the rear, un-
til they reached their current position. The angle θ can have
values ranging from 0◦ (expelled exactly toward the front of
the bow shock) to about 70◦, the angle at which the structure is
observed now, which would mean that it moved from the star
straight to its current position. If we assume that the structure B
was expelled at the same outburst as the freely expanding struc-
ture A discussed above, it should have crossed the distance to
the bow shock surface with a velocity of vexp = 150 km s−1.
The velocity on the bow shock, along its surface, is given by
Eq. (10). If we then integrate over the path, we find a dynam-
ical timescale as a function of θ. The results are displayed in
Fig. 12(b). From this, we find that the dynamical timescale for
structure B must be greater than 1.2 × 104 yr, for the case where
this structure has moved directly from the star to its current po-
sition, and that it is very likely that the dynamical timescale is
in the order of 2 × 104 yr (for θ ≈ 45◦). For this θ, the timescale
would even increase to about 4 × 104 yr, if the expansion ve-
locity would be 50 km s−1. For very small θ, the dynamical
timescale goes to infinity, because the velocity along the bow
shock surface drops to zero near the front (See Fig. 5(b)).
One could debate whether the ejecta can survive the harsh
stellar environment for a few times 104 yr. Indeed, we see that
blobs of gas around young, hot stars are photoionised by the
strong radiation field of the star, after which the ionised gas
is blown away by the stellar wind. Being a Wolf-Rayet star,
WR 124 clearly has such a strong radiation field. However,
the gas cannot been blown away by the wind, since there is
an equilibrium on the surface of the bow shock between the
stellar wind and the ram pressure of the ISM. The gas therefore
cannot escape, other than in the direction toward the back of
the bow shock, as we described. In fact, the emission from M1-
67 that we see in Fig. 3 is in Hα, which indicates that the gas
is already ionised. Since we only consider the dynamics of the
gas here, it is not important what exactly happens to the gas, as
long as it does not affect its global motion.
For structure C in Fig. 11, roughly the same scenario holds
as for structure B. It is very likely to be located on the bow
Fig. 12. Upper panel (a): Geometry plot, showing the possible
paths from the star (asterisk at the origin) to an observed struc-
ture of gas on the bow shock surface (asterisk at left), char-
acterised by the angle θ. Lower panel (b): Travel time for an
ejected structure to reach the observed position as a function of
θ. The dashed line is the time needed to cross the distance from
the star to the bow shock. We assumed an expansion velocity of
vexp = 150 km s−1. The dash-dotted line is the time needed to
travel the second part of the path, along the bow shock, where
the velocity is given by Eq. (10). The solid line is the total travel
time.
shock surface. However, this structure it spread out much more
and therefore it is hard to define its exact location. Furthermore,
it is positioned at the southern cutoff limit of our velocity data.
For these reasons, we do not discuss its dynamical properties
here. We labelled it in Fig. 11 for completeness, because it is
another feature that is clearly visible in the different datasets.
Although there are still large uncertainties in the rough
calculations presented here, it seems very likely that the dy-
namical timescales of the freely expanding structure A and the
shocked structure B do not match. The explanation that both
structures originate from the same bipolar outburst, with an ex-
pansion velocity that is different in different directions seems
unlikely, since the directions to these two structures as seen
from the star do not differ too much. This means that the two
features are likely to originate from different outburst events.
This result strongly suggests a multiple-outburst scenario, with
typical outburst timescales of a few times 104 yr. This scenario
favours an LBV phase rather than an RSG phase that is respon-
sible for the nebula M1-67.
14
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
7. Summary and conclusions
We tried to fit different freely expanding models to the nebula
M1-67 and find that M1-67 is not expanding freely, but that
the outbursts that formed the nebula have been interacting with
the ISM. Evidence for this is given by the fact that the nebula
has a radial velocity that is lower than the radial velocity of
the star, in particular the fact that there is no nebular emission
with a significantly higher radial velocity than that of the star.
The asymmetry in the amount of emission between the high-
velocity and low-velocity sides of the nebula and the asymme-
try in the velocity distribution of the nebula seem to confirm
the fact that there are interactions between the outbursts and
the ISM.
We find that the central star WR 124 has a velocity of
about 180 km s−1 with respect to the surrounding ISM, mov-
ing roughly away from us, which causes a paraboloid-like bow
shock around the star. Because the star is moving away from
us, we are looking into the hollow bow shock from the rear.
Using the simple assumption of momentum conservation, we
can model the geometry of the bow shock and the velocity
of the gas that is moving along its surface. This way we can
roughly derive the orientation of the bow shock, which is de-
termined by the spatial velocities of the star and that of the
ISM surrounding it. The inclination of the main axis of the bow
shock with respect to the line of sight is about 20◦, where the
arrow that points from the star to the front of the bow shock
is mainly pointed away from us and slightly to the south. This
small inclination is responsible for the fact that we cannot dis-
tinguish the bow shock in the HST image in Fig. 3 and that
we see the star projected in the centre of the nebula. The star
sits inside this bow shock, at about 1.3 pc from its front. As
input for our bow shock models, we used the observed prop-
erties of the stellar wind ( M, v∞) and assumed an ISM density
of 1 mH cm−3. These values obviously have quite some uncer-
tainty in them. We therefore once more explicitly mention the
rough nature of these calculations. The assumptions of a con-
tinuous wind and a homogeneous ISM for our models are very
likely to be incorrect. In fact, an ISM with inhomogeneities
could explain why the observed bow shock does not have a
smooth surface.
In general, a run-away star with a wind will form a bow
shock. The LBV eruptions during the late stages of the evo-
lution of massive stars will partly collide with the bow shock
surface, brighten up and be dragged away by the ram pressure
of the ISM, along the surface of the bow shock. In this way,
irregular, or even chaotic nebulae with arcs and rings can be
formed by intrinsically nicely behaving, possibly even spher-
ical, discrete outbursts. The interactions with the ISM make
the derivation of dynamical timescales for the outbursts much
more difficult than in the case of low-velocity stars, which blow
voids of tens of parsecs in the interstellar medium during their
O-star phase, so that outbursts occurring in a later stage will in-
deed expand freely. However, also in the case of a bow shock, a
part of each outburst will move toward the rear end of this bow
shock and remain expanding freely forever. This part might still
be fitted in the straightforward way.
For M1-67 we were able to fit the freely expanding part of
the outburst A in Fig. 11 to r ≈ 40′′ and vexp ≈ 150 km s−1,
which gives the rough dynamical timescale of 8 × 103 yr, as-
suming a distance of 6.5 kpc. This dynamical timescale is of the
same order as the time it would take this outburst to cross the
distance of 1.3 pc from the star to the front of the bow shock,
which means that the far part of this particular outburst might
just have collided with the bow shock. In fact, the collision of
this part of the nebula is required in order to explain the lack
of nebular emission with significantly higher radial velocities
than that of the star.
Part of the nebular emission is found to be on the bow shock
surface. In order to get there, this matter could have followed
different paths, because we do not know in which direction the
material has been ejected from the star. We can calculate the
amount of time for this matter to have reached its observed
location, as a function of the path. We did so for the arc that
is labelled B in Fig. 11, assuming that the expansion velocity
was the same as for the freely expanding structure A, vexp =
150 km s−1. We then find that the age of structure B must be
greater than 1.2 × 104 yr, probably in the order of 2 × 104 yr,
so that this arc most likely originates from a different outburst
than structure A.
The multiple outbursts, the expansion velocity of about
150 km s−1 and the dynamical timescales in the order of 104 yr
are all typical for LBVs. We therefore conclude that an LBV
phase preceded the current Wolf-Rayet phase. The outbursts
that occurred during this LBV phase seem to be the most likely
explanation for the creation of the Wolf-Rayet ring nebula M1-
67 around WR 124.
Acknowledgements. We thank Thierry le F`evre for his contribution to
the freely expanding wind models, Antonella Nota for the long-slit
spectra and Yves Grosdidier for the Fabry-P´erot images.
References
Bertola, F. 1964, PASP, 76, 241+
Burton, W. B. 1988, in Galactic and Extragalactic Radio
Astronomy, 295 -- 358
Canto, J., Raga, A. C., & Wilkin, F. P. 1996, ApJ, 469, 729+
Chu, Y.-H. & Treffers, R. R. 1981, ApJ, 249, 586
Grosdidier, Y., Moffat, A. F. J., Joncas, G., & Acker, A. . 1998,
ApJ, 506, L127
Grosdidier, Y., Moffat, A. F. J., Joncas, G., & Acker, A. 1999,
in ASP Conf. Ser. 168: New Perspectives on the Interstellar
Medium, 453+
Lamers, H. J. G. L. M. & Cassinelli, J. P. 1999, Introduction
to stellar winds (Introduction to stellar winds / Henny
J.G.L.M. Lamers and Joseph P. Cassinelli. Cambridge ;
New York : Cambridge University Press, 1999.
ISBN
0521593980)
Lamers, H. J. G. L. M., Nota, A., Panagia, N., Smith, L. J., &
Langer, N. 2001, ApJ, 551, 764
Langer, N., Hamann, W.-R., Lennon, M., et al. 1994, A&A,
290, 819
Merrill, P. W. 1938, PASP, 50, 350+
Mignard, F. 2000, A&A, 354, 522
Van der Sluys & Lamers: The dynamics of the nebula M1-67 around the star WR 124
15
Minkowski, R. 1946, PASP, 58, 305+
Nota, A. & Clampin, M. 1997,
in ASP Conf. Ser. 120:
Luminous Blue Variables: Massive Stars in Transition, 303+
Nugis, T. & Lamers, H. J. G. L. M. 2000, A&A, 360, 227
Pismis, P. & Recillas-Cruz, E. 1979, Revista Mexicana de
Astronomia y Astrofisica, 4, 271
Sirianni, M., Nota, A., Pasquali, A., & Clampin, M. 1998,
A&A, 335, 1029
Solf, J. & Carsenty, U. 1982, A&A, 116, 54
Stothers, R. B. & Chin, C. 1993, ApJ, 408, L85
-- . 1996, ApJ, 468, 842+
|
astro-ph/9904175 | 1 | 9904 | 1999-04-13T19:19:46 | Gravitational lensing statistics with extragalactic surveys. II. Analysis of the Jodrell Bank-VLA Astrometric Survey | [
"astro-ph"
] | We present constraints on the cosmological constant $\lambda_{0}$ from gravitational lensing statistics of the Jodrell Bank-VLA Astrometric Survey (JVAS). Although this is the largest gravitational lens survey which has been analysed, cosmological constraints are only comparable to those from optical surveys. This is due to the fact that the median source redshifts of JVAS are lower, which leads to both relatively fewer lenses in the survey and a weaker dependence on the cosmological parameters. Although more approximations have to be made than is the case for optical surveys, the consistency of the results with those from optical gravitational lens surveys and other cosmological tests indicate that this is not a major source of uncertainty in the results. However, joint constraints from a combination of radio and optical data are much tighter. Thus, a similar analysis of the much larger Cosmic Lens All-Sky Survey should provide even tighter constraints on the cosmological constant, especially when combined with data from optical lens surveys.
At 95% confidence, our lower and upper limits on $\lambda_{0}-\Omega_{0}$, using the JVAS lensing statistics information alone, are respectively -2.69 and 0.68. For a flat universe, these correspond to lower and upper limits on \lambda_{0} of respectively -0.85 and 0.84. Using the combination of JVAS lensing statistics and lensing statistics from the literature as discussed in Quast & Helbig (Paper I) the corresponding $\lambda_{0}-\Omega_{0}$ values are -1.78 and 0.27. For a flat universe, these correspond to lower and upper limits on $\lambda_{0}$ of respectively -0.39 and 0.64. | astro-ph | astro-ph | A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
02(12.07.1; 12.03.4; 12.03.3)
ASTRONOMY
AND
ASTROPHYSICS
9
9
9
1
r
p
A
3
1
1
v
5
7
1
4
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Gravitational lensing statistics with extragalactic surveys
II. Analysis of the Jodrell Bank-VLA Astrometric Survey
Phillip Helbig1, Daniel Marlow1 ⋆, Ralf Quast2, Peter N. Wilkinson1, Ian W. A. Browne1, and
L. V. E. Koopmans3
1 University of Manchester, Nuffield Radio Astronomy Laboratories, Jodrell Bank, Macclesfield, Cheshire SK11 9DL, UK
2 Universitat Hamburg, Hamburger Sternwarte, Gojenbergsweg 112, D-21029 Hamburg, Germany
3 University of Groningen, Kapteyn Astronomical Institute, Postbus 800, NL-9700 AV Groningen, The Netherlands
Received December 2, 1998; accepted January 12, 1999
Abstract. We present constraints on the cosmological
constant λ0 from gravitational lensing statistics of the
Jodrell Bank-VLA Astrometric Survey (JVAS). Although
this is the largest gravitational lens survey which has been
analysed, cosmological constraints are only comparable to
those from optical surveys. This is due to the fact that the
median source redshifts of JVAS are lower, which leads to
both relatively fewer lenses in the survey and a weaker de-
pendence on the cosmological parameters. Although more
approximations have to be made than is the case for opti-
cal surveys, the consistency of the results with those from
optical gravitational lens surveys and other cosmological
tests indicate that this is not a major source of uncertainty
in the results. However, joint constraints from a combina-
tion of radio and optical data are much tighter. Thus, a
similar analysis of the much larger Cosmic Lens All-Sky
Survey should provide even tighter constraints on the cos-
mological constant, especially when combined with data
from optical lens surveys.
At 95% confidence, our lower and upper limits on
λ0 − Ω0, using the JVAS lensing statistics information
alone, are respectively −2.69 and 0.68. For a flat universe,
these correspond to lower and upper limits on λ0 of re-
spectively −0.85 and 0.84. Using the combination of JVAS
lensing statistics and lensing statistics from the literature
as discussed in Quast & Helbig (1999) the corresponding
λ0−Ω0 values are −1.78 and 0.27. For a flat universe, these
correspond to lower and upper limits on λ0 of respectively
−0.39 and 0.64.
Key words: gravitational lensing -- cosmology: theory --
cosmology: observations
Send offprint requests to: P. Helbig
⋆ Present Address: University of Pennsylvania, Dept. of
Physics and Astronomy, 209 S. 33rd Street, Philadelphia, PA
19104-6396, U.S.A.
Correspondence to: [email protected]
1. Introduction
The use of gravitational lensing statistics as a cosmological
tool was first considered in detail by Turner et al. (1984);
the influence of the cosmological constant was investigated
thoroughly by Fukugita et al. (1992), building on the work
of Turner (1990) and Fukugita et al. (1990). Kochanek
(1996a, and references therein) and, more recently, Falco
et al. (1998, hereafter FKM) have laid the groundwork
for using gravitational lensing statistics for the detailed
analysis of extragalactic surveys. Quast & Helbig (1999,
hereafter Paper I) reanalysed optical surveys from the lit-
erature, for the first time exploring a range of the λ0-Ω0
parameter space large enough to enable a comparison with
other cosmological tests. Here, we use the formalism out-
lined in Paper I to analyse the Jodrell Bank-VLA Astro-
metric Survey (JVAS), the largest completed gravitational
lens survey to date.
Radio surveys offer several advantages over optical sur-
veys (see, e.g., FKM): one doesn't have to worry about
systematic errors due to extinction or a lens galaxy of ap-
parent brightness comparable to that of the lensed images
of the source, the resolution (of followup observations if
not of the survey proper) is much smaller than the typ-
ical image separation, parent catalogues in the form of
large-area surveys exist from which unbiased samples can
be selected and relatively easily observed. Disadvantages
in the radio are due to our relatively poor knowledge of
the flux density-dependent redshift distribution or equiva-
lently the redshift-dependent number-magnitude relation.
For a description of our method see Paper I. The plan
of this paper is as follows. Sect. 2 describes the JVAS
gravitational lens survey. In Sect. 3 we describe the cal-
culations we have done based on the JVAS data. Sect. 4
presents our results, using both the JVAS data alone and
in combination with the results from the optical surveys
analysed in Paper I. Finally in Sect. 5 we compare our re-
sults to those of Paper I and present our conclusions and
our prognosis for the analysis of future large surveys such
as CLASS.
2
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
2. The JVAS Gravitational Lens Survey
2.1. The sample
The Jodrell Bank-VLA Astrometric Survey (JVAS) is a
survey for flat-spectrum radio sources with a flux den-
sity greater than 200 mJy at 5 GHz. Flat-spectrum radio
sources are likely to be compact, thus making it easy to
recognise the lensing morphology. In addition, they are
likely to be variable, making it possible to determine H0
by measuring the time delay between the lensed images.
(See Biggs et al. (1999) for the description of a time delay
measurement in a JVAS gravitational lens system.) JVAS
is also a survey for MERLIN phase-reference sources and
as such is described in Patnaik et al. (1992), Browne et al.
(1998) and Wilkinson et al. (1998). JVAS as a gravita-
tional lens survey, the lens candidate selection, followup
process, confirmation criteria and a discussion of the JVAS
gravitational lenses is described in detail in King et al.
(1999) (see also King & Browne 1996).
In order to have a parent sample which is as large as
possible and as cleanly defined as practical, our "JVAS
gravitational lens survey sample" is slightly different than
the "JVAS phase-reference calibrator sample". For the for-
mer, the source must be a point source and must have a
good starting position (so that the observation was cor-
rectly pointed) while its precise spectral index is not im-
portant. For the latter, only the spectral index is impor-
tant, as the source can be slightly resolved or the observa-
tion can be less than perfectly pointed. Thus, the JVAS as-
trometric sample (Patnaik et al. 1992; Browne et al. 1998;
Wilkinson et al. 1998) contains 2144 sources. To these
must be added 103 sources which were too resolved to be
used as phase calibrators and 61 sources which had bad
starting positions (thus the observations were too badly
pointed to be useful for the astrometric sample), bringing
the total to 2308. This formed our gravitational lens sam-
ple, since these additional sources were also searched for
gravitational lenses (King et al. 1999) (none were found
meeting the JVAS selection criteria).
2.2. The lenses
We have used the gravitational lens systems in Table 1
in this analysis. The JVAS lens B1938+666 (King et al.
1998) was not included because it is not formally a part of
the sample, having a too steep spectral index and having
been recognised on the basis of a lensed extended source as
opposed to lensed compact components. Also, the JVAS
lens B2114+022 (Augusto et al. 1999) was not included
because it is not a single-galaxy lens system.
3. Calculations
A major difference between the analysis of an optical sur-
vey (see Paper I and references therein) and a radio survey
is that in the latter one does not know the redshifts of all
the unlensed sources. One can still use the formalism of
Paper I, however, substituting for the non-lensed objects
in the sample a subsample with known redshifts, multiply-
ing the logarithm of this contribution from the non-lenses
to the likelihood by the ratio of the size of the parent sam-
ple to that of the subsample. Alternatively, one can take
the redshifts from a sample selected according to the same
criteria, assigning these randomly to objects in (a subsam-
ple of) the parent sample for a similar flux density range.
Similarly, one does not know the number-magnitude rela-
tion for the sample and for its extension to fainter flux den-
sities (needed to allow for the lens amplification). Again,
this can be estimated from either a subsample (through
extrapolation) or from another sample selected according
to the same criteria (either through extrapolation or by
having a fainter flux density limit in this other sample;
in the latter case obviously the selection criteria should
be identical to that of the original sample except for the
lower flux density limit).
For this analysis, due to the paucity of the observa-
tional data, we have made rather stark assumptions: the
redshift distribution of the sample is assumed to be iden-
tical to that of the CJF sample (Taylor et al. 1996), inde-
pendent of flux density, and the number-magnitude rela-
tion is assumed to be identical to that of CLASS (Cosmic
Lens All-Sky Survey, Myers et al. 1999), independent of
redshift.
Otherwise, we have followed the procedure outlined in
Paper I, calculating the a priori likelihood of obtaining
the observational data as a function of λ0 and Ω0 and
the a posteriori likelihood for the three different choices of
prior information used in Paper I. We present results both
for the JVAS lens survey and for the combination of the
JVAS results with those from the optical surveys analysed
in Paper I.
4. Results and discussion
The left panel of Fig. 1 shows the constraints on the cos-
mological parameters λ0 and Ω0 based only on the in-
formation obtained from the JVAS lens statistics, while
the right panel shows the joint constraints from the JVAS
lens sample and the optical samples from Paper I. Fig. 2
is identical except that one of the input parameters, the
normalisation of the galaxy luminosity function, was in-
creased by two standard deviations. This gives an idea of
the magnitude of systematic uncertainties. (See the dis-
cussion in Paper I.)
The left plot in the top row of Fig. 3 shows the joint
likelihood of our lensing statistics analysis and that ob-
tained by using conservative estimates for H0 and the age
of the universe (see Paper I). Although neither method
alone sets useful constraints on Ω0, their combination
does, since the constraint from H0 and the age of the uni-
verse only allows large values of Ω0 for λ0 values which
are excluded by lens statistics. Even though the 68% con-
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
3
Table 1. JVAS lenses used in this analysis. Of the information in the table, for this analysis we use only the source redshift zs
and the image separation ∆θ
# images
2 + ring
4
2
4
∆θ[′′]
0.334
2.09
1.56
1.28
zl
0.6847
0.9584
0.599
0.337
zs
0.96
2.639
1.535
3.62
lens galaxy
spiral
elliptical
spiral
?
Name
B0218+357
MG0414+054
B1030+074
B1422+231
2.
1.5
PSfrag replacements
0
Ω
1.
PSfrag replacements
0
Ω
0.5
0
p
F
(
D
)
=
,
0.3
-4.
-2.
0
2.
λ
0
p
F
(
D
)
=
,
0.3
2.
1.5
1.
0.5
0
-4.
-2.
0
2.
λ
0
Fig. 1. Left panel: The likelihood function p(Dλ0, Ω0, ξ0) based on the JVAS lens sample. All nuisance parameters are assumed
to take precisely their mean values. The pixel grey level is directly proportional to the likelihood ratio, darker pixels reflect
higher ratios. The pixel size reflects the resolution of our numerical computations. The contours mark the boundaries of the
minimum 0.68, 0.90, 0.95 and 0.99 confidence regions for the parameters λ0 and Ω0. Right panel: Exactly the same as the left
panel, but the joint likelihood from the JVAS lens sample and the optical samples from Quast & Helbig (1999, Paper I)
fidence contour still allows almost the entire Ω0 range, it
is obvious from the grey scale that much lower values of
Ω0 are favoured by the joint constraints. The upper limit
on λ0 changes only slightly while, as is to be expected, the
lower limit becomes tighter. Right plot: exactly the same,
but including optical constraints from Paper I. The upper
limits on λ0 decrease slightly, while the lower limits im-
prove considerably. The latter is probably due to the fact
that, in addition to just using more data the JVAS sources
are at significantly different redshifts than those from the
optical surveys analysed in Paper I (the JVAS sources are
generally at lower redshift). The former is consistent with
the slightly higher optical depth for radio surveys found
by FKM and will be discussed more below.
The middle row of Fig. 3 shows the effect of including
our prior information on Ω0 (see Paper I). As is to be
expected, (for both the JVAS and combination data sets)
lower values of Ω0 are favoured. This has the side effect
of weakening our lower limit on λ0 (though only slightly
affecting the upper limit). This should not be regarded
as a weakness, however, since including prior information
for λ0 and Ω0 from the constraint from H0 and the age
of the universe as well as for Ω0 itself, as illustrated in
the bottom row of Fig. 3, tightens the lower limit again
(without appreciably affecting the upper limit).
We believe that the right plot of the bottom row of
Fig. 3 represents very robust constraints in the λ0-Ω0
plane. The upper limits on λ0 come from gravitational
lensing statistics, which, due to the extremely rapid in-
crease in the optical depth for larger values of λ0, are
quite robust and relatively insensitive to uncertainties in
the input data (cf. Fig. 2 and the discussion of the effect
of changing the most uncertain input parameter by 2 σ in
Paper I) as well as to the prior information used (compare
the upper, lower and middle rows of Fig. 3). The combi-
nation of data from JVAS and optical surveys leads to
much tighter lower limits on λ0 than using either alone.
The upper and lower limits on Ω0 are based on a number
of different methods and appear to be quite robust (see
Paper I). The combination of the relatively secure knowl-
edge of H0 and the age of the universe combine with lens
statistics to produce a good lower limit on λ0, although
this is to some extent still subject to the caveats mentioned
above.
If one is interested in the allowed range of λ0, one can
marginalise over Ω0 to obtain a probability distribution
for λ0. This is illustrated in Fig. 4 and Table 2.
4
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
2.
1.5
PSfrag replacements
0
Ω
1.
PSfrag replacements
0
Ω
0.5
0
p
F
(
D
)
=
,
0.3
-4.
-2.
0
2.
λ
0
p
F
(
D
)
=
,
0.3
2.
1.5
1.
0.5
0
-4.
-2.
0
2.
λ
0
Fig. 2. Exactly the same as Fig. 1, but the parameter ne is increased by two standard deviations. This parameter, the normal-
isation of the luminosity function of the lens galaxies, is one of the more uncertain input parameters, thus one can get a rough
estimate of the overall uncertainty by comparing this figure and Fig. 1. See the discussion in Paper I
Table 2. Marginal mean values, standard deviations and 0.95 confidence intervals for the parameter λ0 on the basis of the
marginal distributions shown in the top row of Fig. 4
Sample
JVAS
JVAS
JVAS
JVAS
joint
joint
joint
joint
Distribution
p(Dλ0)
p1(λ0D)
p2(λ0D)
p3(λ0D)
p(Dλ0)
p1(λ0D)
p2(λ0D)
p3(λ0D)
Mean
0.13
0.44
−0.29
0.11
0.19
0.24
−0.25
−0.09
standard deviation
1.08
0.77
0.98
0.64
0.70
0.63
0.59
0.48
95% c.l. range
−2.08
−1.05
−2.38
−1.20
−1.17
−0.96
−1.46
−1.08
1.91
1.87
1.17
1.16
1.48
1.46
0.77
0.77
information
1.42
1.32
1.45
1.98
1.95
1.96
Table 3. Mean values and ranges for assorted confidence levels for the parameter λ0 for our a priori and various a posteriori
likelihoods from this work for Ω0 = 0.3. This should be compared to Table 3 in Paper I
Cosmological test
JVAS, p(Dλ0)
JVAS, p1(λ0D)
JVAS, p2(λ0D)
JVAS, p3(λ0D)
JVAS & optical, p(Dλ0)
JVAS & optical, p1(λ0D)
JVAS & optical, p2(λ0D)
JVAS & optical, p3(λ0D)
68% c.l. range
90% c.l. range
95% c.l. range
−0.66
−0.44
−1.38
−0.69
−0.54
−0.63
−1.02
−0.77
0.72
0.80
0.86
0.86
0.26
0.40
0.44
0.52
−1.68
−1.00
−2.81
−1.45
−1.08
−0.95
−1.72
−1.23
0.87
0.92
1.00
0.99
0.44
0.53
0.63
0.63
−2.36
−1.38
−3.70
−1.89
−1.41
−1.18
−2.08
−1.52
0.96
1.00
1.06
1.03
0.54
0.62
0.72
0.70
99% c.l. range
−3.91
−2.27
< −5.00
−2.91
−2.15
−1.72
−2.95
−2.15
1.08
1.09
1.15
1.15
0.70
0.73
0.80
0.79
The comparison values from this work corresponding
to those in Tables 3 and 4 of Paper I are presented in
Tables 3 and 4.
As mentioned in Paper I, to aid comparisons with other
cosmological tests, the data for the figures shown in this
paper are available at
For a "likely" Ω0 value of 0.3 we have calculated the
likelihood with the higher resolution ∆λ0 = 0.01. This is
show in Fig. 5. From these calculations one can extract
confidence limits which, due to the higher resolution in
λ0, are more accurate. These are presented in Table 5 and
should be compared to those for p(Dλ0) from Table 3.
http://multivac.jb.man.ac.uk:8000/ceres
/data_from_papers/JVAS.html
and we urge our colleagues to follow our example.
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
5
2.
1.5
PSfrag replacements
0
Ω
1.
PSfrag replacements
0
Ω
p
F
(
D
)
=
,
0.3
0.5
0
2.
1.5
-4.
-2.
0
2.
λ
0
p
F
(
D
)
=
,
0.3
PSfrag replacements
0
Ω
1.
PSfrag replacements
0
Ω
p
F
(
D
)
=
,
0.3
0.5
0
2.
1.5
-4.
-2.
0
2.
λ
0
p
F
(
D
)
=
,
0.3
PSfrag replacements
0
Ω
1.
PSfrag replacements
0
Ω
0.5
0
p
F
(
D
)
=
,
0.3
-4.
-2.
0
2.
λ
0
p
F
(
D
)
=
,
0.3
2.
1.5
1.
0.5
0
2.
1.5
1.
0.5
0
2.
1.5
1.
0.5
0
-4.
-2.
0
2.
λ
0
-4.
-2.
0
2.
λ
0
-4.
-2.
0
2.
λ
0
Fig. 3. Left column: The posterior probability density functions p1(λ0, Ω0D) (top panel), p2(λ0, Ω0D) (middle panel) and
p3(λ0, Ω0D) (bottom panel). All nuisance parameters are assumed to take precisely their mean values. The pixel grey level
is directly proportional to the likelihood ratio, darker pixels reflect higher ratios. The pixel size reflects the resolution of our
numerical computations. The contours mark the boundaries of the minimum 0.68, 0.90, 0.95 and 0.99 confidence regions for
the parameters λ0 and Ω0. The respective amounts of information obtained from our sample data are I1 = 1.42, I2 = 1.32 and
I3 = 1.45. Right column: Exactly the same as the left panel, but the joint likelihood from the JVAS lens sample and the optical
samples from Quast & Helbig (1999). The respective amounts of information obtained from our joint sample data are 1.98, 1.95
and 1.96. See Paper I for definitions
6
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
PSfrag replacements
)
0
λ
(
p
Ω
F
D
=
,
0.3
PSfrag replacements
)
0
λ
(
F
Ω
p
D
=
,
0.3
-4
-2
0
2
λ
0
-4
-2
0
2
λ
0
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
PSfrag replacements
)
0
λ
(
p
Ω
F
D
=
,
0.3
PSfrag replacements
)
0
λ
(
F
Ω
p
D
=
,
0.3
-4
-2
0
2
λ
0
-4
-2
0
2
λ
0
Fig. 4. Left column: The top panel shows the normalised marginal likelihood function p(λ0D) (light gray curve) and the marginal
posterior probability density functions p1(Dλ0) (medium gray curve), p2(Dλ0) (dark gray curve) and p3(Dλ0) (black curve)
derived from the JVAS analysis. All nuisance parameters are assumed to take precisely their mean values. The bottom panel
shows the respective cumulative distribution functions. Right column: Exactly the same as the left panel, but the joint likelihood
from the JVAS lens sample and the optical samples from Quast & Helbig (1999)
Table 4. Mean values and ranges for assorted confidence levels for the parameter λ0 for our a priori and various a posteriori
likelihoods from this work for k = 0. This should be compared to Table 4 in Paper I
Cosmological test
JVAS, p(Dλ0)
JVAS, p1(λ0D)
JVAS, p2(λ0D)
JVAS, p3(λ0D)
JVAS & optical, p(Dλ0)
JVAS & optical, p1(λ0D)
JVAS & optical, p2(λ0D)
JVAS & optical, p3(λ0D)
68% c.l. range
90% c.l. range
95% c.l. range
99% c.l. range
−0.11
0.13
0.35
0.41
−0.15
0.02
0.39
0.39
0.70
0.75
0.77
0.79
0.45
0.54
0.39
0.51
−0.83
−0.15
0.13
0.25
−0.49
−0.12
0.09
0.18
0.78
0.82
0.83
0.83
0.55
0.61
0.59
0.63
< −1.00
−0.33
0.02
0.16
−0.69
−0.29
0.00
0.09
0.82
0.85
0.85
0.85
0.60
0.64
0.64
0.66
< −1.00
−0.69
−0.21
−0.04
< −1.00
−0.60
< −0.22
−0.09
0.86
0.89
0.88
0.88
0.67
0.70
0.70
0.72
5. Conclusions and outlook
We have used the method outlined in Quast & Helbig
(1999) to measure the cosmological constant λ0 from the
lensing statistics of the Jodrell Bank-VLA Astrometric
Survey. At 95% confidence, our lower and upper limits
on λ0-Ω0, using the JVAS lensing statistics information
alone, are respectively −2.69 and 0.68. For a flat universe,
these correspond to lower and upper limits on λ0 of re-
spectively −0.85 and 0.84. Using the combination of JVAS
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
7
1
0.8
)
3
.
0
=
0.6
1
0.8
)
3
.
0
=
0.6
PSfrag replacements
0
Ω
,
0
λ
D
(
p
0.4
p
PSfrag replacements
-4
-2
0
2
λ
0
-4
-2
0
2
λ
0
Fig. 5. Left panel: The likelihood function as a function of λ0 for Ω0 = 0.3 and with all nuisance parameters taking their default
values, using just the JVAS data. Right panel: The same but plotted cumulatively
1
0.8
)
3
.
0
=
0.6
1
0.8
)
3
.
0
=
0.6
PSfrag replacements
0
Ω
,
0
λ
D
(
p
0.4
p
PSfrag replacements
F
0.2
0
F
0.2
0
0
Ω
,
0
λ
D
(
F
0.4
0.2
0
0
Ω
,
0
λ
D
(
F
0.4
0.2
0
-4
-2
0
2
λ
0
-4
-2
0
2
λ
0
Fig. 6. As Fig. 5 but combining optical and radio data. Left panel: The likelihood function as a function of λ0 for Ω0 = 0.3 and
with all nuisance parameters taking their default values. Right panel: The same but plotted cumulatively
Table 5. Confidence ranges for λ0 assuming Ω0 = 0.3. Unlike the results presented in Table 3, these figures are for a specific
value of Ω0 and not the values of intersection of particular contours with the Ω0 = 0.3 line in the λ0-Ω0 plane. These are more
appropriate if one is convinced thatΩ0 = 0.3 and have been calculated using ten times better resolution than the rest of our
results presented in this work. See Figs. 5 and 6
data set
JVAS
JVAS+optical
68% c.l. range
90% c.l. range
95% c.l. range
99% c.l. range
−0.69
−0.65
0.72
0.30
−1.72
−1.17
0.91
0.49
−2.39
−1.48
0.98
0.57
−3.83
−2.22
1.06
0.70
lensing statistics and lensing statistics from the literature
as discussed in Quast & Helbig (1999) the corresponding
λ0−Ω0 values are −1.78 and 0.27. For a flat universe, these
correspond to lower and upper limits on λ0 of respectively
−0.39 and 0.64. Note that the lower limit is affected more
than the upper limit with respect to the difference between
the JVAS results and those in Paper I and with respect
to combining the JVAS results with those from Paper I.
Our determination is consistent with other recent mea-
surements of λ0, both from lensing statistics and from
other cosmological tests (see Quast & Helbig 1999, Pa-
per I, for a discussion). We confirm the result of Falco
et al. (1998, FKM) that radio surveys give higher val-
ues of λ0 than optical surveys. Cooray et al. (1999) and
Cooray (1999) obtain a 95% confidence upper limit on
λ0 in a flat universe of 0.79 from analyses of the Hubble
8
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
Deep Field and CLASS. However, these analyses suffer
from systematic effects due to our ignorance of the un-
derlying flux density-dependent redshift distribution (or,
equivalently, the redshift-dependent luminosity function)
of the unlensed parent population. As discussed in Cooray
(1999), the value of λ0 obtained from CLASS will decrease
if the mean redshift of the sample is lower than presumed.
Thus, although there is no real conflict at present as the
lower limits on λ0 are not as tight, it seems not unlikely
that a more detailed analysis of CLASS, incorporating
more information about the unlensed parent population,
will result in a value more in line with our value obtained
from the JVAS analysis. Of course, the JVAS analysis also
suffers from systematic effects, but the general agreement
between the results obtained from the analysis of optical
surveys (cf. Paper I and references therein) and radio sur-
veys as presented here and in FKM suggests that these
are not overwhelming. Also, the difference, a higher value
of λ0 from radio surveys, is what one would expect, as
lens systems which go unnoticed will, all other things be-
ing equal, reduce the value of λ0. This could be the case
in optical surveys since it is possible that extinction in
the lens galaxy and the fact that the resolution is only
slightly better than the image separation could lead to
lens systems being missed. Again, the general agreement
does suggest though that these effects are not overwhelm-
ing.
Of course, one could imagine that the agreement is
coincidental, the optical surveys being heavily affected by
extinction and resolution bias and the radio surveys by our
ignorance of the unlensed parent population. However, the
fact that lens statistics in general gives results which are
not in conflict with other cosmological tests (cf. Paper I)
suggests that this is not the case. Moreover, extinction
would bias the results from lens statistics and the m-z re-
lation (e.g. for type Ia supernovae, cf. the results in Tables
3 and 4 of paper I and in the references mentioned there)
in the opposite direction. Thus, their agreement suggests
that both methods have their systematics more or less
under control.
The major source of uncertainty in radio lens surveys
is the lack of knowledge about the redshift distribution
and number-magnitude relation of the source sample (e.g.
Kochanek 1996b). We are currently undertaking the neces-
sary observations to reduce this source of systematic error.
Since the time scale for this project is comparable to that
for the followup of the CLASS survey, there seems little
point in doing a better analysis of JVAS in the future,
especially since CLASS is defined so that JVAS is essen-
tially a subset of it.1 The larger size of the CLASS survey,
1 The definition of both is flat-spectrum between L-band and
C-band, i.e. α > −0.5 where sf ∼ f α, the essential difference
being the lower flux density limit of 200 mJy for JVAS and
30 mJy for CLASS. However, since CLASS is defined based on
newer catalogues (GB6 and NVSS: Gregory et al. 1996; Condon
et al. 1998) than JVAS, there will be some essentially random
coupled with better knowledge of the redshift distribu-
tion and number-magnitude relation of the source sample,
should reduce both the random and systematic errors on
our value of λ0.
Acknowledgements. We thank our collaborators in the JVAS,
CJF and CLASS surveys for useful discussions and for pro-
viding data in advance of publication and many colleagues at
Jodrell Bank for helpful comments and suggestions. We also
thank John Meaburn and Anthony Holloway at the Depart-
ment of Astronomy in Manchester and the staff at Manchester
Computing for providing us with additional computational re-
sources. RQ is grateful to the CERES collaboration for making
possible a visit to Jodrell Bank where this collaboration was
initiated. This research was supported in part by the European
Commission, TMR Programme, Research Network Contract
ERBFMRXCT96-0034 "CERES".
References
Augusto P., Browne I.W.A., Wilkinson P.N., et al., 1999,
MNRAS, in preparation
Biggs A., Browne I.W.A., Helbig P., et al., 1999, MNRAS,
304, 349
Browne I.W.A., Patnaik A.R., Wilkinson P.N., Wrobel J.,
1998, MNRAS, 293, 257
Condon J.J., Cotton W.D., Greisen E.W., et al., 1998, AJ,
115, 1693
Cooray A.R., 1999, A&A, 342, 353
Cooray A.R., Quashnock J.M., Miller M.C., 1999, ApJ,
511, 562
Falco E., Kochanek C.S., Munoz J.A., 1998, ApJ, 494, 47
Fukugita M., Futamase T., Kasai M., 1990, MNRAS, 246,
24
Fukugita M., Futamase K., Kasai M., Turner E.L., 1992,
ApJ, 393, 3
Gregory P.C., Scott W.K., Douglas K., Condon J.J., 1996,
ApJS, 103, 427
King L.J., Browne I.W.A., 1996, MNRAS, 282, 67
King L.J., Jackson N.J., Blandford R.D., et al., 1998, MN-
RAS, 295, L41
King L.J., Browne I.W.A., Marlow D.R., Patnaik A.R.,
Wilkinson P.N., 1999, MNRAS, in press
Kochanek C.S., 1996a, ApJ, 466, 638
Kochanek C.S., 1996b, ApJ, 473, 595
Myers S.T., Rusin D., Marlow D., et al., 1999, in prepa-
ration
Patnaik A.R., Browne I.W.A., Wilkinson P.N., Wrobel
J.M., 1992, MNRAS, 254, 655
differences due to differing quality of observations and variabil-
ity of the sources. All the JVAS lenses mentioned in Table 1
are in the new CLASS sample, which, having no upper flux
density limit, subsumes JVAS. The previous samples CLASS-I
and CLASS-II will be similarly subsumed in the same sense as
JVAS, though the differences here will be slightly larger since
bands other than L and C were used in the preliminary defini-
tion of these samples.
P. Helbig et al.: Gravitational lensing statistics with extragalactic surveys. II
9
Quast R., Helbig P., 1999, A&A, 344, 721, in press
Taylor G.B., Vermeulen R.C., Readhead A.C.S., et al.,
1996, ApJS, 107, 37
Turner E.L., 1990, ApJ, 365, L43
Turner E.L., Ostriker J.P., Gott III J.R., 1984, ApJ, 284,
1
Wilkinson P.N., Browne I.W.A., Patnaik A.R., Wrobel J.,
Sorothia B., 1998, MNRAS, 300, 790
|
astro-ph/9508110 | 2 | 9508 | 1995-10-25T05:00:57 | Galactic Center Molecular Arms, Ring and Expanding Shells.I. Kinematical Structures in Longitude-Velocity Diagrams | [
"astro-ph"
] | Analyzing the (l, b, Vlsr) data cube of 13CO(J=1-0) line emission obtained by Bally et al, we have investigated the molecular gas distribution and kinematics in the central +/-1 deg (150 pc) region of the Galaxy. We have applied the pressing method to remove the local- and foreground-gas components at low velocities in order to estimate the intensity more quantitatively. Two major dense molecular arms have been identified in longitude-radial velocity diagrams as apparently rigidly-rotating ridges. The ridges are spatially identified as two arms, which we call the Galactic Center molecular Arms (GCA). The arms compose a rotating ring of radius 120 pc (the 120-pc Molecular Ring), whose inclination is about 85 deg. The Sgr B molecular complex is associated with GCA I, and Sgr C complex is located on GCA II. These arms are as thin as 13 to 15 pc, except for vertically extended massive complexes around Sgr B and C. The LV behavior of the arms can be qualitatively reproduced by a model which assumes spiral arms of gas. Assuming a small pitch angle for the arms, we tried to deconvolve the LV diagram to a projection on the galactic plane, and present a possible face-on CO map as seen from the galactic pole, which also reveals a molecular ring and arms. We have estimated masses of these molecular features, using the most recent value of the CO-to-H2 conversion factor taking into account its metallicity dependency and radial gradient in the Galaxy. The estimated molecular masses and kinetic energy are about a factor of three smaller than those reported in the literature using the conventional conversion factor. | astro-ph | astro-ph |
Galactic Center Molecular Arms, Ring and Expanding Shells. I
Kinematical Structures in Longitude-Velocity Diagrams
Yoshiaki SOFUE
Institute of Astronomy, The University of Tokyo
Mitaka, Tokyo 181, Japan
[email protected]
(To appear in PASJ Vol. 47, No.5)
Abstract
Analyzing the (l, b, VLSR) data cube of 13CO(J = 1 − 0) line emission obtained by
Bally et al, we have investigated the molecular gas distribution and kinematics in the
central ±1◦ (±150 pc) region of the Galaxy. We have applied the pressing method to
remove the local- and foreground-gas components at low velocities in order to estimate the
intensity more quantitatively. Two major dense molecular arms have been identified in
longitude-radial velocity (l, V ) diagrams as apparently rigidly-rotating ridges. The ridges
are spatially identified as two arms, which we call the Galactic Center molecular Arms
(GCA). The arms compose a rotating ring of radius 120 pc (the 120-pc Molecular Ring),
whose inclination is i ≃ 85◦. The Sgr B molecular complex is associated with GCA I,
and Sgr C complex is located on GCA II. These arms are as thin as 13 to 15 pc, except
for vertically extended massive complexes around Sgr B and C. The (l, V ) behavior of
the arms can be qualitatively reproduced by a model which assumes spiral arms of gas.
Assuming a small pitch angle for the arms, we tried to deconvolve the (l, V ) diagram to a
projection on the galactic plane, and present a possible face-on CO map as seen from the
galactic pole, which also reveals a molecular ring and arms. We have estimated masses of
these molecular features, using the most recent value of the CO-to-H2 conversion factor
taking into account its metallicity dependency and radial gradient in the Galaxy. The
estimated molecular masses and kinetic energy are about a factor of three smaller than
those reported in the literature using the conventional conversion factor.
Key words: Galactic center -- Galaxy -- Interstellar matter -- Molecular gas -- Spiral arms.
1. Introduction
The galactic center region has been extensively observed in the molecular lines, par-
1
ticularly in the CO line emission (Oort 1977; Scoville et al 1974; Liszt 1988; Liszt and
Burton 1978, 1980; Burton and Liszt 1983, 1993; Brown and Liszt 1984; Heiligman 1987;
Bally et al 1987; 1988; Genzel and Townes 1987; Stark et al 1989; Gusten 1989). Besides
the 4-kpc molecular ring, the CO emission is strongly concentrated in the central a few
degree (Dame et al 1987). Moreover, the molecular gas in the central region has a strong
concentration within l < 1◦ (150 pc) where the majority of the nuclear disk gas is con-
fined (Scoville et al 1974; Bally et al 1987; Heiligman 1987). This high concentration of
dense interstellar matter in a small region is also clearly visible in the far IR emission (e.g.,
Cox and Laureijs 1989) and in the CII emission (Okuda et al 1989). The radio continuum
emission also indicates a highly concentrated nuclear disk of ionized gas (Altenhoff et al
1978; Handa et al 1987). On the other hand, the region between galactocentric distances
∼ 200 pc (l ∼ 1◦4) and ∼ 2 kpc (15◦) appears almost empty in the CO emission (Bally et
al 1987; Knapp et all 1985).
The total molecular mass in the l < 1◦ region estimated from the CO emission
amounts to ∼ 1.4×108M⊙• for a traditional CO-to-H2 conversion factor, or, more probably,
∼ 4.6 × 107M⊙• for a new conversion factor (see section 3). On the other hand, the HI mass
within the 1.2 kpc tilted disk (l < 8◦) is only of several 106M⊙• (Liszt and Burton 1980).
Hence, we may consider that the central ∼ 1 kpc region is dominated by the molecular disk
of ∼ 150 pc (∼ 1◦) radius, outside of which the gas density becomes an order of magnitude
smaller.
Various molecular gas features in the central ∼ 100−200 pc region have been discussed
by various authors, such as a disk related to the 1.2 kpc tilted rotating disk (Liszt and
Burton 1980; Burton and Liszt 1992), molecular rings and spiral arms of a few hundred pc
scale (Scoville et al 1974; Heiligman 1987; Bally et al 1987), and the expanding molecular
ring of 200 pc radius (Scoville 1972; Kaifu et al 1972, 1974). On the other hand, Binney
et al (1991) have modeled the "expanding-ring feature" or the "parallelogram" on the
(l, V ) (longitude-velocity) plot in terms of non-circular kinematics of gas by a closed orbit
model in a bar potential.
It is known that the gas in this parallelogram shares only a
small fraction of the total molecular mass in the galactic center: the majority of the gas
composes more rigid-body like features in the (l, V ) plots.
The CO gas in the central 100 - 200 pc regions in nearby galaxies have been observed
by high-resolution mm-wave interferometry, and their distribution and kinematics have
been extensively studied (e.g., Lo et al 1984; Ishiguro et al 1989; Ishizuki et al 1990a,b).
The central gas disks of galaxies appear to comprise spiral arms or circum-nuclear rings
of a few hundred pc size. Such a gaseous behavior can be reproduced to some extent by
theoretical simulations of accretion of gas clouds in a central gravitational potential (e.g.,
Noguchi 1988; Wada and Habe 1992).
In this paper, we revisit the major part of the nuclear molecular disk (l <∼ 1◦)
by analyzing the molecular line data in the premise that the nuclear disk may comprise
accretion ring or spiral structures similar to those found in external galaxies. In this paper
we reanalyze the data cube of the 13CO (J = 1 − 0)-line emission observed by Bally et al
(1987) with the 7-m off-set Cassegrain telescope of the Bell Telephone Laboratory. The
distance to the Galactic Center is assumed to be 8.5 kpc throughout this paper.
2
2. Longitude-Velocity (l, V ) Diagrams
2.1. Data
The angular resolution of the observations with the Bell-Telephone 7-m antenna at
13CO line was 1′.7. The data used here are in a (l, b, VLSR) cube in FITS format. The
cube covers an area of −1◦.1 ≤ l ≤ 0◦.92, −21′ ≤ b ≤ 17′, or 300 pc × 94 pc region
for a 8.5 kpc distance. The velocity coverage is −250 ≤ VLSR ≤ 250 km s−1. The cube
comprises 127, 39, and 183 channels at 1′, 1′, and 2.75 km s−1 intervals, respectively, We
also use the CS line data in a (l, b, V ) cube with dimensions 151, 42, and 163 channels at
intervals 2′, 1′, and 2.75 km s−1, which covers an area of −1◦ ≤ l ≤ 4◦, −25′ ≤ b ≤ 16′,
and −250 ≤ VLSR ≤ 190 km s−1. The intensity scale of the data is the main-beam
antenna temperature approximately equivalent to brightness temperature in Kelvin. The
observational details are described in Bally et al (1987, 1988). We made use of the AIPS
and IRAF software packages for the reduction.
2.2. Subtraction of the Local and Foreground Components
In order to analyze molecular gas features in the Galactic Center region, we first
subtract contaminations by local and foreground molecular clouds at low velocities. Since
13CO line is optically thin for the foreground clouds, the contaminations appear as emission
stripes superposed on the galactic center emission, The subtraction of foreground emission
is essential when we derive the mass and kinetic energy of molecular gas features. Such
a "cleaning" also helps much the morphological recognition of features on the (l, V ) and
(b, V ) diagrams.
Fig. 1a shows an (l, V ) diagram averaged in a latitude range of −17′ ≤ b ≤ 12′. The
diagram is strongly affected by "stripes" at a low velocities elongated in the direction of
longitude with narrow velocity widths, including the 3-kpc expanding arm at −52 km s−1.
In order to eliminate these stripes, we applied the "pressing method" as developed for
removing scanning effects in raster scan observations (Sofue and Reich 1979). We briefly
describe this method below.
-- Fig. 1a, b --
The original (l, V ) map M0 is trimmed by −70 ≤ VLSR ≤ 50 km s−1 to yield M1 where
the local and foreground gas contribution is significant. The trimmed map M1 is smoothed
only in the V direction by 5 channels (14 km s−1) using a boxcar or Gaussian smoothing
task, yielding M2. Smoothed map M2 is subtracted from M1 to yield M3 (= M1 − M2).
Map M3 is then smoothed only in l direction by 20 channels (20′) (boxcar or Gaussian) to
yield M4. This M4 map approximates the contribution from the local gas that is dominated
by elongated features in the longitudinal direction. We then subtract M4 from M1 to obtain
M5 (= M1 − M4). This M5 is, thus, a map in which the local gas contribution has been
roughly subtracted. M5 is then smoothed in V direction by 5 channels. Then we replace
M2 by this smoothed map, and repeat the above procedures twice (or more times) until
3
we obtain the second (or n-th) M5. Finally, the −70 ≤ VLSR ≤ 50 km s−1 part of the
original map M0 is replaced by M5 to yield M6. Now, we have a "pressed" map M6 in
which corrugations due to local gas clouds have been removed out. Fig. 1b shows the thus
obtained map M6 for the same (l, V ) diagram as in Fig. 1a.
We have applied this algorithm (the pressing method) to all (l, V ) and (b, V ) diagrams
in the cube, and created a new (l, b, V ) cube, which is almost free from local and foreground
contaminations. In the present paper we use this new cube. We also applied the pressing
method to remove scanning effects, which had originated during the data acquisition, in
every diagram such as intensity maps in the (l, b) space.
By comparing the original and the thus 'pressed' maps, we estimated the contribution
of the local/foreground emission to be 5% of the total emission, and 9% of the emission
with VLSR < 100 km s−1. Thus, without the subtraction, the mass and energetics would
be overestimated by about 5 to 9%. Moreover, if the gas out of the disk component at
b > 10′ is concerned, this local contribution would amount to more than 10%. Hence,
the subtraction of the foreground emissions is crucial in a quantitative discussion of the
features discussed in this paper.
2.3. "Arms" in Longitude-Velocity (l, V ) Diagrams
Fig. 2 shows (l, V ) diagrams near the galactic plane averaged in 4′ latitude interval
after subtraction of the local/foreground components. Various features found in these
diagrams have been discussed in Bally et al (1987, 1988).
In this paper we highlight
continuous features (ridges) traced in the (l, V ) diagrams. The major structures of the
"disk component" at low latitude (b <∼ 10′=25 pc) are "rigid-rotation" ridges, which we
call "arms". Fig. 3 illustrate these ridges (arms) which can be identified in the diagrams as
coherent structures. In Table 1 we summarize the identified features, and describe below
the individual arms. Heiligman (1987) has used these ridges to derive a rigid rotation
curve. At higher latitudes (b >∼ 10′) the so-called expanding ring features at high
velocities (VLSR > 100 km s−1), which will be discussed in a separate paper.
-- Fig. 2 --
-- Fig. 3 --
-- Table 1 --
2.3.1. Arm I
The most prominent (l, V ) arm is found as a long and straight ridge, slightly above
the galactic plane at b ∼ 2′, which runs from (l, V )=(0◦.9, 80 km s−1) to (−0◦.7, −150
km s−1), and extends to (−1◦.0, −200 km s−1). This arm intersects the line at l = 0◦ at
negative velocity VLSR = −40 km s−1, indicating that the gas is approaching us at l = 0◦.
We call this ridge Arm I. A part of this arm can be traced also below the galactic plane
at b = −0.1◦, running from (l, V )=(0.8◦, 60 km s−1) to (0◦.1, −20 km s−1). Its positive
longitude part is connected to the dense molecular complex Sgr B, which is extended both
in space and velocity, from b = −0.25 to 0.07◦ and VLSR=20 to 100 km s−1.
4
2.3.2. Arm II
Another prominent arm is seen at negative latitude at b ∼ −6′, running from (l, V )=(0◦.1,
60 km s−1) to (−0◦.6, −80 km s−1). We call this ridge Arm II. It is bent at l ∼ 0◦.1 and
appears to continue to (l, V )=(1◦, 100 km s−1), and merges with Arm I at the Sgr B
complex region. The negative longitude part also merges with Arm I, and is connected to
the Sgr C complex. Arm II intersects l = 0◦ at positive velocity of VLSR = 50 km s−1.
2.3.3. Arms III and IV
At positive latitude (b ∼ 0◦.01 to 0◦.2), another arm can be traced running from
(l, V )=(0◦, 140 km s−1) to (−0◦.15, 10 km s−1). Its counterpart to the negative longitude
side appears to be present at (l, V )=(−0◦.45, −120 km s−1) to (−0◦.55, −180 km s−1). We
call this ridge Arm III. Bally et al (1988) called this the "polar arc", and discussed its
connection to Sgr A.
A branch can be traced from (l, V )=(0◦.1, 60 km s−1) to (0◦, −20 km s−1), apparently
being bifurcated from Arm II at l ∼ 0◦.1. This ridge intersects l = 0◦ at negative velocity
(VLSR − 50 km s−1). We call this ridge Arm IV.
2.4. "Rigid-rotation" in (l, V ) Plane and "Arms and Ring" in the Galactic
Plane
We emphasize that "rigid-rotation" ridges in (l, V ) diagrams for edge-on galaxies,
whose rotation curves are usually flat, are generally interpreted as due to spiral arms
and rings. Indeed, the rigid-rotation ridge in the CO (l, V ) diagram of the Milky Way
is identified with the 4-kpc molecular ring (e.g., Dame et al 1987; Combes 1992). Many
edge-on spiral galaxies like NGC 891 are found to show similar (l, V ) ridges in HI and CO,
which are also interpreted to be spiral arms and rings (e.g., Sofue and Nakai 1993, 1994).
The circular rotation velocity as defined by Vrot = (R∂Φ/∂R)1/2, remains greater than
at least 150 km s−1 from the nuclear few pc region till the 1 kpc radius region (Genzel and
Townes 1987). Here, Φ is the potential and R is the distance from the nucleus. Hence,
the actual rotation should not be rigid at all: The rigid-rotation ridges in the (l, V ) plane
such as Arms I to IV in the Galactic Center can thus be more naturally attributed to real
arms and rings.
3. Intensity Distribution and the Galactic Center Arms
3.1. Intensity Maps and Masses
Fig. 4a shows the total intensity map integrated over the full range of the velocity
5
(−250 ≤ VLSR ≤ 250 km s−1). This map is about the same as that presented by Stark
et al (1989), except that the local gas has been removed. Fig. 4b shows the same in grey
scale and a that with the vertical scale in b direction enlarged twice.
-- Fig. 4 --
First of all the intensity map can be used to obtain the molecular mass. However, the
conversion of the CO intensity to H2 mass is not straightforward. We have recently studied
the correlation of the conversion factor X12 for the 12CO(J = 1 − 0) line with the metal
abundance in galaxies (Arimoto et al 1994). We have obtained a clear dependency of X
on the galacto-centric distance R within individual galaxies, which is almost equivalent to
the metallicity dependence. For the Milky Way we have
X12(R) = 0.92(±0.2) × 1020exp(R/Re)
where Re = 7.1 kpc is the scale radius of the disk. Applying this relation to the Galac-
tic center, we obtain a conversion factor at the Galactic center as X12 = 0.92(±0.2) ×
1020 [H2 cm−2/K kms s−1], about one third of the solar vicinity value. We then assume
that the 12CO and 13CO intensities are proportional, and estimate the ratio by aver-
aging observed intensity ratios for the inner Galaxy at l ≤ 20◦ (Solomon et al 1979);
I12CO/I13CO ≃ 6.2 ± 1.0. Then, we obtain a conversion factor for the 13CO line intensity
in the galactic center region as X13(R = 0) ≃ 5.7 × 1020 [H2 cm−2/K kms s−1], and we
use this value throughout this paper.
The correction factor from the H mass to real gas mass is given by µ = 1/X = 1.61,
where X is the hydrogen abundance in weight. Here, the following relation has been
adopted (Shaver et al 1983): Y = 0.28 + (∆Y /∆Z)Z = 0.34 in weight, where Z = 0.02
is the abundance of the heavy elements and ∆Y /∆Z = 3 is the metallicity dependence
of the helium abundance Y in the interstellar matter, and so, the hydrogen abundance
is X = 0.62. So, the surface mass density of molecular gas after correction for the mean
weight of gas is given by
where
σ ∼ 14.6(±3) I/η [M⊙•pc−2],
I ≡Z T ∗
Adv [K km s−1]
(2)
(3)
is the integrated intensity of 13CO(J = 1 − 0) line emission and η = 0.89 is the primary
beam efficiency of the antenna. The total mass of molecular gas (including He and metals)
can be estimated by
M [M⊙•] = 14.6Z I/ηdxdy [K km s−1 pc2].
(4)
Using these relations, we have estimated the total molecular gas mass in the observed
area (−1◦.0 ≤ l ≤ 0◦.92, −21′ ≤ b ≤ 17′) after removing the local and foreground
contribution to be 4.6(±0.8) × 107M⊙•. We have also estimated the total molecular mass
of the "disk" component, which comprises most of the ridge-like features in the (l, V )
diagrams, excluding the expanding ring feature (or the parallelogram) at high velocities
6
(VLSR >∼ 100 − 150 km s−1). The disk component has the mass 3.9 × 107M⊙•, which is
85% of the total in the observed region. On the other hand, the expanding ring (or the
parallelogram) shares only 6.7 × 106M⊙• (15%) in the region at l < 1◦.
3.2. Ring and Arms in Intensity Maps
In order to clarify if each of the arms traced in the (l, V ) diagrams (Fig. 1-3), par-
ticlurly Arms I and II, is a single physical structure in space, we have obtained velocity-
integrated intensity map in the (l, b) plane for each of the arms. Thereby, we integrated
the CO intensity in the velocity ranges as shown in Fig. 5 individually for Arms I and II.
Fig. 6 show the integrated intensity maps corresponding to Arms I and II, together with
a summation of I and II. In Table 1 we summarize the derived parameters.
-- Fig. 5 --
-- Fig. 6 --
3.2.1. Galactic Center Arms I, II
In the intensity map, Arm I can be traced as a single, thin arc-like arm from l = 0◦.9
near the Sgr B complex toward negative longitude at l = −1◦.0. We call this spatial arm
the Galactic Center Arm I (GCA I). The angular extent is as long as 1◦.9 (280 pc) in the
longitudinal direction, whereas the thickness in the b direction is as thin as ∼ 5′ (13 pc;
see section 3.2.2). The Sgr B molecular complex is much extended in the b direction by
about 0◦.4 (60 pc), and composes a massive part of the arm. A "return" of this arm can
be traced from (l, b)=(0◦.9, 0◦) to (0◦.2, −0◦.07), and is more clearly recognized in Fig. 4
at V = 83 ∼ 167km s−1. This can be also clearly seen in the (l, V ) diagram in Fig. 2 at
b ∼ −0◦.1. In the negative l side, the arm appears to be bifurcated at l ∼ −0◦.65, and
linked to Arm II. This can be more clearly observed in Fig. 4 at V = −83 ∼ 0 km s−1.
The intensity in Fig. 6a has been integrated to give a total mass of molecular gas involved
in GCA I (in the velocity range as shown in Fig. 5a) to be M = 1.72 × 107M⊙•.
Arm II can be traced as a a single bright ridge from l ∼ 0◦.3 to −0◦.7, and the
thickness is about 6′ (15 pc), and makes GCA II. The mass of Arm II is estimated to
be 1.35 × 107M⊙•. Thus, the total mass involved in GCA I and II is estimated to be
3.07 × 107M⊙•, and shares almost 67% of the total gas mass in the observed region, and
78% of the disk component.
3.2.2. The 120-pc Molecular Ring
As shown in Fig. 4 and 6, GCA I and II compose a global ring structure, which is
tilted and slightly bent. If we fit the GCA I and II by a ring, its angular extent in the
major axis is 1◦.6 from l = −0◦.7 to 0◦.9, and so, the major axis length (diameter) is 240
pc, and the radius 120 pc. The minor axis length is estimated to be 7′.9 from the maximum
separation between Arm I and II at l ∼ −0◦.2 (see Fig. 7). Therefore, the inclination of
the I+II ring is i = 85◦ from the minor-to-major axis ratio. The center of the ring, as
7
fitted by the above figures, is at (l, b)=(0◦.1, 0◦.0) We call this ring the 120-pc Molecular
Ring.
From these we conclude that the spatial distribution of the molecular gas associated
with the principal ridges in the (l, V ) diagrams comprises a circum-nuclear ring of radius
R ≃ 120 pc inclined by 5◦ from the line of sight.
3.2.3. Cross Section of the Arms
Fig. 7 shows the intensity variation perpendicular to the galactic plane across GCA
I and II averaged from l = 0◦.24 to −0◦.33, where the arms are most clearly separated.
Since the effective resolution of the present data is (θ2 + ∆b2)1/2 = 2′.0, where θ = 1′.7
is the beam width and ∆b = 1′.0 is the grid interval, the arms are sufficiently resolved.
The two peaks in the figure at b = 1′.8 (Arm I) and b = −6′.0 (Arm II) can be fitted by a
Gaussian intensity distributions as (TB, peak, FWHM) = (0.27 K, 5′.3) and (0.33 K, 5′.5),
respectively. Namely, the arms are as thin as 13.0 (GCA I) and 13.5 pc (GCA II). If we
subtract the contributions from these two arm components, the residual intensity in the
whole area in Fig. 7 is only 36% of the total intensity. The intensity coming from the inter-
arm region between the arms shares only 12% of the intensity from the two arms. This
would be an upper limit, as the region displayed in this figure is the weakest part along
the arms without any significant molecular clumps and condensations. Thus, we conclude
that the molecular gas as observed in the CO line emission in the region discussed in this
paper is almost totally confined within the two major arms. Therefore, the central 100 pc
radius region is almost empty, making a hole of molecular gas, except the nuclear few pc
region surrounding Sgr A.
-- Fig. 7 --
3.3. Velocity Field
Fig. 8a shows a velocity field as obtained by taking the first moment of the (VLSR, l, b)
cube, and therefore, an intensity-weighted velocity field. A general rotation characteristics
is clearly seen along the major axis of the ring feature at b ≃ −6′. Sgr C molecular spur is
seen as a negative velocity spur extending toward negative b. GCA III is seen as the tilted
high-velocity plume at (l, b) ∼ (0◦.2, 0◦.1).
In addition to these individual velocity structures, a large-scale velocity gradient in
the latitude direction is prominent in the sense that the positive b side has positive velocity
and negative b side negative velocity. This can be attributed to the fact that the high-
velocity expanding shell (ring) is more clearly seen in positive velocity at b > 0◦, while
the negative velocity part more clearly at b < 0◦ (see section 4). This can be explained by
a tilted nature of the expanding oblate molecular shell, as will be discussed in section 4
based on an analysis of (b, V ) diagrams. In fact, if we construct a velocity field, excluding
the expanding ring features, we obtain a rather regular velocity field as shown in Fig. 8b.
-- Fig. 8 --
8
3.4. Possible Models for the Galactic Center Arms and Ring
We here try to reproduce the (l, V ) diagram based on a simple spiral arm model.
According to the galactic shock wave theory (Fujimoto 1966; Roberts 1969) and the bar-
induced shock wave theory (Sorensen et al 1976; Huntley et al 1978; Roberts et al 1979;
Noguchi 1988; Wada and Habe 1992), flow vectors of gas in the densest part along the
shocked arms are almost parallel to the potential valley that is rigidly rotating at a pattern
speed slower than the galactic rotation. In such shocked flows, the gas cannot be on a closed
orbit, but is rapidly accreted toward the center along deformed spirals.
As the simplest approach to simulate an (l, V ) diagram, we assume that the flow vector
of gas is aligned along a spiral with a constant velocity equal to the rotation velocity in the
potential. Fig. 9a shows a model, where we have assumed two symmetrical spiral arms with
a pitch angle p = 10◦. In addition to a constant circular rotation of gas (Vrot =constant;
flat rotation curve), radial infall motion of Vrotsin p is superposed, so that the gas is flowing
along the arms into the central region. The density distribution in the arms are shown
by the spiral-like contours. The azimuthally averaged density of gas has a hole at the
center, or it corresponds to a ring distribution of gas on which two arms are superposed. A
calculated (l, V ) diagram is shown by the superposed contours with a tilted X shape. The
characteristic features in the observed (l, V ) diagrams can be now qualitatively reproduced.
Fig. 9b and 10c show cases where the spiral arms are oval in shape whose major axis is
inclined by ±30◦ from the nodal line. Such a case may be expected when the oval potential
or a bar in the center is deep enough to produce a non-circular motion. Fig. 9d-f are the
same, but the density distribution along the arms has the maximum at the center and
the pitch angle is taken larger: p = 20◦. Again, the case of a circular rotation appears
to reproduce the observation, while the oval orbit cases result in more complicated (l, V )
plots than the observation. Among these models, the case shown in Fig. 9a or 9b appears
to reproduce the observed characteristics in the (l, V ) plot (e.g. Fig. 3) reasonably well.
The model in Fig. 9d or 9e with the averaged gas density increasing toward the center
may explain observed Arms III and IV. However, the cases corresponding to Fig. 9c and
9f may be excluded.
-- Fig. 9 --
According to the galactic shock wave theory in a spiral density wave or a bar potential,
the shocked gas looses its azimuthal velocity so that the (l, V ) behavior becomes closer to
the potential's pattern speed. As the consequence, the apparent rotation velocity of the
gas along shocked spiral arms is smaller than that from the rotation velocity. This may be
the reason why the observed maximum velocities of the rings/arms (e.g., Arm I near Sgr
B) are less than that expected from the gravitational potential.
3.5. Deconvolution into Projection on the Galactic Plane: A Face-on View
We may thus assume that the molecular gas is on a ring or spiral arms whose pitch
angle is not so large. Then, it is possible to deconvolve the (l, V ) diagram into a spatial
distribution in the galactic plane by assuming an approximately circular rotation. Thereby,
we make use of the velocity-to-space transformation (VST), which has been extensively
9
applied to derive the HI gas distribution in our Galaxy (Oort et al 1957). Suppose that a
gas element is located at a projected distance x (≃ l × 8.5 kpc) along the galactic plane
from the center of rotation, and has a radial velocity v. If the rotation is circular at velocity
V0, the line of sight distance y of the element from the nodal line can be calculated by
y = ±xs(cid:18) V0
v (cid:19)2
− 1.
(5)
The signs must be opposite for Arms I and II. Here, we assume that Arm I is near side,
and Arm II far side, so that the signs are −/+, respectively. The center of rotation is
assumed to be at Sgr A, and v is measured from the intersection velocity at l = −0◦.06 on
each arm ridge.
Fig. 10 shows a thus obtained "face-on" map of the molecular gas for V0 = 150 km
s−1. The arms appear to construct a circum-nuclear ring of radius ∼ 120 pc. Here, we used
(l, V ) diagrams averaged within latitude ranges −2′ ≤ b ≤ 6′ for Arm I and −5′ ≤ b ≤ 3′ for
Arm II, so that vertically extended clumps such as Sgr B complex are only partly mapped
in this figure. During the deconvolution, we used only the arm component concentrated
near the ridges within ±20 km s−1 in velocity (as illustrated in Fig. 5). Diffuse gas and
clumps with velocities far from the arms are not taken into account. The same VST was
applied to the HII regions Sgr B1, B2 and C using their H recombination line velocities
(Downes et al 1980). We plotted their positions in Fig. 10. The HII regions lie along
the arms associated with the molecular complexes, though slightly avoiding the molecular
gas peaks. Sgr B and C appear to be at symmetrically opposite locations with respect to
the nucleus. We have assumed that Arm I is near side. However, in this kind of simple
deconvolution, we cannot distinguish the exact orientation, as is the case of deconvolution
of gas distribution inside the solar circle from kinematical information. Hence, it may be
possible to assume an opposite configuration of the arm locations: Arm I in far side, and
Arm II in near side.
-- Fig. 10 --
The connection of Arms I and II is not clear from this deconvolution. This is mainly
because of the ambiguous position determination near the node, which arises from unknown
precise rotation curve. The error is also large at l <∼ 0◦.1, where we applied interpolation
from both sides along each arm. Obviously, this kind of deconvolution is not unique, but it
was possible here because of the separation of Arms I and II in the (l, b) plane. Therefore,
this deconvolution should be taken as a possible hint to the spatial distribution of gas.
3.6. Comparison with Other Galaxies and Models
Accretion spirals, either shocked or not, and rings of molecular gas have been indeed
observed in the CO line in many extragalactic systems such as IC 342 (Lo et al 1984;
Ishizuki et al 1990a) and NGC 6946 (Ishizuki et al 1990b). The ring structure of molecular
gas of 100 to a few hundred pc size is commonly observed in the central regions of spiral
galaxies (Nakai et al 1987; Ishiguro et al 1989). See Sofue (1991) for a more number of
galaxies with a nuclear molecular ring. Thus, the ring/spiral structure of molecular gas
10
of radius 120 pc in the Milky Way, would be similar to the situation found in external
galaxies.
There have been various numerical simulations of the accretion of gas toward the
central region in spiral and oval potential by gas-dynamical simulations (Sorensen et al
1976; Huntley et al 1978; Roberts et al 1979; Noguchi 1988; Wada and Habe 1992). The
models predict a rapid accretion of gas along spiral orbits, and the gas behavior in these
models somehow mimic the models illustrated in Fig. 9.
A number of simulations of position-velocity diagrams along the galactic plane have
been constructed and compared with the observations, in order to understand larger-scale
(l, V ) diagrams for our Galaxy both in HI and CO (Mulder and Liem 1986; Liszt and
Burton 1978; Burton 1988). Position-velocity diagrams for extragalactic edge-on galaxies
in CO have been extensively studied (Sofue and Nakai 1993, 1994; Sofue 1994) and a
numerical simulation has been attempted to reproduce the PV characteristics based on
the gas dynamics in an oval potential (e.g., Mulder and Liem 1986; Wada et al 1994).
Binney et al (1991) have noticed the "parallelogram" and calculated theoretical (l, V )
diagrams, and have shown the presence of a bar of 2 kpc length in the Galactic bulge.
However, the parallelogram (the expanding ring feature) in Fig. 1b shares only 15% of the
total emission. However, we emphasize that the major structures, which contain 85% of
the molecular mass within 150 pc of the center, are due to the Arms discussed above.
3.7. Relationship with Radio Sources
Fig. 11 shows superposition of a 10-GHz radio continuum map (Handa et al 1987) on
the 13CO and CS intensity maps. We here briefly comment on a global relationship of the
major radio sources with molecular features at a spatial resolution of a few arc minutes.
Detailed internal structures of individual sources are out of the scope of the present paper,
for which the readers may refer to a review by Liszt (1988).
3.7.1. Sgr A
The relationship of molecular features of scales less than a few arc minutes with Sgr
A has been discussed by many authors (e.g., Oort 1977; Bally et al 1987; Gusten 1989).
However, these nuclear features, which are of 1′ (∼ 3 pc) scales, are not well visible in the
present plots in so far as the (l, V ) plots are concerned. We only mention that Arm III
is a largely tilted out-of-plane plume with high positive velocity, which Bally et al (1988)
called the polar arc. Arm IV shows also large velocity gradient, and appears to be an
object related to a deep gravitational potential around the nucleus.
3.7.2. Sgr B
The molecular complex at l ∼ 0◦.6 − 0◦.9 on Arm I is associated with the star forming
regions Sgr B1 at (l, b) = (0◦.519, −0◦.050), and Sgr B2 at (0◦.670, −0◦.036). Sgr B1 and
B2, whose radial velocities in H recombination line emission are VLSR=45 and 65 km s−1,
11
respectively (Downes et al 1980), are also located in the (l, V ) plane at the upper (higher-
velocity) edges of molecular clumps. Thus, the de-convolved positions of these continuum
sources are slightly displaced from the de-convolved arm, as indicated in Fig. 10. The
molecular gas distribution is highly extended in the direction of latitude for about 0◦.4
(60 pc), largely shifted toward the lower side of the galactic plane (b < 0◦). This complex
is also much extended in the velocity space: the velocity dispersion amounts to as high
as 50 km s−1. The internal structure of Sgr B molecular complex has been discussed in
detail in relation to the star formation activity, and it was shown that the molecular gas
is distributed in a shell, spatially surrounding the continuum peak (Bally et al 1988; Sofue
1990; Hasegawa et al 1993). The present ring model is consistent with the CII line (l, V )
diagram as obtained by Okuda et al (1989), which indicates a rotating ionized gas feature
with Sgr B and C on the tangential points of the ring.
-- Fig. 11 --
3.7.3. Sgr C
The star forming region Sgr C is associated with a molecular complex, and is located
on Arm II at l ∼ −0◦.6. However, the spatial proximity is less significant than that for Sgr
B: The radio continuum peak of Sgr C, (l, b, VLSR) = (−0◦.57, −0◦.09), is located at the
western edge of the molecular complex, but displaced by about 6′(15 pc) from the molecular
peak. The LSR velocity of the H recombination line also agrees with the molecular gas
velocity, and so, it is located on the de-convolved arm in Fig. 10. The molecular gas in
this complex is extended vertically, and molecular spurs are found to extend both toward
positive and negative latitude directions. We emphasize that the positive-latitude spur is
clearly associated with the inner edge of the western ridge of the Galactic Center Lobe
observed in the radio continuum emission (Sofue and Handa 1984; Sofue 1985), as is shown
in Fig. 11.
3.7.4. Orbital Displacement vs Alignment of Star Forming Regions and Molec-
ular Arms
The close association of Sgr B and C with GCA I and II may have a crucial implication
for the orbits of gas and stars: If the arms are shock lanes in a bar during a highly non-
circular motion, the HII regions of a million years old should already be displaced from
the molecular arms. Therefore, the fact that Sgr B and C are still near the gas complexes
from which they may have been born (after one or more rotations) can be explained only if
the stars and gas are circularly co-rotating in the arms at a small pitch angle. This would
argue for the validity of the deconvolution process applied in section 3.5.
Consider a spiral arm which is a shocked density wave. Star formation from a molec-
ular cloud will be triggered in the arms. It will take about t ∼ 106 years for proto stars to
form and shine as OB stars, and therefore, until HII regions are produced. On the other
hand, the rotation period of the stars is only ∼ 106 years for r = 100 pc and Vrot = 200
km s−1. According to the density wave theory, the velocity difference between the rotation
velocity and the shocked gaseous arm, which is about the same as the pattern speed of
12
density wave, is of the order of
Vrot − Vp = (Ωrot − Ωp)r.
The azimuthal phase difference between the HII region and the gaseous arm is then
∆φ ∼ (Ωrot − Ωp)t.
(6)
(7)
The phase difference for Sgr B2 and its corresponding molecular peak in Fig. 10
(darkest part in Arm I) is roughly ∆φ ∼ 5◦, and a similar value is found for Sgr C. If
t ∼ 106 yr, we obtain Ωrot − Ωp ∼ 0.1 radian/106 years ∼ 100 km s−1 kpc−1. This is
an order of magnitude greater than the value near the solar circle (∼ 10 km s−1 kpc−1).
For older HII regions (weaker radio sources) the phase difference would be much greater.
Moreover, orbits of stars, and therefore, HII regions, are no longer closed, and must be
largely displaced from the orbits of gas. Thus, the HII regions in the central 100 pc of
the Galaxy, except for young cases as Sgr B2, would not be associated with molecular gas
arms. This will simply explain why the molecular gas features are not directly correlated
with the weaker radio sources in the Galactic center (Fig. 11).
4. Discussion
By analyzing the 13CO line BTL data cube, we have shown that most (85%) of
the total molecular gas within l < 1◦ comprises rigid-body-like structures in the (l, V )
diagrams, which can be attributed to arms on a ring. Moreover, 66% of the total gas in
the region, and 78% of the disk component (b <∼ 10′=25 pc), was found to be confined
in the two major Arms I and II. The spiral/ring structures are consistent with the picture
drawn by Scoville et al (1974) based on the earlier data, while the scale obtained here is
slightly smaller. The structures will be common in external galaxy nuclei in the sense that
the gas distribution is spiral- and ring-like.
Numerical simulations for a few kpc scale disks have suggested that the features would
be understood as the consequence of spiral accretion by a density wave in an oval potential,
either shocked or not. Based on qualitative consideration, we have suggested possible
models to explain the observed (l, V ) features as shown in Fig. 9a.
The molecular mass in the Galactic Center has been derived by usin the most recent
CO-to-H2conversion factor about one third of the conventional value, which has been
obtained by detailed analyses of the dependency on the metallicity as well as on the
galacto-centric distance (Arimoto et al 1994). This has resulted in a factor of three smaller
mass and energetics than the so far quoted values in the literature: The molecular mass
within 150 pc radius from the center is estimated to be only 3.9 × 107M⊙•.
Thus, the molecular gas mass is only a few percent of the total mass in the region
estimated as Mdyn = RV 2
rot/G ∼ 8 × 108M⊙• for a radius R ∼ 150 pc and rotation velocity
Vrot ∼ 150 km s−1. This implies that the self-gravity of gas is not essential in the galactic
center, and a given-potential simulation would be sufficient to theoretically understand the
region.
13
The expanding molecular ring (or the parallelogram) was shown to share only 15
percent of the total gas mass within the central 1◦ region. This feature has been shown
to be extending vertically over ∼ 100 pc above and below the galactic plane (Sofue 1989).
For the very different b distribution, it is a clearly distinguished structure from the arms
and the ring described in this paper. On the (l, V ) plot, the feature can be fitted by an
ellipse of radius 1◦.2 (Bally et al 1987), slightly larger than the disk discussed in this paper
. There have been controversial interpretations about this feature: either it is due to some
explosive event (Scoville et al 1972; Kaifu et al 1972, 1974) or due to non-circular rotation
of disk gas (Burton and Liszt 1992; Binney 1991). We will discuss this feature based on
the present data in a separate paper.
Acknowledgement: The author would like to express his sincere thanks to Dr. John
Bally for making him available with the molecular line data in a machine-readable format.
References
Altenhoff, W. J., Downes, D., Pauls., T., Schraml, J. 1979, AAS 35, 23.
Arimoto, N., Sofue, Y., Tsujimoto, T. 1994, in preparation.
Bally, J., Stark, A.A., Wilson, R.W., and Henkel, C. 1987, ApJ Suppl 65, 13.
Bally, J., Stark, A.A., Wilson, R.W., and Henkel, C. 1988, ApJ 324, 223.
Binney, J.J., Gerhard, O.E., Stark, A.A., Bally, J., Uchida, K.I., 1991 MNRAS 252, 210.
Brown, R.L, and Liszt, H.S. 1984, ARAA 22, 223.
Burton, W. B. 1988, in Galactic and Extragalactic Radio Astronomy, ed. G. L. Verschuur
and K. I. Kellermann, 2nd edition (Springer-Verlag, New York) p 295.
Burton, W. B., and Liszt, H. S. 1983, in Surveys of the Southern Galaxy, ed. W. B. Burton
and F. P. Israel (Reidel Pub. CO, Dordrecht), p. 149.
Burton, W. B., and Liszt, H. S. 1992 AAS 95, 9.
Combes F 1992 ARAA, 29, 195.
Cox, P., Laureijs, R. 1989, in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris
(D.Reidel Publ. Co., Dordrecht) p. 121.
Dame, T. M., Ungerechts, H., Cohen, R. S., de Geus, E. J., Grenier, I. A., et al. 1987 ApJ
32, 706
Downes, D., Wilson, T. L., Beiging, J., Wink,J. 1980, AA Suppl. 40, 379.
Fujimoto, M. 1966, in Non-stable Phenomena in Galaxies, IAU Symp. No 29, ed. Arakeljan
(Academy of Sciences of Armenia, USSR), p.453.
Genzel, R., and Townes, C.H 1987, ARAA 25, 377.
Gusten, R. 1989, in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel
Publ. Co., Dordrecht) p. 89.
Handa, T., Sofue, Y., Nakai, N. Inoue, M., and Hirabayashi, H. 1987, PASJ 39, 709.
Hasegawa, T., Sato, F., Whiteoak, J. B., Miyawaki, R. 1993, ApJ 419, L77.
Heiligman, G. M. 1987 ApJ 314, 747.
Huntley, J. M., Sanders, R. H., and Roberts, W. W., 1978, ApJ 221, 521.
Ishiguro, M., Kawabe, R., Morita, K.-I., Okumura, S. K., Chikada, Y. et al. 1989, ApJ
14
344, 763.
Ishizuki, S. Kawabe, R., Ishiguro, M., Okumura, S. K., Morita, K-I., et al. 1990a Nature
344, 224.
Ishizuki, S., Kawabe, R., Ishiguro, M., Okumura, S. K., Morita, K. -I. et al. 1990b ApJ
355 436.
Kaifu, N., Iguchi, T., and Kato, T. 1974, PASJ 26, 117.
Kaifu, N., Kato, T., and Iguchi, T. 1972, Nature 238, 105.
Knapp, G. R., Stgark, A. A., Wilson, R. W. 1985 AJ 90, 254.
Liszt, H. S. 1988 in Galactic and Extragalactic Radio Astronomy, ed. G. L. Verschuur and
K. I. Kellermann, 2nd edition (Springer-Verlag, New York) p 359.
Liszt,H. S., Burton, W. B. 1978 ApJ 226, 790.
Liszt, H. S., and Burton, W. B. 1980 ApJ 236, 779.
Lo, K. Y., Berge, G. L., Claussen, M. J., et al. 1984, ApJ 282, L59.
Mulder, W.A., Liem, B.T., 1986, AA 157, 148
Nakai, N., Hayashi, M., Handa, T., Sofue, Y., Hasegawa, T., and Sasaki, M., 1987, PASJ
39, 685.
Okuda, H., Shibai, H., Nakagawa, T., Matsuhara, T., Maihara, T., et al. 1989 in The
Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht)
p. 145.
Oort, J. H., Kerr, F. J., Westerhout, G. 1958, MNRAS 118, 379.
Oort, J. H. 1977, ARAA 15, 295
AA Suppl 58, 197.
Roberts, W. W. 1969, ApJ 158, 123.
Roberts, W. W., Huntley, J. M., van Albada, G. D. 1979, ApJ, 233, 67.
Shaver, P. A., McGee, R. X., Newton, L. M., Danks, A. C., Pottasch, S. R. 1983, MNRAS,
204, 53.
Scoville, N.Z. 1972, ApJ 175, L127.
Scoville, N.Z., Solomon, P. M., and Jefferts, K. B. 1974 ApJ 187, L63.
Sofue, Y. 1985 PASJ 37, 697
Sofue, Y. 1989, in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel
Publ. Co., Dordrecht) p. 213.
Sofue, Y. 1990 PASJ 42, 827
Sofue, Y. 1991 PASJ 43, 671
Sofue, Y., and Handa, T. 1984, Nature 310, 568.
Sofue, Y., Nakai, N. 1993 PASJ 45, 139.
Sofue, Y., Nakai, N. 1994 PASJ 46, 147.
Sofue, Y., and Reich, W. 1979, AA Suppl 38, 251.
Solomon, P.M., Scoville, N.Z., and Sanders, D.B., 1979, ApJ 232, L89.
Sφrensen, S. -A., Matsuda, T., and Fujimoto, M. 1976, A. Sp. Sci. 43, 491.
Stark, A. A., Bally, J., Wilson, R. W., Pound, M. W., 1989, in The Center of the Galaxy
(IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht) p. 213.
Tsuboi, M. 1989 in The Galactic Center (IAU Symp. 136), ed. M.Morris (D.Reidel Publ.
Co., Dordrecht) p. 135
Wada, K., Habe, A., Taniguchi, Y., Hasegawa, T. 1994, submitted to Nature
15
Wada, K., Habe, A. 1992 MNRAS 258, 82
16
Table 1: Galactic Center Arms and Ring.
Parameters
Ring (I+II) Arm I
Arm II
Arm III
Arm IV
From (l, V )
(◦, km s−1) . (+0.9,90)
(0.9, 80)
(0.1, 60)
∼ (1, 100)
(0, 140)
(0.1, 60)
To (l, V )
(◦, km s−1) . (−0.65, −140)(−0.7, −150) (−0.6, −80)
(−0.15, 10)
(0, −20)
VLSRat l = 0◦ (km s−1) . . . .
....
∼ (−1, −200)
−40
+50
+70
−50
(◦,◦) . . . . . . . . . (+0.9, 0.0)
From (l, b)
(◦,◦) . . . . . . . . (−0.65, −0.08)(−1.0, −0.2) (−0.65, −0.17)(0, 0)
To (l, b)
(◦) . . . . . . . . . .
b at l = 0◦
(0, 0)
(◦/pc) . . . . . .
Length
0.35/52
Min. b width (◦/pc) . . . . . . .
Max. b width (◦/pc) . . . . . . .
−0.067
0.9/133
0.091/13.5
0.2/30
0.050
1.9/280
0.088/13
0.33/50
(+0.9, −0.1) (0.25, −0.05) (0.25, 0.25)
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
....
(◦/pc) . . . . . . . 1.55/230
(◦/pc) . . . . . . . 0.132/19.5
(◦) . . . . . . . . . . 85◦.1
Maj.ax.len.
Min.ax.len.
Inclination
Ring cen. (l, b) (◦,◦) . . . . . . .(0.12,0.0)
Ring radius
Rot. Velo
(pc) . . . . . . . . 120
(km s−1) . . . . +90/ − 140
....
....
....
....
....
....
....
....
....
....
Mol. Mass†
Remarks
(107M⊙•) . . . . . 3.07
. . . . . . . . . . . . . Circum Nuc. asso. Sgr B Sgr C
1.72
1.35
∗ The distance to the galactic center is assumed to be 8.5 kpc.
† 1.61 times the H2 mass obtained from the 13CO intensity to H2 conversion [see eq. (1)-(3)],
where the metal abundance has been assumed to be twice the solar. This also applies to mass in
Table 2. The statistical error which occurs during intensity integration is only a few %, while the
error arising from ambiguity of the conversion factor is about 20 to 30% (Arimoto et al 1994).
Sgr A?
Sgr A?
Figure Captions
Fig. 1: (a) The (l, V ) diagram of the 13CO (J = 1 − 0) line emission of the central region of
the Milky Way by averaging the data from −0.35 ≤ b ≤ 0◦.17 as obtained with the Bell Telephone
7-m telescope by Bally et al (1987). Contours are in unit of K T ∗
A at levels 0.1×(1, 2, 3, 4, 5, 6, 8,
12, 15, 20, 25).
(b) The same as Fig. 1a, but the local and foreground CO emissions have been subtracted by
applying the "pressing method" (see the text for the procedure). Contour levels are same as in
(a).
Fig. 2: The (l, V ) diagrams averaged in 4′ b interval. Local/foreground emissions have been
removed. 'Rigid-rotation' ridges (arms) are dominant in the disk at b <∼ 10′ (25 pc). Contours
are in unit of K T ∗
A at levels 0.2×(1, 2, 3,..., 9, 10, 12, 14, 16, 18, 20, 25, 30).
Fig. 3: Schematic sketch of the major ridges (arms) in the (l, V ) diagrams.
Fig. 4: (a) Integrated intensity map in the whole velocity range at −250 ≤ VLSR ≤ 250 km
s−1. This is almost the same as the map presented by Stark et al (1989), except that the local
contribution has been subtracted. Contours are in unit of K km s−1 at levels 25×(1, 2, 3, ..., 9,
10, 12, 14, 16, 18, 20, 25, 30).
(b) Same but in a grey-scale representation. For intensity scale, see (a). The bottom figure
shows the same, but the scale in the latitude direction has been doubled. Galactic Center Arm
(GCA) I runs as a long arc in the positive b side; GCA II runs in the negative b side.
Fig. 5: (l, V ) diagrams corresponding to (a) Galactic Center Arms I and (b) II, which were
used to obtain intensity maps of the Galactic Center Arms in Fig. 6. Contours are in unit of K
T ∗
A at levels 0.2×(1, 2, 3, 4, 5, 6, 8, 10, 12, 15, 20, 25, 30, 35).
Fig. 6: Integrated intensity maps corresponding to (a) Galactic Center Arm I, and (b) Arm
II as in Fig. 5. Contours are in unit of K km s−1 at levels 12.5×(1, 2, 3, ..., 9, 10, 12, 14, 16, 18,
20, 25, 30). (c) Arms I+II. Contours are in unit of K km s−1 at levels 25×(as above).
Fig. 7: Intensity variation across Galactic Center Arms I and II perpendicular to the galactic
plane averaged at l = 0◦.24 to −0◦.33, where the arms are most clearly separated.
Fig. 8: (a) A velocity field as obtained by taking the first moment of the (VLSR, l, b) cube
(intensity-weighted mean velocity field). Contour interval is 10 km s−1 Full-line contours are for
positive velocity starting at 0 km s−1. Dashed contours are for negative velocity.
(b) Same as (a), but for the "disk component" with VLSR < 100 km s−1.
Fig. 9: Two-armed spiral model with a spiral infalling motion. Gas density distribution is
shown by spiral-like contours as projected on the galactic plane. Calculated (l, V ) diagram is shown
by tilted X shaped contours. The scales are arbitrary.
(a) Two spiral arms with a pitch angle p = 10◦ are assumed. The azimuthally averaged gas
density has a hole at the center, corresponding to a ring distribution of gas on which two arms
are superposed.
In addition to a constant circular rotation, radial infall of velocity Vrotsin p is
superposed.
(b), (c) The same as (a), but the spiral arms are oval in shape whose major axis are inclined
by ±30◦ from the nodal line.
(d)-(f) The same as (a)-(c), respectively, but the density distribution along the arms has the
maximum at the center and the pitch angle is taken larger: p = 20◦.
Fig. 10: Possible deconvolution of the (l, V ) diagrams for Galactic Center Arms I and II into
a spatial distribution as projected on the galactic plane. Contour interval is 0.25 starting at 0.1 in
an arbitrary unit. Sgr A is assumed to be at the center.
Fig. 11: Superposition of the radio continuum emission at 10 GHz (contours: Handa et al
1987) on (a) 13CO, and (b) CS emission maps (grey scale). Contours are in unit of K TB of 10
GHz continuum brightness at levels 0.1×(1, 2, 3, 4, 6, 8, 10, 15, 20, 25). For CO intensity scale,
see Fig. 4.
|
astro-ph/0511821 | 2 | 0511 | 2005-12-06T08:38:00 | Dark Energy Evolution and the Curvature of the Universe from Recent Observations | [
"astro-ph",
"hep-ph"
] | We discuss the constraints on the time-varying equation of state for dark energy and the curvature of the universe using observations of type Ia supernovae from Riess et al. and the most recent Supernova Legacy Survey (SNLS), the baryon acoustic oscillation peak detected in the SDSS luminous red galaxy survey and cosmic microwave background. Due to the degeneracy among the parameters which describe the time dependence of the equation of state and the curvature of the universe, the constraints on them can be weakened when we try to constrain them simultaneously, in particular when we use a single observational data. However, we show that we can obtain relatively severe constraints when we use all data sets from observations above even if we consider the time-varying equation of state and do not assume a flat universe. We also found that the combined data set favors a flat universe even if we consider the time variation of dark energy equation of state. | astro-ph | astro-ph | astro-ph/0511821
November 2005
5
0
0
2
c
e
D
6
2
v
1
2
8
1
1
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Dark Energy Evolution and the Curvature
of the Universe from Recent Observations
Kazuhide Ichikawa and Tomo Takahashi
Institute for Cosmic Ray Research,
University of Tokyo, Kashiwa 277-8582, Japan
Abstract
We discuss the constraints on the time-varying equation of state for dark energy
and the curvature of the universe using observations of type Ia supernovae from
Riess et al. and the most recent Supernova Legacy Survey (SNLS), the baryon
acoustic oscillation peak detected in the SDSS luminous red galaxy survey and cosmic
microwave background. Due to the degeneracy among the parameters which describe
the time dependence of the equation of state and the curvature of the universe, the
constraints on them can be weakened when we try to constrain them simultaneously,
in particular when we use a single observational data. However, we show that we
can obtain relatively severe constraints when we use all data sets from observations
above even if we consider the time-varying equation of state and do not assume a
flat universe. We also found that the combined data set favors a flat universe even
if we consider the time variation of dark energy equation of state.
1
Introduction
Almost all current cosmological observations indicate that the present universe is accel-
erating. It can be explained by assuming that the universe is dominated by dark energy
today. Although many candidates for dark energy have been proposed so far, we still do
not know the nature yet. Many studies have been devoted to investigate dark energy as-
suming or constructing a specific model and then study its consequences on cosmological
observations such as cosmic microwave background (CMB), large scale structure (LSS),
type Ia supernovae (SNeIa) and so on. On the other hand, many efforts have also been
made to study dark energy in phenomenological way, i.e., as model independent as pos-
sible. In such approaches, dark energy can be parameterized with its equation of state
and constraints on it can be obtained using cosmological observations. Assuming that the
equation of state for dark energy wX is constant in time, current observations give the
constraint as wX ∼ −1 [1, 2, 3, 4, 5, 6]. Although one of the most famous models for dark
energy is the cosmological constant whose equation of state is wX = −1, most models
proposed so far have time-varying equation of state. Thus, when we study dark energy,
we should accommodate such time dependence in some way.
Many recent works on dark energy investigate the time dependence of the dark energy
equation of state using simple parameterizations such as wX = w0 + w1(1−a) with a being
the scale factor of the universe [7, 8]. Parameterizing the dark energy equation of state in
simple ways, the constraints on the time evolution of wX have been considered (for recent
works on this issue, for example, see Ref. [9]). It should be mentioned that, when one
studies the equation of state for dark energy, it is usually assumed that the universe is
flat. It should be also noted that dark energy is usually assumed to be the cosmological
constant when one derives the constraint on the curvature of the universe. However, it has
been discussed that, assuming a non-flat universe, the constraints on wX and the curvature
of the universe can be relaxed to some extent even with the time independent wX from the
CMB data alone [10]. Also, even if we assume a flat universe, there are degeneracies among
the parameters which describe the evolution of dark energy, i.e., the time dependence of
wX when we consider the constraints on dark energy [11, 12]. Furthermore it has been also
discussed that if we remove the prior of a flat universe, the degeneracies becomes much
worse [13]. Since it is very important to study the time dependence of wX to differentiate
the models of dark energy and also the curvature of the universe to test the inflationary
paradigm, we should investigate how the prior on the curvature of the universe affects the
determination of the time-varying equation of state for dark energy and vice versa. Some
works along this line have been done using a specific one-parameter parameterization for
the time-varying wX [14].
In this paper, we study this issue, namely the determination of the evolution of dark
energy and the curvature of the universe, in some detail using widely used parameterization
of the equation of state for dark energy. We consider the constraints from observations
of SNeIa reported in Refs. [4, 6], the baryon acoustic oscillation detected in the SDSS
luminous red galaxy survey [15] and recent CMB observations including WMAP [1]. In the
1
next section, we summarize the analysis method we adopt in this paper. Then we discuss
the constraints from above mentioned observations on the curvature of the universe with
time-varying equation of state for dark energy followed by the analysis on wX without
assuming a flat universe. The final section is devoted to the summary of the paper.
2 Method
In this section, we briefly summarize the method for constraining the parameters which
describe the dark energy evolutions and other cosmological parameters. To study the evo-
lution of dark energy, we use the following parameterization for the time-varying equation
of state [7, 8]
z
wX(z) = w0 +
1 + z
w1 = w0 + (1 − a)w1,
(1)
where z is the redshift. In this parameterization, the equation of state at the present time
is wX(z = 0) = w0 and for the early time it becomes wX(z = ∞) = w0 + w1. Since we are
interested in the late-time acceleration of the universe due to dark energy, we consider the
case where the dark energy dominates the universe only at late time. Thus, in this paper,
we assume
w0 + w1 < 0,
(2)
in order not to include the possibilities of early-time dark energy domination. With this
parameterization, the energy density of dark energy can be written as
1 + z (cid:19) ,
ρX(z) = ρX0(1 + z)3(1+w0+w1) exp(cid:18)−3w1z
(3)
where ρX0 is the energy density of dark energy at present time. The Hubble parameter is
given by
H 2(z) = H 2
0(cid:20)Ωr(1 + z)4 + Ωm(1 + z)3 + Ωk(1 + z)2 + ΩX (1 + z)3(1+w0+w1) exp(cid:18)−3w1z
1 + z (cid:19)(cid:21) ,
(4)
where H0 is the Hubble parameter at the present epoch, Ωi is the energy density of a
component i normalized by the critical energy density and the subscripts r, m, k and X
represent radiation, matter, the curvature of the universe and dark energy, respectively.
To consider the constraints on dark energy and other cosmological parameters, we use the
data from SNeIa, the baryon acoustic oscillation peak and the CMB.
As for SNeIa data, we use the gold data set given in Ref. [4] and the first year data of
the Supernova Legacy Survey (SNLS) released recently [6]. Constraints from SNeIa can
be obtained by fitting the distance modulus which is defined as
M − m = 5 log(cid:18) dL
Mpc(cid:19) + 25.
2
(5)
Here dL is the luminosity distance which is written as
dL =
1 + z
pΩkS(cid:18)pΩkZ z
0
dz ′
H(z ′)/H0(cid:19) ,
(6)
where S is defined as S(x) = sin(x) for a closed universe, S(x) = sinh(x) for an open
universe and S(x) = x with the factor pΩk being removed for a flat universe.
We also use the baryon acoustic oscillation peak detected in the SDSS luminous red
galaxy survey [15]. To obtain the constraint, we make use of the parameter A which is
defined as
A =
√Ωm
(H(z1)/H0)1/3 "
1
z1pΩkS(cid:18)pΩkZ z1
0
dz ′
H(z ′)/H0(cid:19)#2/3
,
(7)
where z1 = 0.35 and A is measured to be A = 0.469 ± 0.017 [15].
of the CMB angular power spectrum [16]. R is given by
For the CMB data, we only use the shift parameter R which determines the whole shift
R =
√Ωm
pΩkS(cid:18)pΩkZ z2
0
dz ′
H(z ′)/H0(cid:19) ,
(8)
where z2 = 1089. It has been discussed that using the CMB shift parameter is a robust way
to include the constraints from observations of CMB [17]. From the recent observations
of CMB including WMAP, CBI and ACBAR, the shift parameter is constrained to be
R = 1.716 ± 0.062 [1, 17]
In this paper, we only consider the effect of the modification of the background evolu-
tion by the change in the cosmological parameters. In fact, the properties of dark energy
can also modify the evolutions of cosmic density fluctuation. When we consider the effects
of dark energy perturbation, we also have to specify the speed of sound of dark energy
component. However, such modification can arise at low multipole region of the CMB
power spectrum where the errors due to cosmic variance are large. Thus the constraints
from CMB on dark energy mostly come from the position of acoustic peaks which can
be described by the shift parameter. Hence we do not consider the perturbation of dark
energy in this paper#1 .
3 Constraints on the Ωm vs. ΩX plane
Now we discuss the implication of the dark energy evolution on the determination of the
curvature of the universe. For this purpose, we derive the constraints from the observations
on the Ωm vs. ΩX plane. First we consider the case with the constant equation of state for
#1However, the speed of sound can be useful to differentiate the models of dark energy. There are some
works which discuss the constraint on the speed of sound. For interested reader, we refer Refs. [18, 19, 20].
3
Figure 1: Constraints on Ωm and ΩX from SNeIa (a), the baryon acoustic oscillation peak
(b), the CMB shift parameter (c) and all data combined (d). For the constraint from
SNeIa data and all data combined, we show the contours obtained from the gold set of
Ref. [4] and SNLS [6] separately. Contours of 1σ (red solid line) and 2σ (green dashed line)
are shown (for SNLS data, 1σ and 2σ contours are shown in blue dash-dotted line and
purple dotted line, respectively). We assumed the cosmological constant as dark energy
and a flat universe here.
dark energy. In Fig. 1, we show the contours of 1σ and 2σ constraints from observations of
SNeIa (a), the baryon acoustic oscillation peak (b) and the CMB shift parameter (c). We
also show the constraint from the combination of all data sets (d). In Fig. 1, we assume
the equation of state as wX = −1. For the constraints from SNeIa, we used the gold data
set from Ref. [4] and the data from SNLS [6] separately. As we can see, although each
constraint from a single observational data has the degeneracy in the Ωm vs. ΩX plane,
when all the data sets are combined, we can obtain a severe constraint as Ωm ∼ 0.3 with
Ωk ∼ 0, which is a well-known result.
Next we consider the case where the equation of state for dark energy is allowed to
vary, but still we keep wX constant in time. In Fig. 2, we show the contours of 1σ and 2σ
constraints after marginalizing over the values of wX . Here we assumed the prior on wX
as −5 ≤ wX ≤ 0. As seen from the figure, the allowed regions become larger compared
to those for the case with wX = −1 when we use a single data set alone. However, by
combining all data, we can obtain almost the same constraint as the case with wX = −1
4
Figure 2: The same as Fig. 1 except that we marginalized over the value of wX. Here we
considered the constant wX.
even though we marginalize over wX. This can be understood by noting that the values
of wX which give the minimum χ2 for fixed (Ωm, ΩX) from each data set are different. To
see this, we also plot contours of constant wX which gives minimum values of χ2 at each
point on the Ωm vs. ΩX plane in Fig. 3. We can clearly see that the favored values of wX
vary for different data sets. Observations we consider here measure some distance scales
to certain redshifts which are determined by the energy density of matter Ωm, dark energy
ΩX and the equation of state wX. If we vary the value of wX, the density parameters Ωm
and ΩX can have more freedom to be consistent with observations since the fit to the data
depends on the combinations of these quantities. For a single observation, when wX is
allowed to vary, larger range of (Ωm, ΩX) can be consistent with observations. However,
when we use all observations, the combinations of Ωm, ΩX and wX which are consistent
with observations become fairly limited. Thus while the allowed regions become larger
for each data set, when we combine all the data, the allowed region converges towards
the concordance model with Ωm + ΩX ∼ 1 and Ωm ∼ 0.3 even if we do not assume the
cosmological constant as dark energy.
Next we discuss the case with the time dependent equation of state parameterized as
Eq. (1). As it has been already pointed out in the literature, even if we assume a flat
universe, there exists a degeneracy among the parameters which describe the evolution of
equation of state for dark energy using a single observational data. It is also known that
5
Figure 3: Contours of constant wX which gives minimum χ2 when we marginalize over
wX for the constraints from SNeIa (a), the baryon acoustic oscillation peak (b), the CMB
shift parameter (c) and all data combined (d). Contours of wX = −2 (red solid line), −1
(green dashed line) and −0.8 (blue dotted line) are shown except for the case where all
three data are combined (panel (d)). For the panel (d), contours of wX = −1.1 (red solid
line), −1 (green dashed line) and −0.9 (blue dotted line) are shown. For SNeIa, the data
from SNLS are used here.
the degeneracy can be removed using more than one observation assuming the flatness.
Here we discuss to what extent the evolution of dark energy equation of state can affect
the determination of Ωm and ΩX when we do not assume a flat universe. In Fig. 4, the
contours of 1σ and 2σ constraints are shown as the same manner as Fig. 1 except that
we varied the both values of w0 and w1 which appear in Eq. (1) and marginalized over
them to obtain the constraint. We assumed the prior on them as −5 ≤ w0 ≤ 0 and
−4 ≤ w1 ≤ 4 under the condition of Eq. (2). For the constraint from SNeIa, the allowed
regions get larger compared to the case with the constant equation of state discussed
above (see Fig. 2). As for the constraint from the baryon acoustic oscillation peak and the
CMB shift parameter, the allowed regions are almost the same as the case with a constant
equation of state. For observations which measure a single distance scale, the fit to the
data does not significantly become better, even if the time evolution of wX is allowed,
due to the degeneracy between w0 and w1. Furthermore, when we use all three different
observations together, the allowed region becomes almost the same as that of the case
6
Figure 4: The same as Fig. 1 except that we consider the time dependent equation of state
for dark energy and marginalize over w0 and w1.
with a constant equation of state. This is because each observation favors different values
of wX. Even if we consider the time-varying equation of state, different combinations of
w0 and w1 are chosen to minimize the value of χ2 for each observation. Thus the allowed
region from the combined data set is almost unchanged although the constraints from a
single observation, in particular from SNeIa, can become weaker.
Here we comment on the effect of the prior on w0 and w1. Although the constraints
from a single data alone somewhat depend on the prior, the allowed region for all data
combined is almost unaffected since the combinations of w0 and w1 favored around the
allowed region from all data sets are far from the the edge of the prior we assumed.
4 Constraints on the Ωm vs. wX plane
In this section, we consider the constraints on the Ωm vs. wX plane. First we show the
constraint on Ωm and w0 in a flat universe with the equation of state for dark energy being
constant. In Fig. 5, contours of 1σ and 2σ constraints from observations of SNeIa (a), the
baryon acoustic oscillation peak (b), the CMB shift parameter (c) and all data combined
(d) are shown. As is well-known, SNeIa and CMB are complementary for constraining
dark energy, which can be seen from the figure. In addition, we can see that the constraint
7
Figure 5: Constraints on Ωm and wX from observations of SNeIa (a), the baryon acoustic
oscillation peak (b), CMB (c) and all data combined (d). We assumed a flat universe and
a constant equation of state here.
from the baryon acoustic oscillation peak is also complementary. Thus we can obtain a
severe constraint using all three data sets.
Next we consider the case with the time-varying equation of state for dark energy.
Here we discuss the case with a flat universe. In Fig. 6, the constraints on Ωm and w0
are shown marginalizing over w1. When we marginalize the values of w1, we assumed the
prior on w1 as −4 ≤ w1 ≤ 4. Similarly to the situation where we constrain Ωm and ΩX
discussed in the previous section, if we consider the time-varying equation of state for dark
energy, the allowed region becomes larger when we use a single data set. However, when
we consider all three data sets, we can obtain a relatively severe constraint. Although
the allowed region of w0 extends to larger values, Ωm is constrained to be Ωm ∼ 0.3
which is the almost same as the case with assuming a constant equation of state. This
is because the values of w1 favored by different data sets given Ωm and w0 are different.
Thus the degeneracy among the parameters describing the time dependence of the dark
energy equation of state can be removed drastically by using all three data sets.
Now we discuss the constraints without assuming a flat universe. First we discuss the
case with wX being constant in time. In Fig. 7, we show the constraints from SNeIa (a),
the baryon acoustic oscillation (b) and all data combined (c). For the last case, the data
from the CMB shift parameter is included. To obtain the constraint, we marginalized
8
Figure 6: The same as Fig. 5 except that we consider the time dependent equation of state
for dark energy as in Eq. (1). To obtain the constraint, we marginalized over w1. Here we
assumed a flat universe.
over Ωk with the prior −0.3 ≤ Ωk ≤ 0.3. Here we do not show the constraint from the
CMB shift parameter alone since we cannot obtain a significant constraint from it in the
parameter region we consider in Fig. 7. This is because, as it has already been pointed
out, the curvature of the universe and wX are strongly degenerate in the CMB power
spectrum [10]. Again, although the constraints from a single data set alone are weakened,
if we consider all data sets, we can obtain almost the same constraint as that with the
case where a flat universe is assumed. Notice that the constraint on w0 is also not changed
much compared to that with flat universe prior. Hence from Fig. 7, we can say that the
prior on the curvature does not affect the determination of the equation of state for dark
energy much.
We also show the case where we consider the time-varying equation of state without
assuming a flat universe. In Fig. 8, the constraints are shown as the same as Fig. 7 except
that we marginalized over Ωk and w1 in this case. The prior on w1 is taken as −4 ≤ w1 ≤ 4.
The CMB shift parameter cannot give a significant constraint in the parameter range we
consider in this case too. As in the previous cases, using each observational data alone,
the allowed region becomes significantly larger compared to those with constant equation
of state and a flat universe. However, when we use all data sets, the allowed region does
not change much. From Figs. 5 and 7, we can conclude that the prior on the curvature
9
Figure 7: Constraints on Ωm and wX from observations of SNeIa (a), the baryon oscillation
(b) and all data combined (c). Here we do not assume a flat universe, but a constant
equation of state for dark energy is assumed. In this figure, we marginalized over Ωk.
of the universe does not affect the constraint on Ωm and w0 much if we use all data sets.
In particular, we can obtain the constraint such that Ωm ∼ 0.3 without assuming a flat
universe. Moreover, the assumption of the constancy of the equation of state for dark
energy also does not affect the constraint on Ωm. On the constraint on the equation of
state, if we consider the time-varying equation of state, the allowed region becomes slightly
larger even if all data combined is used.
5 Constraints on the evolution of dark energy
In this section, we consider the constraints on the time dependence of dark energy equation
of state, i.e., on w0 and w1 without assuming the prior of a flat universe. However first
we discuss the case with a flat universe for the later comparison with the case where a
flat universe is not assumed. In Fig. 9, contours of 1σ and 2σ allowed regions are shown
for the case with a flat universe. Here we fix the energy density of matter as Ωm = 0.28.
Notice that we do not consider the region where w0 + w1 > 0 in order not to include
the possibilities of early-time dark energy domination. As we can see from the figure, the
constraint from SNeIa is stringent compared to baryon acoustic oscillation peak and CMB.
Since the position of the baryon acoustic oscillation peak measures a single distance scale
from z = 0 to z = 0.35 and the CMB shift parameter also gives a single scale from z = 0 to
z = 1089, there is strong degeneracy among the parameters which describe the equation
of state for dark energy. Notice that the distance scales are determined by the integration
of the inverse of the Hubble parameter, which can smear out the information on the time
dependence of the equation of state. This is the reason why the strong degeneracy exists
when only a single scale is considered. However, as for SNeIa data, the degeneracy is
10
Figure 8: The same as Fig. 7 except that we consider the time dependent equation of state
for dark energy and do not assume a flat universe. In this figure, we marginalized over Ωk
and w1.
removed to some extent since we have the distances from z = 0 to various redshifts#2 .
Thus observations of SNeIa can mainly constrain the time dependence of the equation of
state for dark energy.
Next we show the constraints on w0 and w1 in a flat universe without assuming a
particular value for Ωm. To obtain the constraint, we marginalized over Ωm with the prior
0 ≤ Ωm ≤ 0.5. In Fig. 10, we show the constraints on w0 and w1 from SNeIa (a) and all
data combined (b). We do not show the constraints from the baryon acoustic oscillation
peak and the CMB shift parameter because we cannot obtain significant constraints from
them in the parameter region of w0 and w1 we consider here. Also in this case, such severe
degeneracies among the parameters exist because those observations measure a single
distance scale. As in the previous case, when we include all data sets in the analysis,
we can obtain almost the same constraint as that with Ωm being fixed to 0.28. This is
partly because the best fit value of Ωm is near 0.28 when we marginalized over it, but
also because the values of Ωm minimizing χ2 for each observational data set are different.
To see this, in Fig. 11, we show contours of constant Ωm which minimize χ2 when we
marginalize over Ωm for the constraints from SNeIa (a), the baryon acoustic oscillation
peak (b), the CMB shift parameter (c) and all data combined (d). As is clear from the
figure, preferred values of Ωm for each observation are quite different. Thus when all data
are used, we can obtain a severe constraint although a single observation cannot constrain
the equation of state for dark energy. It should be also mentioned that the prior on Ωm
does not affect the constraint on the time dependence of the equation of state when we
use all data combined.
Now we discuss the case with Ωm not being fixed and not assuming a flat universe. In
Fig. 12, the allowed regions are shown after marginalizing over Ωm and Ωk. We assumed
#2Of course, there is a limitation to determine the time dependence of the equation of state using SNeIa
data [11, 12].
11
Figure 9: Constraints on w0 and w1 from SNeIa (a), the baryon acoustic oscillation peak
(b), the CMB shift parameter (c) and all data combined (d). We assumed Ωm = 0.28 with
a flat universe.
the prior on these variables as 0 ≤ Ωm ≤ 0.5 and −0.3 ≤ Ωk ≤ 0.3. In this case, the
constraint is significantly weakened when we consider a single data set alone. We only
report here the constraint from SNeIa (a) and that from all data combined (b) since the
baryon acoustic oscillation peak and the CMB shift parameter cannot give meaningful
constraints on this plane in this case too. As we can see, when we consider the constraint
from all data sets, it significantly becomes severe compared to that from SNeIa alone as
seen from the figure. This is because the favored values of Ωm and Ωk from each data
set are different as in the previous cases. Thus we can conclude that the prior on the
curvature does not affect much the determination of the equation of state for dark energy
if we use all data combined.
6 Summary
We considered the constraint on the curvature of the universe and the equation of state for
dark energy from observations of SNeIa, the baryon acoustic oscillation peak and the CMB
shift parameter. Usually, when one discusses the curvature of the universe, dark energy is
assumed to be the cosmological constant. Moreover, when one considers the constraint on
12
Figure 10: Constraints on w0 and w1 from observations of SNeIa (a) and all data combined
(b). A flat universe is assumed but we marginalized over Ωm.
the evolution of dark energy, in particular the time dependence of the dark energy equation
of state, a flat universe is usually assumed. In this paper, we discussed the constraints
on the curvature of the universe without assuming the cosmological constant and also the
time dependence of the equation of state for dark energy without assuming a flat universe.
We showed that the constraint on the curvature of the universe is significantly relaxed
from a single observation when we allow the time dependence in the dark energy equation
of state. However, it was also shown that, when we use all data sets, the curvature of the
universe or the energy density of matter and dark energy are severely constrained to be
Ωm ∼ 0.3 with a flat universe even if we consider the time-varying equation of state for dark
energy. Observations we consider here measure some distance scales to certain redshifts
which can be determined by the energy density of matter Ωm, dark energy ΩX and the
equation of state wX. If we assume a broad range for wX, the energy density of matter Ωm
and dark energy ΩX can have more freedom to be consistent with observations since the
fit to the data depends on the combinations of these variables. For a single observation,
if wX is allowed to vary more freely, much larger range of values of Ωm and ΩX can be
consistent with observations. However, when we use all observations, the combinations
of Ωm, ΩX and wX consistent with observations become fairly limited. Thus even if the
allowed regions become large for each data set, when we combine all the data, we can
obtain a severe constraint, which is interestingly almost the same region as that we obtain
assuming the cosmological constant as dark energy.
We also investigated the constraint on the time-varying equation of state for dark
energy without assuming a flat universe. Similarly to the situation where we constrained
the curvature with the time-varying equation of state, the allowed region for wX becomes
larger when we use a single observational data. However, if we use all data sets considered
in this paper, we can obtain almost the same constraint as that in the case where a flat
universe is assumed.
Finally, we summarize what we found in this paper. The combination of the current
13
Figure 11: Contours of constant Ωm which gives minimum χ2 when we marginalize over
Ωm for the constraints from SNeIa (a), the baryon acoustic oscillation peak (b), CMB shift
parameter (c) and all data combined (d). Contours of Ωm = 0.3 (red solid line), 0.2 (green
dashed line) and 0.1 (blue dotted line) are shown except for the case where all three data
are combined (panel (d)). For the panel (d), Ωm = 0.29 (red solid line), 0.28 (green dashed
line) and 0.27 (blue dotted line) are shown. For SNeIa, we use the data from SNLS.
observations
• favors a flat universe regardless of the prior on the equation of state for dark energy
with or without time evolution.
• favors Ωm ∼ 0.3 regardless of the flatness prior and the prior on the equation of state
for dark energy with or without time evolution.
• yields constraints on the time evolution of dark energy equation of state regardless
of the prior on Ωm and Ωk.
In future observations, we can obtain much more stringent constraints on the equation
of state for dark energy as well as the curvature of the universe regardless of the prior on
other cosmological parameters. Hence we can expect that we will be able to have much
insight on the nature of dark energy and also the inflationary paradigm.
14
Figure 12: The same as Fig. 10 except that we marginalized over Ωm and Ωk.
Acknowledgment: T.T would like to thank the Japan Society for Promotion of Science
for financial support.
References
[1] D. N. Spergel et al., Astrophys. J. Suppl. 148, 175 (2003) [arXiv:astro-ph/0302209].
[2] M. Tegmark et al.
[SDSS Collaboration], Phys. Rev. D 69, 103501 (2004)
[arXiv:astro-ph/0310723].
[3] J. L. Tonry et al. [Supernova Search Team Collaboration], Astrophys. J. 594, 1
(2003) [arXiv:astro-ph/0305008].
[4] A. G. Riess et al. [Supernova Search Team Collaboration], Astrophys. J. 607, 665
(2004) [arXiv:astro-ph/0402512].
[5] C. J. MacTavish et al., arXiv:astro-ph/0507503.
[6] P. Astier et al., arXiv:astro-ph/0510447.
[7] M. Chevallier and D. Polarski,
Int. J. Mod. Phys. D 10,
213 (2001)
[arXiv:gr-qc/0009008].
[8] E. V. Linder, Phys. Rev. Lett. 90, 091301 (2003) [arXiv:astro-ph/0208512].
[9] Y. Wang
Phys. Rev. Lett.
92
and M. Tegmark,
241302
[arXiv:astro-ph/0403292]; B. Feng, X. L. Wang and X. M. Zhang, Phys. Lett. B 607,
35 (2005) [arXiv:astro-ph/0404224]; H. K. Jassal, J. S. Bagla and T. Padmanabhan,
Mon. Not. Roy. Astron. Soc. 356, L11 (2005) [arXiv:astro-ph/0404378]; D. A. Dicus
and W. W. Repko, Phys. Rev. D 70, 083527 (2004) [arXiv:astro-ph/0407094];
(2004)
15
S. Hannestad and E. Mortsell, JCAP 0409, 001 (2004) [arXiv:astro-ph/0407259];
U. Seljak et al., Phys. Rev. D 71, 103515 (2005) [arXiv:astro-ph/0407372];
F. Giovi, C. Baccigalupi and F. Perrotta, Phys. Rev. D 71, 103009 (2005)
[arXiv:astro-ph/0411702]; A. Upadhye, M. Ishak and P. J. Steinhardt, Phys. Rev.
D 72, 063501 (2005) [arXiv:astro-ph/0411803]; H. K. Jassal, J. S. Bagla and
T. Padmanabhan, Phys. Rev. D 72, 103503 (2005) [arXiv:astro-ph/0506748];
M. Doran, K. Karwan and C. Wetterich, arXiv:astro-ph/0508132; S. Nesseris and
L. Perivolaropoulos, arXiv:astro-ph/0511040; J. Q. Xia, G. B. Zhao, B. Feng, H. Li
and X. Zhang, arXiv:astro-ph/0511625.
[10] J. L. Crooks, J. O. Dunn, P. H. Frampton, H. R. Norton and T. Takahashi, Astropart.
Phys. 20, 361 (2003) [arXiv:astro-ph/0305495].
[11] I. Maor, R. Brustein and P. J. Steinhardt, Phys. Rev. Lett. 86, 6 (2001) [Erratum-
ibid. 87, 049901 (2001)] [arXiv:astro-ph/0007297].
[12] I. Maor, R. Brustein, J. McMahon and P. J. Steinhardt, Phys. Rev. D 65, 123003
(2002) [arXiv:astro-ph/0112526].
[13] Y. Wang and M. Tegmark, in Ref. [9].
[14] Y.
g. Gong
and Y. Z. Zhang, Phys. Rev. D 72,
043518
(2005)
[arXiv:astro-ph/0502262].
[15] D. J. Eisenstein et al., arXiv:astro-ph/0501171.
[16] J. R. Bond, G. Efstathiou and M. Tegmark, Mon. Not. Roy. Astron. Soc. 291, L33
(1997) [arXiv:astro-ph/9702100].
[17] Y. Wang and P. Mukherjee, Astrophys. J. 606, 654 (2004) [arXiv:astro-ph/0312192].
[18] R. Bean and O. Dore, Phys. Rev. D 69, 083503 (2004) [arXiv:astro-ph/0307100].
[19] J. Weller and A. M. Lewis, Mon. Not. Roy. Astron. Soc. 346, 987 (2003)
[arXiv:astro-ph/0307104].
[20] S. Hannestad, Phys. Rev. D 71, 103519 (2005) [arXiv:astro-ph/0504017].
16
|
astro-ph/9706168 | 1 | 9706 | 1997-06-16T21:05:26 | The Globular Cluster M54 and the Star Formation History of the Sagittarius Dwarf Galaxy | [
"astro-ph"
] | We present a deep color-magnitude diagram in the VI passbands of the globular cluster M54, a member of the Sagittarius dwarf galaxy. The data extend below the cluster's main sequence turn-off, allowing us to estimate the cluster's age. We find that M54 is 0.5--1.5 gigayears older than the Galactic globulars M68 and M5. In absolute terms, the age is comparable to the published age estimates of the other member clusters Arp 2 and Terzan 8, but is significantly older than the member cluster Terzan 7. An age estimate of the Sagittarius field population relative to M54 suggests that M54 is \gtrsim 3 Gyr older than the field. We discuss briefly the star formation history of the Sagittarius dwarf galaxy. | astro-ph | astro-ph | The Globular Cluster M54 and the
Star Formation History of the Sagittarius Dwarf Galaxy
Dept. of Astronomy, University of Michigan, Ann Arbor, MI 48109-1090
Andrew C. Layden1,2
Kitt Peak National Observatory, P.O. Box 26732, Tucson, AZ 85726-6732
Ata Sarajedini1
Accpeted for publication in Astrophysical Journal Letters.
ABSTRACT
We present a deep color-magnitude diagram in the V I passbands of the globular
cluster M54, a member of the Sagittarius dwarf galaxy. The data extend below the
cluster's main sequence turn-off, allowing us to estimate the cluster's age. We find that
M54 is 0.5 -- 1.5 gigayears older than the Galactic globulars M68 and M5. In absolute
terms, the age is comparable to the published age estimates of the other member clusters
Arp 2 and Terzan 8, but is significantly older than the member cluster Terzan 7. An
age estimate of the Sagittarius field population relative to M54 suggests that M54 is ∼>3
Gyr older than the field. We discuss briefly the star formation history of the Sagittarius
dwarf galaxy.
Subject headings: galaxies: individual (Sagittarius) -- galaxies: star clusters -- galaxies:
stellar content -- globular clusters: individual (NGC 6715) -- stars: Population II
7
9
9
1
n
u
J
6
1
1
v
8
6
1
6
0
7
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1.
Introduction
Ibata, Gilmore, & Irwin (1994, hereafter referred to as IGI) announced the discovery of a gas-
poor galaxy lying behind the bulge of the Milky Way at a distance of ∼24 kpc. They estimated the
overall stellar content of this galaxy (hereafter referred to as the Sagittarius dwarf galaxy, or Sgr)
1Hubble Fellow.
2Observations obtained while a staff member of the Cerro Tololo Inter-American Observatory, one of the National
Optical Astronomy Observatories (NOAO). NOAO is operated by the Association of Universities for Research in
Astronomy, Inc., under cooperative agreement with the National Science Foundation.
-- 2 --
to be similar in age, metallicity, and total luminosity to the Fornax dwarf spheroidal galaxy. They
also noted that four previously-known globular clusters (M54, Arp 2, Ter 7, and Ter 8) had roughly
the correct distances, positions on the sky, and radial velocities to be members of this galaxy.
The presence of a galaxy so nearby provides an unparalleled opportunity for detailed study
of stellar populations, chemical enrichment history, and galaxy formation and evolution.
Ibata
et al. (1997) provide an excellent review of the ensuing research on Sgr. The results of several of
these studies have increased the likelihood that the four globular clusters are members of Sgr. In
particular, Da Costa & Armandroff (1995) presented new abundances and radial velocities for the
four globular cluster candidates; their velocities agree well with the systemic velocity of Sgr (IGI).
The new isodensity map of Ibata et al. (1997) confirms that Sgr extends over the three outlying
clusters, Arp 2, Ter 7, and Ter 8.
Table 1 summarizes our current knowledge concerning the ages of Sgr and its globular clusters.
The difficulties in comparing absolute ages of stellar systems are well-known (e.g., Chaboyer 1995),
so we have used ages, when available, from the self-consistent work of Chaboyer et al.
(1996,
hereafter referred to as CDS). We use the ages from column 2 of their Table 3, which employ an
MV (RR) − [F e/H] relation nearly identical to that of Lee et al. (1990, hereafter referred to as
LDZ). CDS do not include ages for the Sgr field population, so the estimates for Sgr listed in Table
1 are less homogeneous. The confusion is compounded by the uncertainty in the metallicity of Sgr;
the dependence of derived age on assumed metallicity is well known (e.g., CDS). Still, the studies
agree on an age between 10 and 14 Gyr for the dominant population of Sgr.
The only Sgr globular cluster candidate currently without accurate main sequence turn-off
(MSTO) photometry and an age estimate is M54 (NGC 6715). Sarajedini & Layden (1995, hereafter
referred to as SL95) presented a CCD color-magnitude diagram (CMD) of the bright stars in M54
(MV ∼< +2.0). They found M54 to be metal poor ([F e/H] = −1.79 dex) and to have a blue
horizontal branch typical of old Galactic globular clusters. The recent photometry of M54 by
Marconi et al. (1997) goes deeper, but the photometric errors and incompleteness at the magnitude
of the MSTO preclude an accurate age analysis.
M54 is by far the most luminous of the four Sgr clusters (SL95), and perhaps the most secure
candidate for membership in Sgr, since it lies in the highest density region of that galaxy (IGI) and
has a distance (SL95) and radial velocity (Da Costa & Armandroff 1995) which correspond very
well with those of Sgr. SL95 have speculated that M54 is the "nucleus" of Sgr, akin to the nuclei
found in many dwarf elliptical galaxies. Clearly, we cannot have a complete understanding of the
star formation history of Sgr without information about the age of M54.
2. Observations, Reductions, and the CMD
In order to obtain photometry to the MSTO, we secured deep images of M54 using the CTIO
4.0-m telescope at prime focus during the nights of 1995 June 28 and 29. We employed the Tek#4
-- 3 --
20482 pixel CCD, which provided a scale of 0.43 arcsec pixel−1. The median seeing was 1.2 arcsec.
We measured instrumental stellar magnitudes on each frame using the DoPHOT photometry
program (Schechter, Mateo, & Saha 1993), and transformed them directly to the V I magnitude
system of SL95 using the large number of stars common to both data sets. Photometry from the
five V I frame pairs with the best seeing were assembled, and mean magnitudes and errors (standard
errors of the mean) were computed for each star detected in three or more frame pairs. The details
of the reduction procedure and the photometric data for the resulting 26,485 stars are presented
in Layden & Sarajedini (1997). Comparisons show that these data are accurately tied to the SL95
photometric system.
Figure 1 presents the V I CMDs for (a) all the M54 stars with high quality data, and (b) all
the high quality M54 stars located between 2.5 and 4.3 arcmin from the cluster center. In these
panels, the curves are the fiducial red giant branches (RGBs) of M54 and Sgr derived by SL95.
These curves, together with the CMDs of Sgr and a foreground bulge control field by Mateo et al.
(1995, hereafter referred to as MUSKKK), facilitate the interpretation of our CMD.
The curve on the left is the M54 RGB fiducial; it guides the eye faintward to where the M54
RGB becomes well populated. The M54 RGB turns blueward onto the subgiant branch (SGB) at
(V − I, V ) = (0.9, 20.9) mag, and merges with a column of points ∼0.2 mag blueward of this. As
we will see, this column represents the superimposed MSTO regions of M54 and Sgr.
The curve on the right is the Sgr RGB fiducial. The lower RGB of Sgr is not as well populated
as that of M54 in this figure, but there appears to be an excess of points roughly parallel to the
M54 RGB which terminates at (V − I, V ) = (1.0, 20.9) mag, and which presumably turns blueward
onto the SGB at this point (see MUSKKK). The MSTO of Sgr in the MUSKKK field occurs at
(V − I, V ) ≈ (0.75, 21.4) mag, the same region as the MSTO of M54 in our data. The plume of
stars at V − I = 0.75 and 20.2 < V < 20.9 mag corresponds to the young (4 Gyr) Sgr population
discovered by MUSKKK.
Other prominent sequences in Figure 1a include the M54 blue HB (SL95), the Sgr red HB
clump (MUSKKK, SL95), and a population of blue stragglers or very young stars belonging either
to M54 or Sgr (0.2 < V − I < 0.6, 19 < V < 21 mag). The MUSKKK bulge control field coincides
well with the column of stars at V − I ≈ 0.9 and V < 19.5 mag which sweeps redward at fainter
magnitudes across the M54 and Sgr lower RGBs. Thus most of the scatter with V > 20 and
V − I > 1.0 mag is attributed to the foreground bulge.
One important qualitative statement about the relative ages of M54 and Sgr can be made at
this point. The MSTOs of these populations appear to be coincident at V − I ≈ 0.75 mag. The
reddenings are identical since the populations lie in the same field. Yet the metallicity of M54 is at
least 0.5 dex lower than that of Sgr, so for the MSTOs to coincide, M54 must be older than Sgr.
-- 4 --
3. Comparison with Cluster Fiducial Sequences
A simple estimate of the age of M54 can be made by directly comparing our photometry with
the fiducial sequences of other clusters. High quality V I CCD photometry exists in the literature
for M68 ([F e/H] = −2.09) and M5 ([F e/H] = −1.40), which bracket M54 in metallicity. Figure 2
shows our data plotted with the fiducial sequence of M68 from Walker (1994, dashed line) and the
fiducial of M5 from Sandquist et al. (1996, solid line). All the cluster data were registered to the
observational HR Diagram using the V (HB), [F e/H], and E(V − I) values given in Table 2 along
with the relation MV (RR) = 0.17[F e/H] + 0.82 (LDZ).
The fiducial sequence comparison reveals that the age of M54 is comparable to those of M68
and M5. Since M54 is almost exactly between M68 and M5 in metallicity, one expects the data
for M54 to lie midway between the M68 and M5 fiducials. However, the M54 data, particularly for
the SGB, appears to be skewed slightly toward the M5 fiducial. This suggests that M54 may be
slightly older than M68 or M5. In the next section, we will quantify this age difference.
4.
Isochrone Fitting
Another method for measuring globular cluster ages is isochrone fitting. Figure 3 shows the
Revised Yale Isochrones (RYI; Green, Demarque, & King 1987) for Y = 0.23, [F e/H] = −1.50,
and ages of 10 -- 18 Gyr, superimposed on the data from Figure 1b. The isochrones were shifted to
the observed plane using the V (HB) and E(V − I) values listed in Table 2, and the LDZ relation
between MV (RR) and [F e/H] (assumed for consistency with the RYI, see King et al. 1988). The
ridge-line of M54 SGB stars suggests an age of 13 -- 14 Gyr. The 12 Gyr isochrone forms an envelope
about the MSTO points, setting a hard lower limit for the age of M54 under the stated assumptions.
We used isochrones with [F e/H] = −1.50 because the RYI employ scaled-solar abundance
ratios, whereas observations suggest that Galactic globular clusters have an enhancement of α-
elements of [α/F e] ≈ +0.4 dex (e.g., Pagel & Tautvaisiene 1995). Salaris et al. (1993) showed
that for a given iron abundance, scaled-solar isochrones 0.29 dex more metal-rich in [F e/H] closely
mimic α-enhanced isochrones with [α/F e] = +0.4 dex. RYI with [F e/H] = −1.79 indicate ages
1 -- 2 Gyr older than those shown here.
Analogous RYI fits to the data of M68 (Walker 1994) and M5 (Sandquist et al. 1996), again
using the parameters from Table 2 and the "α-enhanced" metallicities, produced ages of ∼13 Gyr
for M68 and ∼12 Gyr for M5. As in Sec. 3, M54 appears to be comparable in age to these clusters,
or perhaps slightly older.
We are concerned by the tendency for the isochrones to be bluer than the data at V ∼> 22. This
could be due to (1) differential incompleteness in our data as a function of color, (2) a tendency for
Sgr stars to dominate the red side of the lower main sequence and thus to bias the data redward, or
(3) inadequacies in the isochrones or adopted reddening and distance modulus. Though adopting
-- 5 --
a larger reddening (e.g., by 0.05 mag ≈ 2σ) corrects the main sequence color problem and makes
the derived age younger (∼2 Gyr), it degrades the fit to the lower RGB. Adjusting the reddening
and distance modulus in concert enables us to obtain a better overall fit; the age obtained is 15
Gyr for E(V − I) = 0.18 and a distance modulus 0.15 mag smaller than employed in Figure 3.
Our isochrone age estimates are supported by estimates based on the luminosity of the subgiant
branch. The difference between the magnitude of the subgiants at a well defined color and that of
the horizontal branch is similar for M54, M68, and M5. When calibrated using the RYI, we find
that M54 is 1 -- 2 Gyr older than M68 and 0 -- 1 Gyr older than M5. Details of this procedure are
presented in Layden & Sarajedini (1997).
Given the uncertainties in determining absolute ages, we would like to compare our age for
M54 to that of the Sgr field population in a relative sense. In Figure 1, the lower RGB of the Sgr
field population appears to terminate abruptly at V ≈ 20.9 mag. RYI with metallicities and ages
with lower RGBs terminating at this magnitude can be used to place an upper limit on the age of
the Sgr field. For [F e/H] = −0.50 (SL95), we find a maximum age of 6 Gyr. For [F e/H] = −1.2
and [α/F e] = +0.4, we find an age of 9 Gyr. For [F e/H] = −1.2 and [α/F e] = +0.0, we find an
age of 11 Gyr. The latter is the oldest age obtainable for Sgr which is consistent with currently
quoted abundance estimates. This age is also in good agreement with the Sgr ages listed in Table
1. Clearly, M54 must be older than the Sgr field stars by ∼>3 Gyr.
5. Discussion
All three of the methods discussed above suggest that M54 is 0.5 -- 1.5 Gyr older that the
comparison clusters M68 and M5. CDS find the age of M5 to be typical of Galactic globular
clusters of its metallicity, while M68 may be somewhat younger than average. Thus, M54 has an
age typical of Galactic globulars of its metallicity (see Figure 1 of CDS).
The absolute age estimates discussed in Sec. 4, using the LDZ relation between MV (RR) and
[F e/H] and [α/F e] = +0.4, suggest that M54 has an age of ∼14 Gyr. Comparing this with the ages
of the other Sgr globulars shown in Table 1 indicates that M54, Ter 8, and Arp 2 are all comparably
old (for more details, see Layden & Sarajedini 1997), while Ter 7 is significantly younger. Given
the uncertainties in the existing photometry, we cannot rule out the possibility that the three old
clusters in Sgr are coeval.
Comparing our absolute age for M54 with the age estimates for the dominant Sgr field popu-
lation shown in Table 1 suggests that M54 is older than the metal-rich field population in which it
is embedded. This result is supported by our analysis in Sec. 4, where we estimated the maximum
age of the Sgr field as a function of assumed [F e/H], and found that M54 is at least 3 Gyr older
than Sgr.
Taken at face value, these ages suggest that the metal-poor clusters represent the earliest epoch
-- 6 --
of significant star formation in Sgr. Vigorous star formation in the field appears to have begun
several Gyr later. Given this age difference, it seems reasonable to expect that gas expelled from
evolving metal-poor cluster stars enriched the interstellar medium and thus the first generation
of Sgr field stars. This explains, at least in part, why the Sgr field is so much more metal-rich
than the old clusters. As was the case for many of the Galactic satellite dwarf spheroidals (e.g.,
Smecker-Hane et al. 1994), Sgr managed to retain a significant portion of its gas for many Gyr,
enabling the formation of Ter 7, and of the ∼4 Gyr field population discussed by MUSKKK and
represented by the blue plume of stars above the MSTO in Figure 1. Given the age and abundance
of Ter 7, it is perhaps more appropriate to compare this cluster with the "populous clusters" of
the SMC or ESO121-SC03 in the LMC (Da Costa 1991) than with traditional globular clusters.
Finally, we note that the HB morphologies of the three old Sgr globulars are quite blue for their
metallicity (SL95, Buonanno et al. 1995, Ortolani & Gratton 1990), in better agreement with the
Galactic globular clusters than those of the Fornax dwarf galaxy. In this respect, Sgr may be a
better example of a surviving "building block" of the Galactic halo than Fornax (see Zinn 1993).
We thank Mario Mateo for his thoughtful comments. A.C.L. was supported by NASA grant
number HF-01082.01-96A, and A.S. was supported by NASA grant number HF-01077.01-94A,
from the Space Telescope Science Institute, which is operated by the Association of Universities for
Research in Astronomy, Inc. under NASA contract NAS5-26555.
REFERENCES
Buonanno, R., Corsi, C.E., Fusi Pecci, F., Richer, H.B., & Fahlman, G.G. 1995, AJ, 109, 650
Chaboyer, B. 1995, ApJ, 444, L9 |
astro-ph/9508089 | 1 | 9508 | 1995-08-18T23:37:05 | Galactic extinction and Abell clusters | [
"astro-ph"
] | In this paper, we present the results of comparing the angular distribution of Abell clusters with Galactic HI measurements. For most subsamples of clusters considered, their positions on the sky appear to be anti-correlated with respect to the distribution of HI column densities. The statistical significance of these observed anti-correlations is a function of both richness and distance class, with the more distant and/or richest systems having the highest significance (~3 sigma). The lower richness, nearby clusters appear to be randomly distributed compared to the observed Galactic HI column density. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. , { ( )
Galactic extinction and Abell clusters
R. C. Nichol
and A. J. Connolly
Department of Astronomy and Astrophysics, University of Chicago, S. El lis Ave, Chicago, Il linois , USA
Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD , USA
ABSTRACT
In this paper, we present the results of comparing the angular distribution of Abell
clusters with Galactic HI measurements. For most subsamples of clusters considered,
their positions on the sky appear to be anti-correlated with respect to the distribution
of HI column densities. The statistical signi(cid:12)cance of these observed anti-correlations is
a function of both richness and distance class, with the more distant and/or richest sys-
tems having the highest signi(cid:12)cance (' (cid:27)). The lower richness, nearby clusters appear
to be randomly distributed compared to the observed Galactic HI column density.
Key words: surveys-galaxies:clusters:general-dust,extinction-large-scale structure in
the universe
August
5
9
9
1
g
u
A
8
1
9
8
0
8
0
5
9
/
h
p
-
o
r
t
s
a
INTRODUCTION
Obscuration due to dust in our own Galaxy is a ma jor sys-
tematic problem for most areas of extragalactic astronomy.
In particular, it is a severe problem for (cid:13)ux{limited, optical
surveys of cosmological ob jects (e.g. clusters of galaxies) as
these surveys are used for statistical studies of the large-
scale structure in the universe (Nichol et al. ). A spu-
void of clusters reported by Bahcall & Soneira ( ). Fur-
thermore, all investigations of the e(cid:11)ects of extinction on
the Abell cluster distribution have only implemented inter-
nal consistency checks, by utilizing the observed surface den-
sity of clusters compared to that expected from a uniform
distribution. This does not, however, account for genuine
large-scale structure in the cluster distribution which will
certainly confuse the issue.
rious clustering signal can be introduced into these surveys
With the recent availability of high quality,
large{
via patchy Galactic extinction. This false signal may then
area independent extinction indicators, like the Stark et al.
be interpreted as evidence for large scale structure.
( ) HI radio map of the Galaxy, it is now timely to revisit
The Abell catalogue (Abell ) remains the most
the question of Galactic extinction and its e(cid:11)ects on the ob-
widely used survey of clusters for statistical studies of the
served cluster distribution. It is now possible to check for
cluster large-scale distribution. Abell was well aware of the
signi(cid:12)cant anti{correlations between the observed positions
potential e(cid:11)ects of extinction on his catalogue and corrected
and classi(cid:12)cations of clusters and the measured obscuration
all his galaxy magnitudes for the estimated e(cid:11)ect. He con-
as derived from such independent, external
indicators. In
cluded that Galactic obscuration plays a role in the observed
this paper, we present the results of such an investigation
distribution of clusters of galaxies and noted that the sur-
since the e(cid:11)ects of Galactic extinction may represent the
face density of clusters decreased rapidly as a function of
largest systematic bias confronting statistical analyses of the
Galactic latitude, as well as in a few areas of anomalously
cluster distribution and constrains attempts to construct ho-
low cluster density at high latitudes (e.g. an area at l =
mogeneous, complete samples of clusters.
rising as high as b = + ).
Subsequent analyses of the Abell catalogue have taken
Abell's advice to heart, by including a Galactic latitude se-
lection function (P (b)) of the form
HI RADIO DATA
P (b) =
;
()
The Stark et al. ( ) HI radio map is the cleanest survey of
:((cid:0)cosec jbj)
Galactic atomic neutral hydrogen (HI) presently available. It
where b is Galactic latitude (see, for example, Abell ;
is complete above a declination of (cid:0) and was constructed
Bahcall & Soneira ; Postman, Huchra & Geller ).
using the AT&T Bell Laboratories foot horn re(cid:13)ector at
Yet few investigators have worried about possible Galactic
Crawford Hill. The telescope has a FWHM beamwidth of
(cid:14)
longitudinal dependences of Galactic extinction (even the
and is relatively free from side-lobe contamination which is
anomalous patches highlighted by Abell), and if they have,
a ma jor advantage over previous work. As much as % of
they have only considered the extreme cases e.g. the large
the measured radiation in a particular direction can be due
R. C. Nichol and A. J. Connol ly
Figure . The real and random cumulative distributions for the
D > (top) and R = & D (cid:20) (bottom) subsamples. It is
to contamination from the side-lobes which are typically lo-
cated tens of degrees away from the pointing direction. Such
problems make it very di(cid:14)cult to map accurately the am-
bient interstellar hydrogen and can lead to false detections
(Hartmann ). One of the main drawbacks of the Stark
et al. map is its' low resolution. However, for the work pre-
sented here on the cluster distribution, it is unlikely that
this will be a ma jor constraint since the surface density of
Abell clusters is low ((cid:24) : per deg
).
The map of Stark et al. was used to interpolate the
value of the HI column density (atoms cm
) for any given
(cid:0)
line-of-sight in the sky north of (cid:14) = (cid:0)
. A recent compar-
(cid:14)
ison of such an interpolation method against data obtained
from higher resolution HI observations has shown that it
works reasonably well (Elvis, Lockman & Fassnacht ).
In % of the cases, the absolute di(cid:11)erence between the in-
terpolated and measured HI column densities is less than
(cid:24) (cid:2)
atoms cm
, or, (cid:24) % of the typical HI values
(cid:0)
used below. This error is an order of magnitude smaller than
the observed spread in HI values seen over the region of the
sky considered here and therefore, is expected to have little
e(cid:11)ect on our results. For the remaining % of cases, the dis-
tribution of the di(cid:11)erences between the two values does show
a signi(cid:12)cant tail with the higher resolution measurements
having higher HI values than those interpolated from Stark
et al. (the largest discrepancy being (cid:24) (cid:2)
atoms cm
,
(cid:0)
or, (cid:24) %). These discrepant points appear to have no
strong correlation with the observed HI values, although
most lie at large HI column densities. Therefore, it would
appear that the Stark et al. data, in high HI column den-
sity regions, occasionally underestimate the true HI column
density. Therefore, if we detect any anti{correlation between
the clusters and the HI column densities it will be an under-
estimate.
HI-ABELL CORRELATIONS
For the analysis presented here, we concentrate on the origi-
nal northern Abell survey (Abell ) constraining the de-
clinations of the clusters to be above (cid:14) = (cid:0)
(the Abell
(cid:14)
catalogue stops at (cid:14) = (cid:0)
). Furthermore, we only consider
(cid:14)
apparent from these (cid:12)gures that the more distant Abell clusters
clusters above a Galactic latitude of b = + . This latitude
tend to be at lower HI column densities than that expected from
cut is more severe than those suggested by Abell (Table
random lines{of{sight, while the lower richness systems appear
in his paper), since we wish to investigate areas of the sky
randomly distributed compared to the Galactic HI. Table con-
commonly believed to be free of obscuration e(cid:11)ects.
tains the results of KS tests on these distributions. The inset
The HI column density values interpolated from the
plots shows the distributions of Galactic latitudes for these Abell
Stark et al. map were used as tracers of the Galactic ex-
clusters and the random datasets.
tinction, assuming a constant dust{to{gas ratio. A typical
conversion between HI column density and visual extinction
is
(cid:0)
A
= : (cid:2)
N (H I )
()
B
where A
is the B magnitude extinction and N (H I ) is the
B
HI column density (Nichol & Collins ). Only relative ex-
tinctions between di(cid:11)erent parts of the sky are important,
since we are not attempting to de(cid:12)ne the absolute complete-
ness of the Abell cluster catalogue. The validity of using HI
column densities, and the Stark et al. data, as a tracer of
the Galactic extinction was assessed by Boulanger & Perault
( ) using high resolution far-infrared maps derived from
the IRAS satellite. At (cid:22)m, most of the di(cid:11)use emission
Galactic extinction and Abel l clusters
Figure . Histogram of mean HI column densities for ran-
Figure . The real and random cumulative distributions for the
dom runs of the D > Abell cluster subsample. Also plotted is
D > &
(cid:20) l (cid:20)
Abell subsample. The inset plot again
(cid:14)
(cid:14)
the mean determined for the real cluster dataset and the standard
shows the Galactic latitude distributions for the Abell clusters
error on that measurement. This demonstrates that such a low
and the random dataset.
mean value of HI is not expected from random lines-of-sight.
seen by IRAS is assumed to be thermal radiation from dust
data sample. A demonstration of this is shown in Fig.. In all
in the Galaxy. They discovered that at high Galactic lati-
cases, the distribution of random samplings was Gaussian, as
tudes (b > ) the measured HI emission and IRAS (cid:22)m
expected from the Central Limit Theorem. The signi(cid:12)cance
(cid:13)ux were well correlated, both as a function of Galactic lat-
of the observed Abell HI mean value can then be derived
itude and longitude. Moreover, the correlation was linear
from the mean and standard deviation of these generated
constraints and using the same number of clusters as the real
thus justifying a constant dust{to{gas ratio. We have used
distributions.
the HI data in preference to the IRAS data because of con-
tinuing uncertainties in the zodiacal light subtraction from
the IRAS data.
DISCUSSION
For various selections of Abell clusters (see Table ), the
interpolated HI Stark et al. values towards all the clusters
We have constrained the analysis presented here to areas of
were computed and the observed distribution of these HI
the sky believed to be free of extinction problems. However,
values was then compared to the observed HI distribution for
it is clear from Table that for most of the Abell subsamples
randomly selected lines-of-sight. In all cases, the number of
considered the distribution of clusters appears to be anti{
random directions was twenty times greater than that of the
correlated with respect to the implied distribution of Galac-
real data and they were selected under the same coordinate
tic extinction. All the samples examined here have a lower
constraints as the real data discussed above. In addition,
mean HI column density than that computed for random
the random directions were constrained to follow the same
lines-of-sight. The statistical signi(cid:12)cance of these departures
observed Galactic latitude dependence which was achieved
from random expectations varies between the samples.
by (cid:12)tting the observed surface density of clusters for each of
Taking all Abell clusters above b > + , irrespective of
the di(cid:11)erent selections in Table (see Fig. ). This ensured
their richness (R) and distance class (D; see Abell for
that the random catalogues covered the same part of the sky
the full de(cid:12)nition of these), it appears that their positions
and included the same, known selection bias.
are signi(cid:12)cantly anti-correlated with respect to the implied
The real and random HI cumulative distributions for
HI column density. This is not due to the known strong
each of the di(cid:11)erent cluster subsamples were compared using
Galactic latitude selection function since this has already
a Kolmogorov-Smirnov (KS) test and a probability derived
been incorporated into the analysis. If we now investigate
that the two distributions were drawn from the same parent
this anti-correlation as a function of richness (Table ), the
distribution. The results of these KS tests are shown in Table
smaller D (cid:20) and R (cid:21) sample still shows a signi(cid:12)cant
, while Fig. shows two examples of the real and random
anti{correlation as de(cid:12)ned by the lower than expected mean
cumulative distributions.
HI value and the high KS probability ( .% probability
As a further test, the mean HI column density and stan-
that the real and random HI distributions were not drawn
dard error were calculated for each of the samples of Abell
from the same parent distribution). The signi(cid:12)cance of the
clusters given in Table . In addition, for each sample, a
anti{correlation diminishes as a function of richness, with
distribution of random mean HI values was constructed by
the R = clusters having a mean HI value and KS proba-
randomly sampling the HI data under the same coordinate
bility consistent with a random dataset. In contrast, when
R. C. Nichol and A. J. Connol ly
Table . The results of comparing the positions of Abell clusters with measured HI column densities from the Stark et al. ( ) survey
in units of
atoms cm
. Various Abell subsamples are considered and presented here along with total number of clusters used in the
(cid:0)
analysis. The KS probability and mean observed Abell HI are also shown. The last column contains the expected mean HI as derived
from randomly sampling the data.
Sample
N
P
hH I i
Mean hH I i
K S
b > +
.
(cid:6)
(cid:6)
b > + , D (cid:20) & R (cid:21)
.
(cid:6)
(cid:6)
b > + , D (cid:20) & R (cid:21)
.
(cid:6)
(cid:6)
b > + , D (cid:20) & R =
.
(cid:6)
(cid:6)
b > + , D >
.
(cid:6)
(cid:6)
b > +
.
(cid:6)
(cid:6)
b > +, D (cid:20) & R (cid:21)
.
(cid:6)
(cid:6)
b > +, D (cid:20) & R =
.
(cid:6)
(cid:6)
b > +, D >
.
(cid:6)
(cid:6)
b > j j, Postman et al. sample
.
(cid:6)
(cid:6)
b > j j, Postman et al. sample, R =
.
(cid:6)
(cid:6)
b > j j, Postman et al. sample, R (cid:21)
.
(cid:6)
(cid:6)
b > + , D > &
(cid:20) l (cid:20)
.
(cid:6)
(cid:6)
(cid:14)
(cid:14)
b > + , D > &
< l (cid:20)
.
(cid:6)
(cid:6)
(cid:14)
(cid:14)
b > + , D > &
< l (cid:20)
.
(cid:6)
(cid:6)
(cid:14)
(cid:14)
clusters with D > are investigated, irrespective of richness
a high mean extinction value and a low surface density of
(most are R > ), it is apparent than a large fraction of the
clusters (as originally highlighted by Abell), our analysis in-
observed anti-correlation between the whole Abell sample
dicates that the positions of the clusters in this region
and the HI column densities comes from this distant cluster
are not preferentially located in areas of lower than average
sample. This is consistent with Abell's original hypothesis.
extinction; they are e(cid:11)ectively randomly scattered with re-
Similar trends are also seen if we repeat this analysis only
spect to the extinction. The
(cid:20) l (cid:20)
region has a
(cid:14)
(cid:14)
for clusters at Galactic latitudes of b > +, re-enforcing
higher surface density of clusters than expected from ran-
the fact that this is not a Galactic latitude e(cid:11)ect.
dom which could be used as an indicator of low extinction.
We have also carried out an analysis on the sample of
However, our analysis tentatively suggests that the positions
Abell clusters used by Postman et al. ( ), which is the
of these clusters are anti{correlated with the HI column
largest sample of Abell clusters with measured redshifts.
density. We carried out an analysis for the area bounded by
The sample constitutes all Abell ( ) northern clusters,
(cid:20) l (cid:20)
and
(cid:20) b (cid:20)
which coincides with the
(cid:14)
(cid:14)
(cid:14)
(cid:14)
irrespective of richness, with an m
(cid:20) : (where m
void discussed by Bahcall & Soneira ( ) and found the
is the magnitude of the tenth cluster galaxy). This sample
same level of signi(cid:12)cance for the observed anti-correlation.
has been extensively used to study the large{scale cluster-
This suggests that extinction may play a part in the distri-
ing properties of clusters. The (cid:12)ndings of our analysis are
bution of clusters in this part of the sky.
consistent with those mentioned above for the more generic
This analysis does indicate the severe problems in using
samples of Abell clusters considered here and indicates that
(cid:13)uctuations in the observed surface density of clusters as an
the observed angular distribution of this important Abell
indicator of extinction. It is extremely di(cid:14)cult is separate
subsample maybe in(cid:13)uenced by Galactic extinction.
extinction{induced clustering from real
large{scale struc-
As a further test, the data was cut into three separate
ture. Furthermore, it highlights potential longitudinal gra-
Galactic longitude segments (see Table ) to test for any
dients in the distribution of Abell clusters. Such an e(cid:11)ect is
longitudinal dependence on the observed anti-correlation i.e.
not normally considered in statistical studies of the cluster
was any particular region of the sky responsible for all the
distribution and could introduce false large{scale signals if
signal? We carried out such an analysis using the D > sub-
the gradients are caused by extinction.
sample as this had the most signi(cid:12)cant anti{correlation. The
The work presented here
implies
subtle
anti{
(cid:14)
(cid:14)
< l (cid:20)
region of the data coincides with the area
correlations between the classi(cid:12)cations of Abell clusters and
originally highlighted by Abell and contains an anomalously
the observed distribution of HI column densities. It is rela-
low surface density of clusters, while the
< l (cid:20)
(cid:14)
(cid:14)
tively straightforward to understand why the more distant
segment covers an area discussed by Bahcall & Soneira
clusters appear to be discovered in regions of lower than
( ) which appears to have an apparent void of Abell clus-
average inferred extinction, since in these areas the pho-
ters.
tographic plates are more likely to reach fainter (cid:13)ux lim-
Once again, all three segments have a lower observed
its thus aiding their detection and classi(cid:12)cation. The anti-
mean HI value than that expected from random directions,
correlation between richer Abell systems and HI column den-
yet the only one with a compelling statistical signi(cid:12)cance is
sities is less obvious, but is probably a combination of several
(cid:14)
(cid:14)
the
(cid:20) l (cid:20)
region. We present the cumulative real
factors; an increased e(cid:14)ciency in counting and classifying
and random distributions for this segment in Fig. . This
galaxies in the lower relative extinction regions, an inap-
is contrary to what may have been expected in light of the
propriate background subtraction and a larger Abell radius
above remarks. Although the
(cid:20) l (cid:20)
region has
compared to other clusters of similar intrinsic properties in
(cid:14)
(cid:14)
Galactic extinction and Abel l clusters
higher extinction regions. As an example, . magnitudes of
relative extinction can boost the richness of an Abell clus-
ter by upto % compared to similar clusters nearby on
the sky. Extinction{induced di(cid:11)erences in the background
galaxy correction can account for approximately % of this,
while the remainder comes from a larger inferred Abell ra-
dius which is proportional to m
and would thus be arti(cid:12)-
cially fainter (smaller radius) in the higher extinction region.
Finally, this work does highlight the problems of ex-
tinction in constructing and statistically analysing cluster
surveys. In recent years, several groups have begun to con-
struct automated samples of clusters from digitised galaxy
counts e.g. Lumsden et al. ( ). Such an automated ap-
proach will easily allow the e(cid:11)ects of extinction, as derived
from independent information, to be included into the se-
lection and classi(cid:12)cation of clusters thus removing the need
to correct for it later. This point is emphasized by carry-
ing out an identical correlation analysis on the automated
Edinburgh/Durham Cluster Catalogue (EDCC; Lumsden et
al. ). Although this catalogue was not explicitly cor-
rected for extinction using external
indicators, an initial
step in its construction involved the subtraction of a sky
frame which represented the large{scale (cid:13)uctuations (scales
of (cid:24) :
) seen in the original galaxy distribution (Lumsden
(cid:14)
et al. ). This process certainly helped in reducing the
e(cid:11)ects of extinction, since the typical coherence length of the
patchy extinction is '
!
(Nichol & Collins ).
(cid:14)
(cid:14)
No correlation, signi(cid:12)cant or not, was found between the
observed HI column densities and the richnesses, distances
or positions of the clusters.
ACKNOWLEDGMENTS
Thanks to Rich Kron for his initial insight. AJC acknowl-
edges partial support from NSF grant AST- , while
we both acknowledge partial support from ADP grant NAG
- . This research has made use of some data obtained
through the High Energy Astrophysics Science Archive
Research Center Online Service, provided by the NASA-
Goddard Space Flight Center. Finally, we are very grateful
to Stuart Lumsden, the referee, for his helpful comments on
the paper.
REFERENCES
Abell G.O., , ApJS, ,
Bahcall N.A., Soneira R.M., , ApJ, ,
Bahcall N.A., Soneira R.M., , ApJ, ,
Boulanger F., Perault M., , ApJ, ,
Elvis M., Lockman F.J., Fassnacht C., , ApJS, ,
Hartmann D., , PhD Thesis, Univ. of Leiden
Lumsden S.L., Nichol R.C., Collins C.A., Guzzo L., ,
MNRAS, ,
Nichol R.C., Collins C.A., Guzzo L., Lumsden S.L., ,
MNRAS, , P
Nichol R.C., Collins C.A., , MNRAS, ,
Postman, M., Huchra J.P., Geller M.J., , ApJ, ,
Stark A.A., Gammie C.F., Wilson R.W., Balley J., Linke
R.A., Heiles C., Hurwity M., , ApJS, ,
|
0812.3682 | 1 | 0812 | 2008-12-18T23:39:38 | Submillimeter Continuum Properties of Cold Dust in the Inner Disk and Outflows of M82 | [
"astro-ph"
] | Deep submillimeter (submm) continuum imaging observations of the starburst galaxy M82 are presented at 350, 450, 750 and 850 micron wavelengths, that were undertaken with the Submillimetre Common-User Bolometer Array (SCUBA) on the James Clerk Maxwell Telescope in Hawaii. The presented maps include a co-addition of submm data mined from the SCUBA Data Archive. The co-added data produce the deepest submm continuum maps yet of M82, in which low-level 850 micron continuum has been detected out to 1.5kpc, at least 10% farther in radius than any previously published submm detections of this galaxy. The overall submm morphology and spatial spectral energy distribution of M82 have a general north-south asymmetry consistent with H-alpha and X-ray winds, supporting the association of the extended continuum with outflows of dust grains from the disk into the halo. The new data raise interesting points about the origin and structure of the submm emission in the inner disk of M82. In particular, SCUBA short wavelength evidence of submm continuum peaks that are asymmetrically distributed along the galactic disk suggests the inner-disk emission is re-radiation from dust concentrations along a bar (or perhaps a spiral) rather than edges of a dust torus, as is commonly assumed. Higher resolution submm interferometery data from the Smithsonian Submillimeter Array and later Atacama Large Millimeter Array should spatially resolve and further constrain the reported dust emission structures in M82. | astro-ph | astro-ph |
Version June 7, 2018, in press at AJ
Submillimeter Continuum Properties of Cold Dust
in the Inner Disk and Outflows of M 82
Lerothodi L. Leeuw1,2,3 and E. Ian Robson4
ABSTRACT
Deep submillimeter (submm) continuum imaging observations of the starburst
galaxy M 82 are presented at 350, 450, 750 and 850 µm wavelengths, that were
undertaken with the Submillimetre Common-User Bolometer Array (SCUBA)
on the James Clerk Maxwell Telescope in Hawaii. The presented maps include a
co-addition of submm data mined from the SCUBA Data Archive. The co-added
data produce the deepest submm continuum maps yet of M 82, in which low-
level 850 µm continuum has been detected out to 1.5 kpc, at least 10% farther in
radius than any previously published submm detections of this galaxy. The over-
all submm morphology and spatial spectral energy distribution of M 82 have a
general north-south asymmetry consistent with Hα and X-ray winds, supporting
the association of the extended continuum with outflows of dust grains from the
disk into the halo. The new data raise interesting points about the origin and
structure of the submm emission in the inner disk of M 82. In particular, SCUBA
short wavelength evidence of submm continuum peaks that are asymmetrically
distributed along the galactic disk suggests the inner-disk emission is re-radiation
from dust concentrations along a bar (or perhaps a spiral) rather than edges of
a dust torus, as is commonly assumed. Higher resolution submm interferome-
tery data from the Smithsonian Submillimeter Array and later Atacama Large
Millimeter Array should spatially resolve and further constrain the reported dust
emission structures in M 82.
1Space Science and Astrophysics Branch, NASA Ames Research Center, MS 245-6, Moffett Field, CA
94035; [email protected].
2Department of Physics & Electronics, Rhodes University, PO Box 94, Grahamstown 6140, South Africa;
[email protected].
3SA SKA / MeerKAT, Lonsdale Building, Lonsdale Road, Pinelands, 7405, South Africa;
[email protected].
4Astronomy Technology Centre, Royal Observatory, Blackford Hill, Edinburgh EH9 3HJ, United King-
dom; [email protected].
-- 2 --
Subject headings: dust, extinction - galaxies: individual (M 82) - galaxies: star-
burst - radiation mechanisms: thermal - submillimeter
1.
Introduction
Massive ejections of gas and dust originating in galactic nuclei have been observed at
scales of a few kpc in optical emission lines, submillimeter (submm) molecular lines, mid-
infrared (MIR), ultraviolet (UV), submm and radio continuum emission, and soft X-rays
(e.g., Watson et al. 1984; Heckman et al. 1990; Devine & Bally 1999; Hoopes et al. 2005;
Engelbracht et al. 2006). One explanation for the outflows is that a high supernova rate in
the galactic nucleus heats up the surrounding gas to high temperatures with sound speeds
exceeding the escape velocity of the galaxy, creating a wind that expands outward from the
galaxy (Chevalier & Clegg 1985). The wind entrains cosmic rays, warm and cool gas, as
well as cool dust, making the outflow directly visible in many wavebands. The outflows are
usually oriented along the minor axes of the galaxies and are thus most easily observed in
edge-on galaxies.
In local starbursts and high-z Lyman Break galaxies with high enough
global star-formation rate per unit area, the superwinds are common and responsible for
expelling metals from these galaxies and enriching the inter-galactic medium (IGM) and
therefore have implications in the evolution of galaxies and the IGM (see, e.g, Heckman
2003; Veilleux et al. 2005, for recent reviews).
M 82 (NGC 3034) is a nearby and popular object in which to investigate the physical
association between galactic nuclei and large-scale outflows. The galaxy is edge-on with
an inclination of about 10◦ at position angle 72◦ and is classified as IrrII. At an estimated
distance of 3.63 Mpc (as determined for M 81 by Freedman et al. (1994)), M 82 has optical
dimensions of 11′.2 × 4′.3, i.e. ∼ 11.8 × 4.5 kpc (image scale is ∼17.6 arcsec−1). The nuclear
region, within 4′ ×2′ about the major axis of the galactic disk, has numerous point sources or
emission concentrations, some of which originate from supernovae and massive star clusters
and have been detected from the X-ray to radio wavebands. Layers of dust filaments laden
these inner regions producing severe optical extinction and copious infrared to submm re-
radiated emission.
New Hubble Heritage Team optical images obtained with a deep 6-point mosaic in B
(0.45 µm), V (0.55 µm), I (0.81 µm), and Hα (0.65 µm) filters of the Advance Camera for
Surveys (ACS) on board the Hubble Space Telescope (HST ) exhibit detailed, filamentary
outflows of the M 82 (Mutchler et al. 2007), especially in Hα. As described above, it is
thought that this outflow is being driven by the copious formation of massive stars (or a
starburst) and subsequent explosions of supernovae. The starburst outflow not only provides
-- 3 --
the ejection mechanism for the material from the galactic nucleus, but also heats the gas
and ionizes the hydrogen, causing it to glow with the red light of Hα emission line.
These new optical images as well as earlier detailed natural-color composite images of
M 82 obtained with HST (see, e.g., de Grijs 2001) and the Subaru Telescope (Ohyama et al.
2002) show more than 100 compact groupings of about 105 stars in very bright star clusters
sprinkled throughout the galaxy's central region, prominent dust lanes that crisscross the
disk, knotty filaments of ionized gas that have rich nebular spectra that are not especially
enriched in nitrogen, hydrogen gas in a strong galactic wind that is clearly below the galactic
center and to the right of the central region, along with many other regions of varying star-
formation environments in the nuclear parts of this galaxy. The huge clusters of massive
stars, numerous X-ray and radio detected supernovae, gas concentrations, optically-dramatic
dust filaments, galactic winds and other active nuclear features have been attributed to a
large burst of star formation 107 to 108 years ago, that was probably triggered by a tidal
interaction with the nearby spiral galaxy M 81 and dwarf starburst galaxy NGC 3077 (see,
e.g., Yun et al. (1994) for evidence of H I tails linking the three and Forster Schreiber (2000)
for a recent review).
The proximity of this galaxy makes it possible to observe the region of interaction be-
tween the star-formation regions and the halo, including expanding shells or bubbles and
"chimneys" that are producing a clearer picture of the localized driving mechanisms for
the outflows (e.g., Heckman et al. 1990; Wills et al. 1999; Westmoquette et al. 2007). The
proximity also allows the detections of low-level emission in the halo and consequently the
determination of the amounts and possible origin of material that results in this emission
(e.g., Seaquist & Clark 2001; Engelbracht et al. 2006). Studying the contents and interac-
tions between the star-formation regions and the halo is important for understanding their
role in the evolution of M 82 and may provide clues to general galaxy evolution as well as
details of the composition of the intergalactic material.
This paper focuses on the Submillimetre Common-User Bolometer Array (SCUBA)
maps of the copious submm re-radiated emission that results from the dust-laden, star-
forming disk, as well as large-scale, low-level emission that is associated with the outflows
in the halo of M 82. Submm continuum observations of the dusty central regions in M 82
were previously obtained with the submm continuum receiver UKT 14 on the James Clerk
Maxwell Telescope (JCMT) by Hughes et al. (1990, 1994) and later with SCUBA, also on
the JCMT, by Leeuw et al. (1999) and Alton et al. (1999). The current study was intended
to extend these previous imaging submm observations in spatial extent, sensitivity and to
all the four submm wavelengths available with the SCUBA array (see Section 2). The study
includes a co-addition of data mined from the SCUBA Data Archive.
-- 4 --
The co-added data produce the deepest submm continuum maps yet of M 82, in which
low-level emission is detected out to 1.5 kpc for the first time in the submm continuum of
this galaxy. The deep maps are used in a detailed morphological study of the nuclear and
large-scale detections (see Sections 3 and 5), including a focused comparative analysis to op-
tical (see Section 3.2) and high-resolution CO (1-0) (see Section 3.1) morphology. The maps
are also used in the computation the first submm spatial spectral energy distribution (SED)
of separate locations in the nuclear star-forming region of M 82 (see Section 5). These obser-
vational results are used in the discussion of the origin and structure of submm continuum
morphology and spatial SED of M 82 and reviewed in the context of relevant interpretations
by other researchers, including those who use data from other wavelengths (see Section 5).
In particular, (a) the commonly assumed interpretation that the double emission peaks that
are seen in the mm to infrared continuum are due to emission from the edges of an inclined,
dusty-molecular torus is challenged (see Section 5.1), (b) an analytical review of CO results
is undertaken to assess if CO emission may significantly contaminate the continuum observed
in SCUBA filters (see Section 4), and (c) a morphological comparison is conducted to check
whether the localized outflows that are reported in high resolution radio, CO, and SiO maps
respectively by Wills et al. (1999), Weiss et al. (1999), and Garc´ıa-Burillo et al. (2001) can
be seen in the SCUBA maps (see Section 5.3). Finally, the overall implication of the results
are discussed and possible future work is outlined (see Section 6).
2. Observations
SCUBA 850, 750, 450 and 350 µm imaging observations of M 82 were obtained with
the telescope pointed at the 2.2 µm infrared nuclear peak of the galaxy using positions from
Dietz et al. (1986). Jiggle mapping observations were conducted with the secondary chopping
in azimuth at 7.8 Hz and with a throw of 120 ′′. The imaging observations employed the
common 64-point jiggle pattern with a 3 ′′ offset between each position, giving fully sampled
images with both arrays.
Because M 82 has been a popularly observed source with SCUBA, additional maps that
were obtained with a chop throw of 120 ′′ by other observers were mined from the SCUBA
Data Archive in order to co-add the related data and maximize the signal-to-noise in the final
maps. Data sets from the SCUBA Archive were separately flux calibrated and corrected for
JCMT pointing errors and then co-added, with each observation being weighted according
to its relative integration time and the noise in the map. The 450-850 µm dual mapping
wavebands have been used more commonly than the 350-750 µm combination, and therefore
the archival maps constitute about 85% and 35% of the respective co-added, total-integration
-- 5 --
time for the 450-850 µm and 350-750 µm dual maps.
The imaging data analysis was undertaken using the dedicated SCUBA data reduction
software SURF (Jenness et al. 1998), as well as KAPPA, GAIA and CONVERT software
packages provided by the Starlink Project1. The data reduction consisted of first flatfielding
the array images, and then correcting for atmospheric extinction. Next, pixels significantly
noisier than the mean were blanked-out; and, after initial inspection of raw images, pixels
containing relatively little flux from the source were used to correct for correlated sky noise
in each individual jiggle-map.
The atmospheric opacity, τ , was determined from skydips made with SCUBA at intervals
during the observations. Were SCUBA-skydip measurements were not available, the τ at
the SCUBA filters was extrapolated from the continuously measured τ at 225 GHz, obtained
courtesy of the Caltech Submillimeter Observatory (CSO) radiometer and using relations
listed in Table 1. At the JCMT, these relations are empirically derived and periodically
updated and improved as more data, especially since the commissioning of SCUBA, have
been obtained (e.g., Archibald et al. 2002)2. All the data were calibrated using instrumental
gains determined primarily from nightly beam maps of Mars and Uranus, or, alternatively,
the JCMT secondary calibrators.
Table 1: The CSO relations used on the data presented in this paper
τ850 = 4.3 × (τCSO − 0.007)
τ450 = 23.9 × (τCSO − 0.01)
τ450 = 6.5 × (τ850 − 0.03)
3. The General Submm Continuum Morphology
Figure 1 shows the 850, 750, 450 and 350 µm co-added maps of M 82 that were obtained
in jiggle mapping observing mode, respectively at 14′′.5, 11′′.4, 8′′.5, and 6′′.7 resolution and
about 8, 100, 225, and 650 mJy/beam sensitivity. The 850 and 750 µm images have a single
emission peak that is centered about 9′′ west of the galactic nucleus, while the 450 and
350 µm maps have two emission peaks centered about 10′′ and 6′′ respectively east and west
1The Starlink Project is run by the Council for the Central Laboratory of the Research Councils on behalf
of the Particle Physics and Astronomy Research Council of the United Kingdom.
2This
latest
documentation
is
available
at
the
Joint Astronomy Centre web
site,
http://www.jach.hawaii.edu
-- 6 --
of the nucleus along the galactic disk. The peak in the west is slightly elongated along the
galactic disk and is brighter than the eastern peak, showing east-west asymmetry about the
nucleus. In the 750µm image, the single emission peak seen in the 850µm image begins to be
resolved-out into the double peaks seen in the 450 and 350µm images (e.g. Figures 1, bottom
panel). This is expected, as the 750µm observations have a resolution that is intermediate
between that of the 850 and 450µm. When the maps at shorter wavelengths (e.g., 350µm)
are smoothed to resolutions similar to those at the longer wavelengths (e.g., 850µm), the two
submm peaks that are resolved at the shorter wavelengths become visible just as one peak,
showing that indeed the different high-brightness morphologies depicted in the maps are
due to the respective resolutions at different wavelengths. The overall extended morphology
has an elliptical shape with the major axis position angle of 72◦, i.e.
roughly the same
as the galactic disk in the nuclear region. This general morphology is similar to previous
continuum observations at mm (e.g., Kuno & Matsuo 1997; Thuma et al. 2000), submm
(Hughes et al. 1994; Leeuw et al. 1999; Alton et al. 1999) and mid-infrared (Telesco et al.
1991) wavelengths, as well as to CO line transitions (e.g., Nakai et al. 1987; Thuma et al.
2000) observations of comparable resolution.
It is noted that although the maps presented in Figure 1 have similar features to those
in published submm continuum maps by Hughes et al. (1994) and Alton et al. (1999), the
very sensitive 850 µm map presented here also shows that the 850 µm emission (that is
expected to be from cold dust) extends by at least 10′′ (∼ 176 pc) radius farther into the
halo than detected by those authors. Furthermore, Figure 2 depicts the 850 µm continuum
emission and integrated CO(2-1) line intensity maps respectively presented in Figure 1 of
this paper and Figure 5 of work by Thuma et al. (2000) to show that the 850 µm emission
is as extended as the CO(2-1) emission, contrary to claims by Thuma et al. (2000) that the
CO(2-1) emission is more extended than the cold dust emission of M 82, and therefore made
the galaxy exceptional in this regard.
3.1. Submm vs. High-Resolution CO (1-0) Morphology
Figure 3 shows black contours of the integrated-intensity-CO (1-0) data, that were ob-
tained with the Berkeley Illinois Maryland Association (BIMA) interferometer by Shen & Lo
(1995), overlaid on the SCUBA 450 and 850 µm continuum images by aligning the sky co-
ordinates in the SCUBA maps to those of the BIMA ones. The BIMA maps are at resolution
2′′.4 × 2′′.6 and are plotted to investigate the spatial correspondence between the submm
continuum morphology and the high-resolution CO features. The top panels in Figure 3 are
plotted at full resolution of the BIMA CO contours and the bottom panels are with the CO
-- 7 --
contours smoothed to beam sizes similar to the SCUBA-850 and -450 µm beams (∼ 14′′.5
and 8′′.5, respectively). Because dust often occurs mixed-in with gas in star-forming regions,
the BIMA maps are expected to give an indication of how dust emission may appear at
higher resolution and perhaps also some insight about the structure seen in the SCUBA
images. The BIMA maps probably provide the best observational, high-resolution evidence
of structure in the inner disk that is associated with dust, as submm continuum observations
of dust in M 82 are not currently publicly available at a resolution higher than 6′′.7, that is
obtainable with SCUBA at 350 µm.
The first obvious difference between the SCUBA and high-resolution-BIMA maps is
that the dust emission peak indicated by a red contour is resolved into two peaks centered
about 9′′ apart in the CO maps. Of these two CO western peaks, the one closer to the
nucleus is co-spatial with the 450 µm (red contour) and slightly east of the 850 µm (red
contour) western peaks. The CO peak farthest from the nucleus is actually the brightest
of all the CO peaks and lies on the eastern edges of the submm 850 and 450 µm high-
brightness lobes, not co-incident with the brightest submm peaks. Distinct from submm
continuum, the lower level CO emission near the very western CO peak fans-out westerly
at position angle 85◦, diverting from the 72◦ -- position-angle of the submm continuum lobes
and disk along the major axis, as well as that of the CO within a 10′′ radius from the
galactic nucleus. These differences in the CO and dust features are evident in both the full
resolution and smoothed CO maps plotted in Figure 3, suggesting that although the dust and
CO generally appear mixed in this star-forming region, in fact there are differences in their
spatial distributions. The different concentrations are most probably locations of varying
gas-to-dust densities or star-formation environments. Consistent with the finding here of
different CO-dust concentrations, evidence of a star-formation history (e.g., de Grijs 2001)
and gas density (e.g., Petitpas & Wilson 2000) that clearly varies from the east to west of
the galaxy has been reported in M 82.
Higher resolution submm observations of dust in M 82 that should be possible with
the Smithsonian Submillimeter Array (SMA) and later Atacama Large Millimeter Array
(ALMA) will provide direct observational evidence to further test how different the CO and
dust emission trace each other at the small scales shown in the BIMA maps. Those future
observations will also test if dust emission has more complex morphology than has currently
been detected, as is suggested by the increasing structure that is seen in the SCUBA maps
going from low to high resolutions. Further discussion of the origin and structure of the
submm emission peaks is detailed in Section 5.
-- 8 --
3.2. Submm vs. Infrared and Optical Morphology
The spatial investigation of submm versus infrared and optical morphology in M 82
is important because complex optical morphology that is seen in this galaxy, with visual
extinction values (AV ) that range from about 3 to 25 (e.g., Alonso-Herrero et al. 2001), is
thought to result from obscuration of optical light by large, cold dust grains that are heated
by stars (among other things) and re-radiate in the infrared to submm wavelengths. Evidence
of star-formation history (e.g., de Grijs 2001), gas density (e.g., Petitpas & Wilson 2000),
and submm and CO emission peaks (this work, e.g. Section 3.1) that clearly varies from
the east to west of the galaxy strongly suggest that any associated dust lanes must vary not
only in their geometric structure but also in their heating mechanisms and composition.
3.2.1. Strong Optical-Obsuration Patches that Correspond with Submm and CO Peaks
The panels in Figure 4 show the B-band maps of M 82 obtained by the Hubble Legacy
Team (Mutchler et al. 2007), overlayed with the 450-µm continuum-emission contours shown
in Figures 1 and CO (1-0) interferometry data by Shen & Lo (1995). The overlays were made
by aligning the sky co-ordinates of the Hubble maps with those of the SCUBA and BIMA
ones. The respective resolutions in the Hubble, SCUBA, and BIMA maps are ∼ 0.09′′
∼ 8′′.5, 2′′.5. For best contrast, the optical intensities in this figure are inversely plotted,
and therefore the light patches are strong extinction features. The panels show that all the
submm and CO peaks of M 82 are spatially coincident with very strong optical-obscrution
patches seen within the intense star-forming region about 30′′ × 15′′ of the galactic nucleus
and 2.2 µm peak (Dietz et al. 1986, marked by a cross in the figures). The brightest submm
peak that is about 9′′ southwest of the nucleus and the CO peaks in this region stretch along
the same direction as prominent optical dust lanes that lie east-west along the galactic disk
major axis and co-spatial to these dust lanes out to about 15′′ from the galactic nucleus.
Interferometery maps of the central CO(1-0) peaks shown in the left panel of Figure 4 also
demonstrate the high-resolution, spatial co-incidence between the CO(1-0) peaks and very
dense optical-obscrution patches about the nucleus. The most western and brightest CO
peak also co-coincides with a dense optical-obscuration patch.
Within a radius of about 10′′ about the galactic nucleus, or about two arcseconds north
of the 2.2 µm peak Dietz et al. (1986), there is intense B−band optical emission. This region
is between the two submm peaks and thus has relatively low submm intensity or, if indeed
the submm emission is from cold dust re-radiation, low density or heating of cold dust. This
region also has mid-IR emission indicative of star-formation clusters (Lipscy & Plavchan
2004), though of less mid-IR brightness and lower mid-IR color temperature than the star-
-- 9 --
formation clusters southwest of this position, i.e. at the location of the southwestern submm
peak. The remarkable spatial co-incidence between the submm as well as high-resolution CO
peaks and very dense optical-obscuration patches about the nucleus, and the co-incidence
of intense B−band optical emission with location of relatively low submm emission or cold
dust column density, support suggestions in this paper that the submm emission in the inner
disk of M 82 originates from re-radiation of dust concentrations or clouds of physically dif-
ferent star-formation environments (e.g., Achtermann & Lacy 1995; Forster Schreiber 2000;
de Grijs 2001), rather than the commonly assumed interpretation of a dusty torus about the
nucleus (see Section 5.1 below).
In color images of M 82 (see, e.g., de Grijs 2001; Westmoquette et al. 2007), the opti-
cal emission associated with submm peaks and high-brightness, diffuse re-radiation in the
inner disk of this galaxy has a blue-brownish hue, strongly suggesting that hot, young,
blue stars are the main heating source for the dust in this inner region. The young stars
also produce infrared emission and emission lines associated with intense star-formation this
galaxy (e.g., Achtermann & Lacy 1995; Alonso-Herrero et al. 2001; Lipscy & Plavchan 2004;
Westmoquette et al. 2007).
3.2.2. Strong Optical-Obsuration Patches with No Corresponding Submm and CO Peaks
Although all the submm and CO peaks correspond to optical features in the inner disk
of M 82, as described above, the contrary is not true; i.e, many prominent dust, optical
filaments, clouds, and lanes have no clear corresponding submm emission counterparts. In
particular, the submm emission is basically smooth at the locations of 1) north-south fila-
ments that run below, through, and flare above the southwestern submm peak that is about
8′′ from the galactic nucleus; 2) east-west dark lanes that run west along the galactic major
axis and continue through to about 35′′ west of the galactic nucleas (or 25′′ west of the
southwestern submm peak); 3) huge, dramatic complex of optically-obscuring clouds, fila-
ments and lanes that extends the entire minor axis of M 82's disk and covers an area greater
than a diameter of 25′′ south of the disk just east of the eastern submm peak that is about
10′′ from the galactic nucleus; and 4) light, optically-obscuring clouds and filaments in a
'low'-extinction region known as a starburst-remnant (e.g., de Grijs 2001) that lies about
30′′ to 60′′ northeast of the galactic nucleus. In another region of low-optical extinction and
radiation that is about 60′′ to 120′′ northeast of the galactic nucleus, submm emission has
currently not been detected where very light optically-obscuring filaments are evident.
The lack of correspondence between the very dark optical clouds, filaments, and lanes
with bright submm emission suggests that these optical features are due to obscuration by
-- 10 --
cool dust grains that are on the near side of the galaxy and at large distances from the nuclear
region, where they are heated by a very dilute stellar radiation field. These foreground dust
clouds evidently have enough column densities to obscure optical light in the line of sight;
however, they only re-emit very low-level emission that shows no striking features in the
current submm maps. That the complex optical morphology seen in Figure 4 is primarily
due to foreground dust is supported by the fact that the obscuration is more dramatic in
the shorter wavelength B−band than longer I −band images (L. L. Leeuw et al. 2009 in
preparation). Obscuring dust is expected to be optically thicker at the shorter wavelength
and therefore cause more optical extinction and thus appear more prominently at the shorter
wavebands.
One explanation for the submm low-level continuum having a relatively smooth mor-
phology is that, because this emission is optically thin, the detected radiation at a particular
submm wavelength represents the total emission from the entire galactic column of dust in
the line of sight. This is different from the optical morphology of dust because the dust is
typically seen obscuring stellar light, and therefore only the dust in certain spatial strati-
fications (usually the foreground) of the line-of-sight columns is observed. In other words,
morphology due to spatial-depth or stratification of similar dust grains that are heated by a
dilute radiation field is most often more obviously seen in optical obscuration than in submm
emission. These morphological effects will of course depend on the sensitivities and resolu-
tions of the instruments used. In a low optical extinction and radiation region about 60′′
to 120′′ northeast of the galactic nucleus, for example, the lack of any submm detection to
date may simply because current continuum instruments have not been sensitive enough to
easily detect low-level, dust re-radition that my correspondence to low-level extinction and
stellar heating in those regions. The relatively larger SCUBA beam could also beam-smear
and thus erase small features detected in the higher resolution optical maps.
3.2.3. The Outflowing Wind
An alternative explaination for the submm morphology of M 82 being smooth as op-
posed to distrupted like the optical morphology depicted in Figure 4 is that the submm
low-level emission is primarily due to dust entrained in outflowing gas and physically dif-
ferent from the optical extinction features that don't have any currently detected submm
counterparts. Figure 5 shows the low- and high-brightness features of the Hα maps obtained
by the Hubble Legacy Team (Mutchler et al. 2007), overlayed with SCUBA 850 and 450-µm
continuum-emission contours as shown in Figures 1. For best contrast, the optical intensities
are inversely plotted and the light patches (e.g. across the center of the image in the right
-- 11 --
panel) are foreground extinction features. The Hα emission is plotted saturated to high-light
the large-scale and base of the outflowing Hα wind, so not all the obscuration patches that
criss-cross optical maps of M 82 are depicted here. The overlay in the left panel demonstrates
the spatial co-incidence between the large-scale low-brightness Hα emission and 850-µm con-
tinuum north and south of the disk, while the right panel depicts the origin of the large-scale
Hα emission in the intense star-formation inner disk of M 82 about the submm emission
peaks (especially near the southwestern peak), where a ∼ 130 pc expanding "superbubble"
has been discovered in CO (Weiss et al. 1999) and ionized gas (Wills et al. 1999).
Recent large-scale, high-resolution OVRO observations by Walter et al. (2002) detected,
resolved molecular CO(1-0) streamers in and below M 82's disk that have different kinemat-
ical signatures to its outflowing gas. Some of the streamers are well correlated with optical
obscuration features and form the basis of some prominent tidal H I features (Yun et al.
1993) that are thought to provide evidence that the gas within the optical disk of M 82 is
disrupted by the interaction of M 82 with M 81 and likely triggers the starburst activity in
M 82's center (Walter et al. 2002). The detection of resolved mid-IR to submm emission that
corresponds to the optical and gas streamers and is perhaps a physically separate component
to the outflowing gas and dust in M 82 should be possible with sensitive and high-resolution
mid-IR to submm imaging instruments using Spitzer, ALMA and the SMA. New Spitzer ob-
servations reported by Engelbracht et al. (2006) did indeed detect extended mid-IR emission
not only in the ouflow of M 82 but also in its halo. The extended halo mid-IR emission could
be from material ejected into the halo by the outflowing wind or from the interaction of M 82
with M 81 (Engelbracht et al. 2006) . Future observations with these sensitive instruments
and their detailed data analysis have the potential to 1) directly uncover the disruption
of cold (and warm) dust distribution by the interaction of M 82 with M 81, 2) decompose
submm dust emission in the disk, outflows, and streamers of M 82 and better constrain the
properties of cold dust in these seperate components (see Section 5.3), and 3) elucidate the
role or consequence of the dust in the interaction of M 81 and M 82 (e.g., Yun et al. 1993)
and any connected triggering and evolution of the star-forming in M 82 (cf. Walter et al.
2002), and the re-processing of galactic dust in general.
4. Possible CO Contamination of the Submm Continuum?
It is worth noting that the dust morphology that is mapped in the submm continuum
from M 82 may be substantially enhanced by CO emission from this galaxy. A recent flux
comparison between CO(3-2) emission and 850 µm continuum in M 82 showed that CO
makes a 47% (i.e. high) contribution to the integrated continuum in this SCUBA band
-- 12 --
(Seaquist & Clark 2001). From analyzing collated CO(4-3) observations together with those
of CO at lower transitions, Guesten et al. (1993) concluded that the line strengths in M 82
increased as one went to higher transitions, indicating that the higher transitions must
provide significant cooling in the galaxy. All SCUBA bands have roughly the same widths,
i.e. 30 GHz, and therefore the CO contribution to the higher frequency continuum would
be expected to be equally or more significant than that reported for the 850 µm band by
Seaquist & Clark (2001). However, because the submm continuum in M 82 has a thermal
spectrum (see Section 5.2), and therefore the submm fluxes increase with frequency, the
CO contribution to the higher frequency continuum may be less than the estimates for
the 850 µm band. For SED analysis in this paper (e.g., Section 5.2), the CO percentage
contribution to the measured flux is assumed to be the same across the SCUBA bands and
no correction for it is made in the presented data.
Observations using a new high frequency, Fabry-Perot spectrometer on the JCMT have
lead to clear detections of the high transition CO(7-6) in M 82 and NGC 253 (Bradford et al.
1999), the first such detection in any extragalactic sources. The analysis by Guesten et al.
(1993) and the detections of CO(7-6), whose transition line lies in the 450 µm filter bandpass,
suggest that other higher transition lines, such as the 13CO(8-7) line that lies at the centre
of the SCUBA 350 µm band, could be very strong in M 82, supporting the proposition above
that the high frequency SCUBA images may have significant contribution from CO. In this
light, the morphology seen in the SCUBA images is a direct probe of the galactic cooling
and the general interactions of active star formation and the ISM in M 82.
Although it is not obvious if the CO contribution to the higher frequencies of SCUBA
will be less than or as significant as estimates by Seaquist & Clark (2001), it is clear that
the CO contamination to higher-frequency continuum warrants investigation. Future work
on this galaxy will attempt to acquire the data of the CO lines in the 450 and 350 µm
bands and make a quantitative comparison of these data in order to determine the possible
contributions of CO to the high-frequency-SCUBA data. Such work is important (among
other things) in the determination and interpretation of submm SED and thus the nature of
dust emission in M 82 (see Section 5.2 below).
5. Origin of Submm Continuum and Spectral Energy Distribution
5.1. Source and Structure of the Submm Continuum in the Inner Disk
The radiation from M 82 at the radio (e.g., Seaquist & Odegard 1991; Wills et al. 1999),
mm (e.g., Hughes et al. 1990) and submm-to-infrared (e.g., Telesco et al. 1991; Hughes et al.
-- 13 --
1994; Alton et al. 1999) wavelengths is respectively dominated by synchrotron emission from
supernovae, free-free emission from ionized gas, and thermal re-radiation from dust heated
by young stars. In this light, the double peaks seen in the mm to infrared continuum have
commonly been interpreted as due to emission from the edges of an inclined, dusty-molecular
torus, that -- as a result of their geometry on the plane of the sky and optically-thin nature of
radiation -- have relatively high optical-depths in the line of sight (e.g., Hughes et al. 1994).
In Figure 1, and other mm-to-infrared maps of similar or worse resolution (such as those
from IRAS), the double peaks are not resolved and appear as a single, elongated lobe that
is brightest in the southwest. Like the galactic disk, the lobe (or peaks when resolved-out)
has a position angle of roughly 72◦ and -- in the tori interpretation (e.g., Shen & Lo 1995) --
an inclination of ∼ 10◦.
The peaks of emission seen in the mm to infrared have alternatively been interpreted sim-
ply as dust and molecular concentrations along the galactic disk, perhaps in a bar structure
(e.g., Neininger et al. 1998; Westmoquette et al. 2007) that may have an expanding "super-
bubble" of gas centered at supernova remnant 41.9+58 (e.g., Weiss et al. 1999; Wills et al.
1999). This interpretation is supported by at least three reasons. First, the east-west, dou-
ble peaks have now been seen in maps of both optically thin and thick CO emission (e.g.,
Petitpas & Wilson 2000). This is reasonable if the emission is from a structure that consti-
tutes concentrations or clouds of dust but is contrary to what is expected if the emission
is from a structure with tori geometry that, like the galactic disk, is thought to be highly
inclined (e.g., Shen & Lo 1995). For optically thin radiation, it will be possible to detect
emission from the inner parts of the imaged structure, and either dust clouds or indeed edges
of an inclined torus would manifest as regions of relatively higher optical-depth or brighter
optically-thin emission in the line of sight (e.g., Hughes et al. 1994). However, as also noted
by Neininger et al. (1998) and Petitpas & Wilson (2000), for optically thick radiation, it
will be possible to directly detect emission only from the foreground surface of the imaged
structure. In that case, the dust clouds will be seen as two emission peaks, while the torus
(or bar) will manifest as an elongated, barlike emission of roughly the same optical depth or
optically-thick brightness.
Second, the two main peaks seen in maps of similar resolution as the 450 and 350 µm
images in Figures 1 are not symmetric. In maps of better resolution and sensitivity (e.g.,
Shen & Lo 1995, see the CO contours in Figure 3), the peak west of the nucleus has a
morphology clearly different from the eastern peak and can be resolved into two or three
structures. High-resolution maps obtained with the Very Large Array (VLA) by Wills et al.
(1999) showed that the western peak is associated with locations of supernova explosions
of higher intensity and earlier evolutionary stage than the eastern peak and confirmed the
discovery by Weiss et al. (1999) of an expanding "superbubble" that is centered near the
-- 14 --
location of M 82's brightest supernova remnant, 41.9+58, and the submm western peak.
The varying supernova intensities and ages across the disk of M 82 is supported by high
resolution HST imaging of stellar clusters that indicates that the regions near the eastern
and western submm peaks have different star-formation histories (e.g., de Grijs 2001). In this
light, the submm peaks indicate concentrations of dust environments associated with different
locations of varying supernova and star-formation activity, not the commonly assumed dust
torus.
Third, observations of line ratio gradients indicate that the average temperature across
the lobe increases from the northeast to southwest, while the density increases in the op-
posite direction (e.g., Petitpas & Wilson 2000). Further evidence of the higher temperature
or, at least, column density in the southwest is seen in the lope-sided 850 and 750 µm lobe
and the double-peak 450 and 350 µm lobes in which the southwest parts are the brighter. A
torus that probably houses and is heated by an active galactic nucleus (e.g., Muxlow et al.
1994), would be expected to have a temperature that decreases from its inner to outer walls.
One explanation for the higher temperature and lower density in the the southwest is linked
with star-formation activity that both heats and depletes the interstellar medium (ISM) at
this location (Wills et al. 1999), or simply western and eastern regions of two different star-
formation physical environments (e.g., Achtermann & Lacy 1995; Forster Schreiber 2000;
de Grijs 2001; Lipscy & Plavchan 2004). Evidence in this paper in terms of clearly asym-
metric submm emission seen in almost all the presented intensity maps seems to disfavor the
commonly assumed interpretation of a dusty torus in M 82.
5.2. Fluxes and Spectral Energy Distribution Analysis
Section 4 above raised the possibility that CO emission may contaminate the continuum
of M 82 observed in SCUBA filters. For one, significantly different contributions in the
SCUBA bands imply different corrections to the measured fluxes and would affect the SED
analysis using those fluxes.
If the differences are significant the SED computation and
analysis should in theory only be determined after correcting or accounting for the CO
contamination.
It is currently not obvious if the contamination to the higher-frequency
continuum will be less or higher than the 47% estimated to the SCUBA-850 µm band by
Seaquist & Clark (2001); even though it is clear that it may also be important (see Section 4).
While relevant CO data need to be acquired in the future to make a quantitative calculation
of the relative contributions in the SCUBA bands, for the practical determination of the
SEDs of M 82 in this paper, it is assumed that the CO contribution in the SCUBA bands is
the same.
-- 15 --
Table 2: SCUBA continuum fluxes for specific locations in M 82 (The listed measured flux densities
include roughly 47% possible contribution from C0 as discussed in the text.)
Locations parallel to M 82's
disk position angle of 72◦
@Peak Flux
30′′ × 15′′, about nucleus
70′′ × 40′′, about nucleus
30′′ × 15′′, 15′′N of nucleus
30′′ × 15′′, 15′′S of nucleus
30′′ × 15′′, 30′′N of nucleus
30′′ × 15′′, 30′′S of nucleus
Flux (Jy)
@350µm
18.6 ± 5.6
49.3 ± 13.7
63.4 ± 18.9
15.9 ± 4.8
12.8 ± 3.8
6.2 ± 1.9
3.1 ± 0.9
Flux (Jy) Flux (Jy) Flux (Jy)
@450µm @750µm @850µm
13.0 ± 2.6
1.2 ± 0.1
2.3 ± 0.2
28.8 ± 5.6
3.8 ± 0.4
35.9 ± 7.0
7.6 ± 1.5
0.6 ± 0.1
0.6 ± 0.1
8.1 ± 1.6
0.2 ± 0.02
2.8 ± 0.5
1.8 ± 0.4
0.1 ± 0.1
1.8 ± 0.4
3.2 ± 0.5
7.7 ± 1.5
1.0 ± 0.2
0.7 ± 0.1
0.3 ± 0.1
0.2 ± 0.1
Table 2 shows the submm fluxes measured at specific locations across M 82 using
SCUBA. The listed errors include calibtation uncertainties. The submm dominant emit-
ting region in M 82 is within 30′′ × 15′′ about the nucleus and is associated with the most
intense star-formation in the galaxy. Assuming the primary source of the submm emission is
dust re-radiation, the measured fluxes from this galaxy were fitted with the following thermal
function:
Fν = ΩBν(T )[1 − exp(−(
λo
λ
)β)],
(1)
where Ω is the solid angle for the emitting region, Bν(T ) the Planck function at temperature
T , λo the wavelength at which the optical depth is unity (λo = 7.8 µm, Hughes et al. (1994)),
and β the emissivity index of the grains. Due to the limited frequency sampling in the data,
the temperature T and emissivity index β were the only parameters that were statistically
determined in Equation 1.
Table 3: SCUBA derived dust emission properties for locations in M 82
Locations parallel to M 82's
disk position angle of 72◦
30′′ × 15′′, about nucleus
30′′ × 15′′, 15′′N of nucleus
30′′ × 15′′, 15′′S of nucleus
70′′ × 40′′, about nucleus
T
(K)
36 ± 14
31 ± 3
31 ± 15
27 ± 5
β
Ω
(sr)
2.0 ± 0.1
2.2 ± 0.1
2.3 ± 0.2
1.8 ± 0.1
1.66e − 8
1.66e − 8
1.66e − 8
9.05e − 8
Table 3 lists the derived dust T and β, as well as the Ω associated with the specified
emitting regions in M 82. For the listed rectangular locations, the Ω is determined from
the longest side of the rectangle. The average derived T is ∼ 31 K, with a minimum and
maximum of ∼ 27 K and ∼ 36 K respectively. The highest T 's are from regions of the highest
-- 16 --
surface brightnesses, presumably corresponding to regions of intense star-formation, while
the coldest T 's are at regions farthest from the nucleus. The β values range from 1.8 to
2.3 and are highest at regions farthest from the galactic nucleus. The higher β values are
associated with larger and colder dust grains; therefore, the spatial SED analysis here points
to colder grains more prevelant with increasing distance from the galactic center of M 82,
particularly along the minor axis of the galactic disc (see Section 5.3 below).
If the CO contamination in the SCUBA bands were not constant for M 82, as assumed in
this paper, an SED with SCUBA fluxes may have looked different than the result above and
possibly required a different interpretation. For example, a CO contamination that increased
with frequency, as suggested from an analysis of lower-transition CO and CO(4-3) data of
M 82 by Guesten et al. (1993) (see Section 4), would mean that SCUBA fluxes uncorrected
for CO contamination lead to the SED indicating lower dust temperatures and/or higher
emissivity indices than was actually the case. A proper correction for any CO contamination
is therefore needed before a more definitive SCUBA SED analysis can be conducted for
M 82 and is worth attempting in the future, when CO data in all the SCUBA wavebands
(especially the 350 and 450µm ones) are available.
5.3.
Implications of the SCUBA data and SEDs on the Outflow of Cold Dust
In has been noted that the continuum morphology in the inner-halo of M 82 has a
general north-south asymmetry, that is consistent with the north-south asymmetric X-ray
and Hα winds (e.g., Watson et al. 1984, and see Section 3.2.3) and the associated UV, optical,
molecular and indeed infrared to mm structures that have been reported in M 82 (e.g.,
Seaquist & Clark 2001). Therefore, a simple interpretation of the asymmetric morphology
of the submm continuum in the halo of M 82 and is that it is a manifestation of an outflow
of dust from the inner disk to the halo.
Seaquist & Odegard (1991) presented one of the earlier, extensive evidence of the disk-
to-halo outflows from spectral index distribution computations using radio continuum maps
at several wavelengths between 0.33 and 4.9 GHz (90 and 6 cm). They found spectral indices
between -- 0.3 and -- 0.6 in the inner-disk region, steepening to about -- 1.0 at a radius of about
1 kpc along the minor axis, and concluded that these were from relativistic synchrotron-
emitting electrons that were being scattered against infrared photons emitted in the inner-
disk region of M 82. Recently, Wills et al. (1999) used high resolution VLA continuum data
between 1.4 and 5 GHz and computed spectral indices from −0.6 to −0.8 about 20′′ north
of the disk, in localized nuclear sites of the outlows that they call 'chimneys'. These values
are consistent with the results by Seaquist & Odegard (1991) in the same wavebands, and
-- 17 --
Wills et al. (1999) also interpreted them as indicating synchrotron emission from relativistic
electrons entrained in the wind.
A thermal component in the filaments has previously been suggested based on 'tentacles'
observed in Ne II maps of M 82 (Achtermann & Lacy 1995), which presumably are also a
manifestation of the outflow phenomenon. In comparative SED analysis of 30′′ × 15′′ regions
centered in galactic nucleus and two others 15′′ north and south of it, the regions north
and south had respectively cooler temperatures and higher emmisivity indices than the
central region. This spatial SED analysis is consistent with the submm emission coming
from a thermal source with a temperature decrease and emmisitivity index increase along
the minor axes of the disk of M 82. The change of the dust properties along the minor axes
has a similar direction to the radio spectral index gradient shown by Seaquist & Odegard
(1991) and Wills et al. (1999) and is consistent with the north-south asymmetric Hα winds
(e.g., Shopbell & Bland-Hawthorn 1998) of M 82 that has been shown to be co-spatial with
the submm morphology (see Figure 5 and Section 3.2.3). In this light, the submm continuum
maps (and SED changes along the minor axes) indicate an outflow of dust grains that are
ejected from the inner disk by, or entrained in, the starburst winds.
One explanation for the outflow being asymmetric was given by Shopbell & Bland-Hawthorn
(1998), who suggested that if the star-forming disk is slightly shifted up from the galactic
disk, this would imply that there is less covering material in the north and would make
collimation difficult, resulting in an immediate blow-out of material in the north. Detections
confirming that collimation is better to the south of M 82 have been made of large-scale
emission extending to 1.5 kpc, and more extended in the south, not only in optical line maps
(e.g., Devine & Bally 1999), but also in CO(2-1) and CO(3-2) respectively by Thuma et al.
(2000) and Seaquist & Clark (2001). Another valuable result of the co-addition of SCUBA
archive data in this paper is that the most sensitive submm maps that are displayed in
Figure 1 show submm extended emission that is associated with the outflows on scales that
for the first time match the 1.5 kpc CO detections noted above.
Recently reported radio 'chimneys' (e.g., Wills et al. 1999), that are about 20′′ north
of the disk and hypothesised to signify local blow-outs of material by supernova-driven
winds, are not obvious in the SCUBA maps. However, prominent SiO features associ-
ated with the localized radio outflows have now been detected in mm-heterodyne obser-
vations obtained with the Institut de Radio Astronomie Millimetrique interferometer (e.g.,
Garc´ıa-Burillo et al. 2001). These authors explain the SiO detections in a framework of
shocked chemistry at the sites of the gas ejections from the starburst disk. For the moment
it appears that the shocked gas has proved to be a better probe of the localized outlows than
the direct observations of dust emission in the mm to submm continuum.
-- 18 --
Localized sites of the outlows or 'emission spurs', although long sought-after and some-
times reported in mm to submm continuum and CO maps (e.g., Hughes et al. 1994; Shen & Lo
1995; Leeuw et al. 1999; Alton et al. 1999), have not been reliably reproduced in the different
mm to submm observations. All the continuum maps in this paper also have some low-level
'spurs', but almost none are reproduced at exactly the same locations and extend to the same
degrees between any two different observations. This would suggest that the spurs might
be artefacts in the maps. However, if the mm and submm spurs are emission from dust
outflows (and filaments, clouds or lanes) that are of different optical depths or compositions
and re-radiate low-level emission in relatively narrow wavebands, the submm spurs may not
be reproduced in maps at certain wavelengths and sensitivities.
It was noted in Section 3.2 that dust in M 82 is most probably within components of
physically varying locations and origin or simply at different spatial-depths or stratifica-
tions. The recent high-resolution, interferometry observations of M 82's molecular gas by
Walter et al. (2002) discovered CO(1-0) 'streamers' and decoupled previously observed CO
outflows (e.g., Nakai et al. 1987; Seaquist & Clark 2001) from the streamers (e.g., Yun et al.
1993), clarifying the spatial distribution and origin of the molecular gas in this galaxy. Sim-
ilar, future observations with recently commissioned sensitive and high-resolution submm
imaging instruments such as Spitzer and the SMA should provide tighter constraints on the
spatial and optical depth properties of dust, test the reality of the reported submm spurs,
and possibly detect dust re-radiation streamers, clarifying the implications or associations
of all these features to dust outflows and recycling in M 82 (c.f. Section 3.2).
6. Summary of Results and Future Work on M 82
SCUBA 350, 450, 750 and 850 µm imaging observations have been presented of the
dust-laden, star-forming inner disk and large-scale, low-level emission that is associated with
the outflows in the halo of M 82. The displayed maps include co-added data that were
mined from the SCUBA Data Archive, resulting in the deepest submm continuum maps of
M 82. The 850 µm morphology has a single emission peak that is centered about 9′′ west
of the galactic nucleus, while the 450 and 350 µm maps have two emission peaks centered
about 10′′ and 6′′ respectively east and west of the nucleus along the galactic disk, similar to
previous continuum observations at mm, submm and mid-infrared wavelengths, as well as
to CO line transitions and Hα observations of comparable resolution (see Section 3). In the
750 µm image, the single emission peak seen in the 850µm image (see Figure 1 is predictably
beginning to be resolved-out into the double peaks seen in the 450 and 350µm images. Low-
level emission is detected out to 1.5 kpc for the first time in the 850 µm continuum of this
-- 19 --
galaxy, i.e. at least ∼ 160 pc radius farther-out than other recent studies.
The deep maps were used in a detailed morphological study of the disk and large-
scale detections, including a comparative analysis of submm to optical morphology (see
Section 3.2). The overall, extended submm morphology of M 82 generally resembles the
optical picture in that the disk emission has an apparently elliptical shape whose major axis
is clearly aligned with that of the galactic disk at position angle ∼ 72◦. However, the submm
morphology is much smoother than the optical picture, and some prominent dust cloud and
filamentary lanes that are seen in the optical are not obvious in the submm continuum. One
simple explanation for the submm versus optical correspondence (or lack of it) is the different
resolutions and sensitivies of the presented observations. If the optical features emit submm
emission, it is possible the emission is at a level lower than the current mapped SCUBA
sensitivities or smeared and thus erased by the relatively larger SCUBA beam. This can
be varified by future higher resolution and more sensitive submm observations that should
become possible with ALMA.
A comparative analysis was further conducted of submm to high-resolution CO (1-0)
morphology (see Section 3.1). Some resolved peaks in the CO maps could be associated
with unresolved features in the submm maps. However, there are differences that show the
CO and dust emission do not trace each other in a very simple way and suggest that although
the dust and CO generally appear mixed in the central star-forming region, in fact there
are differences in their spatial distributions. The different concentrations are most probably
locations of varying gas-to-dust densities or star-formation environments, that have been
reported in M 82 (e.g., de Grijs 2001; Petitpas & Wilson 2000). This can be varified by
future higher resolution observations of dust in M 82 that should be possible with the SMA
and later ALMA.
The SCUBA maps were also used in a computation of the first submm spatial SED
analysis of locations within and outside the central star-forming region of M 82 (see Sec-
tion 5.2) and in the discussion of the origin and structure of submm maps (see Section 5).
In particular, (a) the commonly assumed interpretation that the double emission peaks that
were seen in the mm to infrared continuum are due to emission from the edges of an in-
clined, dusty-molecular torus was challenged (see Section 5.1), (b) an analytical review of
CO results was undertaken to assess if CO emission might significantly contaminate the
continuum observed in SCUBA filters (see Section 4), and (c) a morphological comparison
was conducted to check whether the localized outflows that were reported in radio and SiO
maps respectively by Wills et al. (1999) and Garc´ıa-Burillo et al. (2001) could be seen in the
SCUBA maps (see Section 5.3).
Evidence in this paper in terms of clearly asymmetric inner-disk submm emission seen
-- 20 --
in the presented intensity seems to disfavor the commonly assumed interpretation of a dusty
torus in M 82. Arguments were presented to explane the inner-disk submm maps of M 82
in the context of emission from a rather complex distribution of dust concentrations that
are in regions of different star-formation environments, as has been reported from various
studies using data at other wavelengths (e.g., Achtermann & Lacy 1995; Forster Schreiber
2000; de Grijs 2001).
It is not obvious if the CO contribution to the higher frequencies of SCUBA will be less
than or higher than the 47% estimated to the SCUBA-850 µm band by Seaquist & Clark
(2001). However, it is clear that the CO contamination to higher-frequency continuum may
also be significant and warrants detailed investigation. Future work on this galaxy will
attempt to acquire the data of the CO lines in the 450 and 350 µm bands and make a
quantitative comparison of these data in order to determine the possible contributions of
CO to the high-frequency-SCUBA data.
The overall submm low-level morphology has a general north-south asymmetry that is
similar to the Hα winds and CO and X-ray outflows that have been detected in M 82 (e.g.,
Shopbell & Bland-Hawthorn 1998). The submm spatial SED analysis also shows thermal
properties that change along the minor axis of the galaxy disk, similar to the radio spectral
index gradient by Seaquist & Odegard (1991) and Wills et al. (1999) and consistent with
the north-south asymmetric, large-scale X-ray and Hα winds. Therefore, the current re-
sults support the simple interpretation (e.g., Leeuw et al. 1999; Alton et al. 1999) that the
asymmetric morphology in the submm maps is a manifestation of corresponding outflows of
dust grains from the galactic disk into the halo. As noted above, this work has presented
low-level 850 µm continuum emission out to 1.5 kpc for the first time in this galaxy, i.e. at
least ∼ 160 pc radius farther-out than other recent studies.
7. Acknowledgments
Lerothodi L. Leeuw acknowledges a NASA Postdoctoral Fellowship that supported the
final write-up of this work at NASA Ames Research Center.
Achtermann, J. M. & Lacy, J. H. 1995, ApJ, 439, 163
REFERENCES
Alonso-Herrero, A., Rieke, M. J., Rieke, G. H., & Kelly, D. M. 2001, Ap&SS, 276, 1109
-- 21 --
Alton, P. B., Davies, J. I., & Bianchi, S. 1999, A&A, 343, 51
Archibald, E. N., Jenness, T., Holland, W. S., Coulson, I. M., Jessop, N. E., Stevens, J. A.,
Robson, E. I., Tilanus, R. P. J., Duncan, W. D., & Lightfoot, J. F. 2002, MNRAS,
336, 1
Bradford, C. M., Stacey, G. J., Nikola, T., Swain, M. R., Bolatto, A. D., Jackson, J. M.,
Savage, M. L., & Davidson, J. A. 1999, BAAS 31, 1477.
Chevalier, R. A. & Clegg, A. W. 1985, Nature, 317, 44
de Grijs, R. 2001, Astronomy and Geophysics, 42, 12
Devine, D. & Bally, J. 1999, ApJ, 510, 197
Dietz, R. D., Smith, J., Hackwell, J. A., Gehrz, R. D., & Grasdalen, G. L. 1986, AJ, 91, 758
Engelbracht, C. W., Kundurthy, P., Gordon, K. D., Rieke, G. H., Kennicutt, R. C., Smith,
J.-D. T., Regan, M. W., Makovoz, D., Sosey, M., Draine, B. T., Helou, G., Armus,
L., Calzetti, D., Meyer, M., Bendo, G. J., Walter, F., Hollenbach, D., Cannon, J. M.,
Murphy, E. J., Dale, D. A., Buckalew, B. A., & Sheth, K. 2006, ApJL, 642, L127
Forster Schreiber, N. M. 2000, New Astronomy Review, 44, 263
Freedman, W. L., Hughes, S. M., Madore, B. F., Mould, J. R., Lee, M. G., Stetson, P.,
Kennicutt, R. C., Turner, A., Ferrarese, L., Ford, H., Graham, J. A., Hill, R., Hoessel,
J. G., Huchra, J., & Illingworth, G. D. 1994, ApJ, 427, 628
Garc´ıa-Burillo, S., Mart´ın-Pintado, J., Fuente, A., & Neri, R. 2001, ApJL, 563, L27
Guesten, R., Serabyn, E., Kasemann, C., Schinckel, A., Schneider, G., Schulz, A., & Young,
K. 1993, ApJ, 402, 537
Heckman, T. M. 2003, in "Galaxy Evolution: Theory & Observations", eds. V. Avila-Reese,
C. Firmani, C. S. Frenk, & C. Allen, Revista Mexicana de Astronomia y Astrofisica
Conference Series, Mexico, Vol. 17, 47
Heckman, T. M., Armus, L., & Miley, G. K. 1990, ApJS, 74, 833
Hoopes, C. G., Heckman, T. M., Strickland, D. K., Seibert, M., Madore, B. F., Rich, R. M.,
Bianchi, L., Gil de Paz, A., Burgarella, D., Thilker, D. A., Friedman, P. G., Barlow,
T. A., Byun, Y.-I., Donas, J., Forster, K., Jelinsky, P. N., Lee, Y.-W., Malina, R. F.,
Martin, D. C., Milliard, B., Morrissey, P. F., Neff, S. G., Schiminovich, D., Siegmund,
-- 22 --
O. H. W., Small, T., Szalay, A. S., Welsh, B. Y., & Wyder, T. K. 2005, ApJL, 619,
L99
Hughes, D. H., Gear, W. K., & Robson, E. I. 1994, MNRAS, 270, 641
Hughes, D. H., Robson, E. I., & Gear, W. K. 1990, MNRAS, 244, 759
Jenness, T., Lightfoot, J. F., & Holland, W. S. 1998, Proc. SPIE, 3357, 548
Kuno, N. & Matsuo, H. 1997, PASJ, 49, 265
Leeuw, L. L., Robson, E. I., & Hughes, D. H. 1999, in "The Stellar Content of Local Group
Galaxies", eds. P.A. Whitelock & R. Cannon, ASP Series, San Francisco, CA, USA,
Vol. 192, 330
Lipscy, S. J. & Plavchan, P. 2004, ApJ, 603, 82
Mutchler, M., Bond, H. E., Christian, C. A., Frattare, L. M., Hamilton, F., Januszewski,
W., Levay, Z. G., Mountain, M., Noll, K. S., Royle, P., Gallagher, J. S., & Puxley,
P. 2007, PASP, 119, 1
Muxlow, T. W. B., Pedlar, A., Wilkinson, P. N., Axon, D. J., Sanders, E. M., & de Bruyn,
A. G. 1994, MNRAS, 266, 455
Nakai, N., Hayashi, M., Handa, T., Sofue, Y., Hasegawa, T., & Sasaki, M. 1987, PASJ, 39,
685
Neininger, N., Guelin, M., Klein, U., Garcia-Burillo, S., & Wielebinski, R. 1998, A&A, 339,
737
Ohyama, Y., Taniguchi, Y., Iye, M., Yoshida, M., Sekiguchi, K., Takata, T., Saito, Y.,
Kawabata, K. S., Kashikawa, N., Aoki, K., Sasaki, T., Kosugi, G., Okita, K., Shimizu,
Y., Inata, M., Ebizuka, N., Ozawa, T., Yadoumaru, Y., Taguchi, H., & Asai, R. 2002,
PASJ, 54, 891
Petitpas, G. R. & Wilson, C. D. 2000, ApJL, 538, L117
Seaquist, E. R. & Clark, J. 2001, ApJ, 552, 133
Seaquist, E. R. & Odegard, N. 1991, ApJ, 369, 320
Shen, J. & Lo, K. Y. 1995, ApJL, 445, L99
Shopbell, P. L. & Bland-Hawthorn, J. 1998, ApJ, 493, 129
-- 23 --
Telesco, C. M., Joy, M., Dietz, K., Decher, R., & Campins, H. 1991, ApJ, 369, 135
Thuma, G., Neininger, N., Klein, U., & Wielebinski, R. 2000, A&A, 358, 65
Veilleux, S., Cecil, G., & Bland-Hawthorn, J. 2005, ARA&A, 43, 769
Walter, F., Weiss, A., & Scoville, N. 2002, ApJL, 580, L21
Watson, M. G., Stanger, V., & Griffiths, R. E. 1984, ApJ, 286, 144
Weiss, A., Walter, F., Neininger, N., & Klein, U. 1999, A&A, 345, L23
Westmoquette, M. S., Smith, L. J., Gallagher, III, J. S., O'Connell, R. W., Rosario, D. J.,
& de Grijs, R. 2007, ApJ, 671, 358
Wills, K. A., Redman, M. P., Muxlow, T. W. B., & Pedlar, A. 1999, MNRAS, 309, 395
Yun, M. S., Ho, P. T. P., & Lo, K. Y. 1993, ApJL, 411, L17
-- . 1994, Nature, 372, 530
This preprint was prepared with the AAS LATEX macros v5.2.
-- 24 --
Fig. 1. -- Deep maps of the 850 (top left - a), 750 (top right - b), 450 (bottom left - c),
and 350 µm (bottom right - d) continuum emission centered at the near-infrared nucleus of
M 82. The 850, 750, 450, and 350 µm contours on the respective maps are [16, 32, 48, 64,
96, 128, 200, 300, 400, 600, 800, 1000, 1175, and 1350], [200, 400, 600, 800, 1000, 1300,
1600, and 1900], [450, 900, 1450, 2000, 3000, 4000, 5000, 6000, and 6800], and [1300, 1950,
3250, 5200, 7000, 8500, and 10000] mJy/beam; the respectively rms uncertainties are ∼ 8,
100, 225, and 650 mJy/beam, and the resolutions are ∼ 14′′.5, 11′′.4, 8′′.5, and 6′′.7. The
white circle in each map indicates the approximate size of the SCUBA beam at the plotted
wavelength. The keys are grayscale-coded intensities in Jy/beam, and the X and Y axes are
J2000 coordinates.
-- 25 --
Fig. 2. -- Deep maps of the 850 µm continuum emission (left - a) and integrated CO(2-1) line
intensity (right - b) of M 82 respectively as presented in Figure 1 of this paper and Figure 5
of work by Thuma et al. (2000). The respective beams for the two observations are plotted
and the rest of the keys, contour lines, axes are as presented in the indicated figures. A scale
bar of ∼ 1.5 kpc is shown on the maps.
-- 26 --
Fig. 3. -- The central 76′′ × 56′′ region of M 82 at 850 (left panels - a & c) and 450 µm
(right panels - b & d) overlaid with the integrated-intensity (black contours) of the CO (1-0)
interferometry data by Shen & Lo (1995). The top panels are plotted at full resolution of the
CO contours and the bottom panels are with the CO contours smoothed to beam sizes similar
to the SCUBA-850 and -450 µm beams (∼ 14′′.5 and 8′′.5, respectively). The white 850 and
450 µm contours on the respective maps are [200 to 1175] and [900 to 6000] mJy/beam, as
displayed in Figure 1. The red 850 and 450 µm contours are at 1350 and 6800 mJy/beam
and are plotted to highlight the position of the respective western submm peak. The cross
marks the 2.2 µm peak (Dietz et al. 1986). The keys are color-coded intensities in Jy/beam
and the X and Y axes are J2000 coordinates.
-- 27 --
Fig. 4. -- The left (a) and right (b) panels respectively show the B-band maps of M 82
obtained by the Hubble Legacy Team (Mutchler et al. 2007), respectively overlayed with
the 450-µm continuum-emission contours shown in Figures 1 and CO (1-0) interferometry
data by Shen & Lo (1995). For best contrast, the optical intensities are inversely plotted
and therefore the light patches are extinction features. The centeral cross and open squares
respectively mark positions of the 2.2 µm peak by Dietz et al. (1986) and mid-infrared, star-
forming clusters by Lipscy & Plavchan (2004). The X and Y axes are J2000 coordinates.
-- 28 --
Fig. 5. -- The left (a) and right (b) panels respectively show the low- and high-brightness
features of the Hα maps obtained by the Hubble Legacy Team (Mutchler et al. 2007), over-
layed with SCUBA 850 and 450-µm continuum-emission contours as shown in Figures 1.
The overlays were made by aligning the sky co-ordinates of the Hubble maps with those of
the SCUBA ones. The respective resolutions in the Hubble, SCUBA 850-µm, and 450-µm
maps are . 1′′, ∼ 14′′.5, and ∼ 8′′.5. The overlay in the left panel demonstrates the spa-
tial co-incidence between the large-scale low-brightness Hα emission and 850-µm continuum,
while the right panel depicts the origin of the large-scale Hα emission in the intensive star-
formation inner region of M 82's disk about the submm emission peaks. For best contrast,
the optical intensities are inversely plotted and the light patches (e.g. across the center of
the image in the right panel) are foreground extinction features. The X and Y axes are J2000
coordinates.
This figure "f1a.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f1b.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f1c.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f1d.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f2a.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f2b.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f3a.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f3b.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f3c.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f3d.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f4a.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f4b.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f5a.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
This figure "f5b.jpeg" is available in "jpeg"(cid:10) format from:
http://arxiv.org/ps/0812.3682v1
|
astro-ph/9908267 | 1 | 9908 | 1999-08-24T13:39:06 | How Observations of Circumstellar Disk Asymmetries Can Reveal Hidden Planets: Pericenter Glow and its Application to the HR 4796 Disk | [
"astro-ph"
] | Recent images of the disks of dust around the young stars HR 4796A and Fomalhaut show, in each case, a double-lobed feature that may be asymmetric (one lobe may be brighter than the other). A symmetric double-lobed structure is that expected from a disk of dust with a central hole that is observed nearly edge-on (i.e., close to the plane of the disk). This paper shows how the gravitational influence of a second body in the system with an eccentric orbit would cause a brightness asymmetry in such a disk by imposing a "forced eccentricity" on the orbits of the constituent dust particles, thus shifting the center of symmetry of the disk away from the star and causing the dust near the forced pericenter of the perturbed disk to glow. Dynamic modeling of the HR 4796 disk shows that its 5% brightness asymmetry could be the result of a forced eccentricity as small as 0.02 imposed on the disk by either the binary companion HR 4796B, or by an unseen planet close to the inner edge of the disk. Since it is likely that a forced eccentricity of 0.01 or higher would be imposed on a disk in a system in which there are planets, but no binary companion, the corresponding asymmetry in the disk's structure could serve as a sensitive indicator of these planets that might otherwise remain undetected. | astro-ph | astro-ph | How Observations of Circumstellar Disk Asymmetries Can Reveal Hidden
Planets: Pericenter Glow and its Application to the HR 4796 Disk
M. C. Wyatt, S. F. Dermott, C. M. Telesco, R. S. Fisher,
K. Grogan, E. K. Holmes and R. K. Pina
Department of Astronomy, University of Florida, Gainesville, FL 32611
ABSTRACT
Recent images of the disks of dust around the young stars HR 4796A (Jayawardhana
et al. 1998; Koerner et al. 1998; Schneider et al. 1999; Telesco et al. 1999) and
Fomalhaut (Holland et al. 1998) show, in each case, a double-lobed feature that may
be asymmetric (one lobe may be brighter than the other). A symmetric double-lobed
structure is that expected from a disk of dust with a central hole that is observed
nearly edge-on (i.e., close to the plane of the disk). This paper shows how the
gravitational influence of a second body in the system with an eccentric orbit would
cause a brightness asymmetry in such a disk by imposing a "forced eccentricity" on
the orbits of the constituent dust particles, thus shifting the center of symmetry of
the disk away from the star and causing the dust near the forced pericenter of the
perturbed disk to glow. Dynamic modeling of the HR 4796 disk shows that its ∼ 5%
brightness asymmetry could be the result of a forced eccentricity as small as 0.02
imposed on the disk by either the binary companion HR 4796B, or by an unseen planet
close to the inner edge of the disk. Since it is likely that a forced eccentricity of 0.01
or higher would be imposed on a disk in a system in which there are planets, but no
binary companion, the corresponding asymmetry in the disk's structure could serve as
a sensitive indicator of these planets that might otherwise remain undetected.
Subject headings: binaries: visual -- celestial mechanics, stellar dynamics --
circumstellar matter -- planetary systems -- stars: individual (HR 4796)
9
9
9
1
g
u
A
4
2
1
v
7
6
2
8
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1.
Introduction
A new generation of astronomical instrumentations is now making it possible to image the
thermal emission from the disks of dust that surround some stars. The disks that have been
imaged around main sequence stars (i.e., the Vega-type, or debris, disks) have aroused considerable
interest because it is hoped that interpretation of their structure will provide valuable information
about the evolution of planetary systems, possibly even leading to the indirect detection of planets
hiding in the disks. One such disk is that around the A0V star HR 4796A. Mid-IR (10-20 µm)
images of the HR 4796 disk (Jayawardhana et al. 1998; Koerner et al. 1998; Telesco et al. 1999)
-- 2 --
show that its emission is concentrated in two lobes, one on either side of the star, indicating that
the disk is being observed nearly edge-on, and that its inner region is almost completely devoid of
dust. The same double-lobed feature is seen in NICMOS (1.1 µm) images of the disk's scattered
light (Schneider et al. 1999). The size of the disk's central cavity, which was previously inferred
from the star's spectral energy distribution (Jura et al. 1993), is approximately 40 AU in radius
from the star, about the same size as the solar planetary system. Since the age of the HR 4796
system, tsys ≈ 10 Myr (Stauffer, Hartmann, & Barrado Y Navascu´es 1995; Jura et al. 1998) places
it at a stage in its evolution when the formation of any planets is expected to be almost complete
(e.g., Lissauer 1993), many authors have speculated that the central cavity could be indicative of
planetary formation in this inner region (Jura et al. 1995; Jayawardhana et al. 1998; Koerner et
al. 1998; Jura et al. 1998). A similar double-lobed feature has been observed in the disk around
the star Fomalhaut, with a similar explanation proposed for its origin (Holland et al. 1998). Some
authors, however, remain skeptical about the existence of planets in these systems (e.g., Kalas
1998).
Observations of the HR 4796 disk show a further interesting feature: the disk's lobes appear
to be of unequal brightness, although this observed asymmetry is of low statistical significance
(∼ 1.8σ). Mid-IR observations of the disk are described in a companion paper by Telesco et
al. (1999, hereafter T99), and their IHW18 (18.2 µm) waveband observation (their Fig. 4b) is
reproduced in this paper in Fig. 8a. Their observation suggests that the NE lobe (on the left
of the image in Fig. 8a) is ∼ 5% brighter than the SW lobe. This lobe asymmetry may also be
apparent in the NICMOS (1.1 µm) images of the disk (Schneider et al. 1999), and in the ∼ 20
µm images of Koerner et al. (1998). The brightnesses of the Fomalhaut disk's lobes also appear
to be asymmetric (Holland et al. 1998). There are many possible explanations for the HR 4796
disk's lobe asymmetry. This paper describes a model of the T99 observation that provides one
possible dynamical explanation for the asymmetry: that it is the long-term consequence of the
gravitational perturbations of one or more massive bodies on the disk (i.e., the consequence of the
system's "secular perturbations").
Quite apart from any speculation about a nascent planetary system in the inner region of the
HR 4796 disk, we know that the disk must have been gravitationally perturbed, since HR 4796 is
a visual binary system. The M dwarf star HR 4796B, located at a projected distance of 517 AU
(Jura et al. 1998), is the common proper motion companion of HR 4796A (Jura et al. 1993). An
understanding of how asymmetries in the observed structure of the zodiacal cloud, the tenuous
disk of dust in the solar system, are linked to the solar system's planets (Dermott et al. 1999),
shows that if there is at least one massive perturber in the HR 4796 system that is on an eccentric
orbit, then the system's secular perturbations would have caused the disk's center of symmetry to
be offset from the star (Dermott et al. 1998). This offset would mean that the material in one of
the disk's observed lobes is closer to the star than that in the other lobe; consequently this lobe
would be hotter and brighter. The aim of this paper is to ascertain how large the perturbations
would have to be to cause the observed 5% asymmetry, and to discuss whether perturbations of
-- 3 --
this magnitude are physically realistic, or even to be expected, in this system.
Since this paper is based on the interpretation of a circumstellar disk observation, it starts
in §2 with a consideration of how information about a disk is stored in an observation. §3 gives
a comprehensive discussion of the physical processes that govern the dynamical evolution of a
disk's particles, and sets up a broad theoretical framework (the "dynamic disk") with which to
investigate a disk's structure. §4 then shows how secular perturbations cause a disk's structure
to have offset and warp asymmetries, and provides evidence that these asymmetries have been
observed in the structure of the zodiacal cloud. §5 uses the theoretical discussions of §§2-4 to
create an offset model for the HR 4796 disk that matches the 18.2 µm brightness distribution
observed by T99. The interpretation of this model is discussed in §6.
2. The Observable Circumstellar Disk
The observed brightness of a circumstellar disk comes from two sources: starlight that has
been absorbed by the disk particles and re-emitted as thermal radiation (primarily at mid-IR,
far-IR, and submillimeter wavelengths at λ > 5 µm), and starlight that has been scattered by
the disk particles (primarily at optical and near-IR wavelengths at λ < 2 µm). This paper only
discusses a disk's thermal emission.
2.1. Thermal Emission of a Single Disk Particle
A particle of diameter D, that is at a distance r from a star, is heated by the stellar radiation
to a temperature T that can be calculated from the equilibrium between the energy that the
particle absorbs and that which it re-emits as thermal radiation. This temperature depends on
the particle's optical properties (Gustafson 1994):
T (D, r) = [hQabsiT⋆ /hQabsiT (D,r)]1/4 ∗ Tbb,
(1)
where this equation must be solved iteratively, since the particle's temperature appears on both
sides of the equation, hQabsiT⋆ and hQabsiT (D,r) are the particle's absorption efficiency averaged
over the stellar spectrum (which can be approximated as that of a black body radiating at the
star's effective temperature T⋆) and the spectrum of a black body radiating at a temperature T ,
and Tbb is the equilibrium temperature of the particle if it were a black body:
Tbb = 278.3qa⊕/r ∗ (4σ/A)(L⋆/L⊙)1/4,
(2)
where Tbb is given in K, σ/A is the ratio of the particle's cross-sectional area to its surface area
(e.g., spherical particles have σ = πD2/4 and A = πD2, giving σ/A = 1/4), and L⋆ and L⊙ are
the luminosities of the star and the Sun.
-- 4 --
If this particle is at a distance R⊕ from the Earth, the contribution of its thermal emission to
the flux density at a wavelength, λ, received at the Earth is given by:
Fν (λ, D, r) = Qabs(λ, D)Bν [λ, T (D, r)]Ω(D),
(3)
where Bν is the Planck function, and Ω = σ/R2
cross-sectional area of the particle.
⊕ is the solid angle subtended at the Earth by the
2.2. Definition of Disk Structure
A circumstellar disk consists of particles with a range of sizes, compositions and morphologies.
Throughout this paper, however, disk particles are assumed to have the same composition and
morphology, and a particle's size is characterized by its diameter, D. Disk particles span a range
of sizes from Dmin, the smallest (probably submicron-sized) particles sustainable for a given disk,
up to Dmax, the largest (probably kilometer-sized) members of the disk that were formed from the
proto-planetary disk. The spatial distribution of these particles can be defined by n(D, r, θ, φ),
where n(D, r, θ, φ)dD is the volume density (number per unit volume) of particles in the size
range D ± dD/2 at a location in the disk defined by r, the radial distance from the star, θ, the
longitude relative to an arbitrary direction, and φ, the latitude relative to an arbitrary reference
plane. However, since it is a particle's cross-sectional area that is apparent in an observation
(eq. [3]), a disk's observable structure is better defined in terms of σ(D, r, θ, φ) = n(D, r, θ, φ)σ,
the cross-sectional area per unit volume per unit diameter.
The definition of a disk's structure can be simplified by assuming the size distribution of its
particles to be independent of θ and φ:
σ(D, r, θ, φ) = ¯σ(D, r)σ(r, θ, φ),
where ¯σ(D, r)dD is the proportion of the total cross-sectional area of the disk at r that is in
particles in the size range D ± dD/2, and
σ(r, θ, φ) = Z Dmax
Dmin
σ(D, r, θ, φ)dD
is the spatial distribution of cross-sectional area of particles of all sizes in the disk.
(4)
(5)
2.3. Line of Sight Brightness of a Disk Observation
To determine the observed brightness of a circumstellar disk in, for example, one pixel of
an image of the disk, two components of the observation must be defined: the vector, R, which
extends from the observer to the disk, and which describes how the line of sight intersects the disk
-- 5 --
in terms of r, θ, and φ; and the solid angle of the observation, Ωobs, where Ωobs = d2
width dpix radians.
pix for pixels of
Consider a volume element along this line of sight that is at a location in the disk defined by
r, θ, φ, and that has a length dR; the element volume is dV = ΩobsR2
⊕dR. The contribution of
the thermal emission of the particles in this element to the disk's brightness in the observation is
given by:
dFν (λ, r, θ, φ)/Ωobs = Z Dmax
Dmin
= P (λ, r)σ(r, θ, φ)dR,
Qabs(λ, D)Bν [λ, T (D, r)]σ(D, r, θ, φ)dDdR,
(6)
(7)
(8)
where equation (7) uses the simplification for the disk structure given by equation (4), and
P (λ, r) = Z Dmax
Dmin
Qabs(λ, D)Bν [λ, T (D, r)]¯σ(D, r)dD.
Thus, the brightness of this element is not affected by the solid angle of the observation, neither is
it affected by the distance of the element from the Earth.
Equation (8) can also be written as:
P (λ, r) = hQabs(λ, D)Bν [λ, T (D, r)]iσ(D,r);
(9)
i.e., P (λ, r) is the combination of the particles' optical properties given by Qabs(λ, D)Bν [T (D, r), λ]
averaged over the disk's cross-sectional area distribution, σ(D, r). Thus, P (λ, r) can also be given
by:
P (λ, r) = Qabs(λ, Dtyp)Bν [λ, T (Dtyp, r)],
(10)
where Dtyp is the size of disk particle that characterizes (and hence dominates) the disk's emission.
This characteristic particle size could be different in different wavebands, as well as at different
distances from the star, but always lies in the range Dmin < Dtyp < Dmax; its value can be found
by considering the relative contribution of particles of different sizes to P (λ, r). However, unless
the optical properties of the particles prevent it, the particles that dominate a disk's emission are
also those that contribute most to its cross-sectional area, i.e., they are those that dominate the
disk's structure, σ(r, θ, φ). This is usually the case for mid-IR (λ = 10 − 20 µm) observations, such
as those of T99 (see e.g., §5.2).
The total brightness of the disk in this observation is the integral of equation (6) over R.
Thus, a disk observation is composed of three parts: the disk's structure, the optical properties of
the disk particles, and the orientation of the disk to the line of sight.
2.4. Real Circumstellar Disk Images
An image of the disk is made up of many pixels, each of which has a different line of sight
vector, R, and corresponding brightness. The image at the detector has been convolved with
-- 6 --
the observational point spread function (PSF), a combination of the seeing conditions and the
telescope and instrumental optics, which, in the diffraction-limited case, for an ideal instrument,
can be approximated as gaussian smoothing with FWHM = λ/D, where this D is the diameter of
the telescope. The observed image also contains photospheric emission from the point-like star, as
well as random noise fluctuations. A useful image of the disk can be recovered by subtracting the
image of the star, but the accuracy of this subtraction depends on how well both the PSF and the
stellar flux density are known. Further smoothing of the image to increase the signal-to-noise may
also prove useful.
3. The Dynamic Disk
This section outlines the theoretical framework upon which later discussion of circumstellar
disk structure (such as how secular perturbations affect this structure) is based. A disk is a
dynamic entity, the constituent particles of which are undergoing constant dynamical and physical
evolution; §3.1 gives an extensive discussion of the physical processes acting on disk particles. §3.2
then summarizes our understanding of the dynamic disk by showing how disk particles can be
categorized according to the dominant physical processes affecting their evolution.
3.1. Physical Processes
If we are to make generalizations about the physical processes relevant to circumstellar disk
evolution, then the zodiacal cloud is the best example of a circumstellar disk on which to base this
understanding, since its properties have been determined with a certain degree of confidence. This
confidence stems from the wealth of observational, theoretical, and physical evidence describing
its present state, its evolutionary history, and the physical environment of the system it is in.
The physical processes that are described in this section are those thought to have dominated
the evolution of the zodiacal cloud, supposedly since the Sun reached the main sequence (e.g.,
Leinert & Grun 1990; Gustafson 1994). These processes should serve as an adequate basis for an
understanding of the evolution of the circumstellar disks around other main sequence stars.
There is, however, one obvious distinction between the zodiacal cloud and exosolar dust disks.
The emission observed from the zodiacal cloud is dominated by that from dust in the inner solar
system (within 5 AU of the Sun), which has its origins in the asteroid belt (Dermott et al. 1984;
Grogan et al. 1997) and the short period comets (Sykes et al. 1986). In contrast, the emission
observed from exosolar dust disks is dominated by that originating from dust in regions analogous
to the Kuiper belt in the solar system, i.e., > 30 AU from the star (e.g., Backman & Paresce
1993). Relatively little is known about the Kuiper belt, but, like the inner solar system, it appears
to be populated with many asteroid- or comet-like objects that are probably the remnants of the
solar system's planetary formation phase (e.g., Jewitt 1999).
-- 7 --
3.1.1. Gravity
The dominant force acting on all but the smallest disk particles is the gravitational attraction
of the star:
Fgrav = GM⋆m/r2,
(11)
where G is the gravitational constant, M⋆ is the mass of the star, m is the mass of the particle,
and r is the distance of the particle from the star. All material is assumed to orbit the star on
Keplerian (elliptical) orbits, with other forces acting as perturbations to these orbits. The orbit of
a particle is defined by the following orbital elements: the semimajor axis, a, and eccentricity, e,
that define the radial extent and shape of the orbit; the inclination, I, and longitude of ascending
node, Ω, that define the plane of the orbit (relative to an arbitrary reference plane); and the
longitude of pericenter, ω, that defines the orientation of the orbit within the orbital plane (relative
to an arbitrary reference direction). The orbital period of the particle is given by:
tper = q(a/a⊕)3(M⊙/M⋆),
(12)
where tper is given in years, a⊕ = 1 AU is the semimajor axis of the Earth's orbit, and M⊙ is the
mass of the Sun. At any instant, the location of the particle in its orbit is defined by the true
anomaly, f , where f = 0◦ and 180◦ at the pericenter and apocenter, respectively, and its distance
from the star, r, and its velocity, v, are defined by:
r = a(1 − e2)/(1 + e cos f ),
v = qGM⋆(2/r − 1/a).
(13)
(14)
Since at any given time a particle could be at any point along its orbit, its contribution to the
distribution of material in a disk can be described by the elliptical ring that contains the mass of
the particle spread out along its orbit, the line density of which varies inversely with the particle's
velocity (eq. [14]). Each disk particle has an orbit defined by a different set of orbital elements,
with a contribution to the spatial distribution of material in the disk that can be described by a
corresponding elliptical ring. Thus, a disk's structure can be defined by the distribution of orbital
elements of its constituent particles, n(D, a, e, I, Ω, ω), where n(D, a, e, I, Ω, ω)dDdadedIdΩdω is
the number of disk particles with sizes in the range D ± dD/2, and orbital elements in the range
a ± da/2, e ± de/2, I ± dI/2, Ω ± dΩ/2, ω ± dω/2. A disk's orbital element distribution can be
quantified in terms of the evolutionary history of the system and the physical processes acting on
the disk's particles. Using techniques such as those described in §5.1.2, the resulting structure can
then be compared with the disk's observable structure, σ(D, r, θ, φ), to link a disk observation
with the physics of the particles in that disk. It is often convenient to discuss the dependence of
the orbital element distribution on the different parameters separately; e.g., the distribution of
semimajor axes, n(a), is defined such that n(a)da is the total number of particles with orbital
semimajor axes in the range a ± da/2.
-- 8 --
3.1.2. Collisions
A typical disk particle is created by the break-up of a larger "parent" body, either as the
result of a collision with another body, or simply by its disintegration. This parent body could
have been created by the break-up of an even larger body, and the particle itself will most likely
end up as a parent body for particles smaller than itself. This "collisional cascade" spans the
complete size range of disk material, and the particles that share a common ancestor are said to
constitute a "family" of particles.
The size distribution that results from this collisional cascade can be found from theoretical
arguments (Dohnanyi 1969):
n(D) ∝ D2−3q,
(15)
where q = 11/6; this distribution is expected to hold for disk particles that are large enough not to
be affected by radiation forces (§3.1.3). A disk with this distribution has its mass, m(D) = n(D)m,
concentrated in its largest particles, while its cross-sectional area, σ(D) = n(D)σ, is concentrated
in its smallest particles. Collisions in such a disk are mostly non-catastrophic (see eqs. [A2]-[A6]),
and a particle in this disk is most likely to be broken up by a particle that has just enough mass
(and hence energy) to do so. This in turn means that collisional fragments have velocities, and
hence orbital elements, that are almost identical to those of the original particle; i.e., in the absence
of other forces, all members of the same family have identical orbits. Due to the interaction of the
competing physical processes, the size distribution of disk particles that are affected by radiation
forces is only really understood qualitatively (§§3.1.3 and 3.2.2); their distribution is particularly
important, since, in general, a disk's cross-sectional area (and hence its observable structure) is
concentrated in these smaller particles.
The importance of collisions in determining a particle's evolution depends on its collisional
lifetime, which is discussed in Appendix A. The collisional lifetime of the particles that constitute
most of a disk's cross-sectional area (i.e., those that are expected to characterize the disk's mid-IR
emission, §2.3), can be approximated by (eqs. [A9] and [A12]):
tcoll(Dtyp, r) = tper(r)/4πτef f (r),
(16)
where tper(r) is the average orbital period of particles at r (eq. [12] with a replaced by r), and
τef f (r) is the disk's effective face-on optical depth (eq. [A8]), which would be equal to the disk's
true optical depth if its particles had unity extinction efficiency. The collisional lifetime of particles
with D > Dtyp can be considerably longer than that of equation (16) (see e.g., eq. [A15]), and a
disk's largest particles, those for which tcoll(D, r) > tsys, may not have suffered any catastrophic
collisions since they were first created; such particles are primordial particles. The cascades of very
young disks may still contain a significant proportion of primordial particles; i.e., their cascades
may not be fully evolved.
The collisional cascade theory is well-supported by evidence from the zodiacal cloud. The
size distribution of the largest (D > 3 km) members of the zodiacal cloud's collisional cascade,
-- 9 --
the observable asteroids, is well-approximated by equation (15) (Durda & Dermott 1997; Durda,
Greenberg, & Jedicke 1998); the distribution of the very largest (D > 30 km) asteroids deviates
from this distribution, however, because of the transition from strength-scaling to gravity-scaling
for asteroids larger than ∼ 150 m (Durda et al. 1998). The size distribution of the zodiacal cloud's
medium-sized (1 mm < D < 3 km) members is also expected to follow equation (15) (Durda &
Dermott 1997), but there is no observational proof of this, since these members are too faint to be
seen individually, and too few to be studied collectively (Leinert & Grun 1990). There is, however,
proof that the zodiacal cloud's collisional cascade extends from its largest members down to its
smallest dust particles: the shapes of the "dust band" thermal emission features (Low et al. 1984)
correspond to those expected from the small (1 − 1000 µm) particles resulting from the break-up,
some time ago, of a few very large asteroids, the largest fragments of which are still observable as
the asteroids in the Themis, Koronis, and, possibly, Eos families (Dermott et al. 1984; Grogan et
al. 1997). The size distribution of the zodiacal cloud's smallest (D < 1 mm) dust particles (e.g.,
Leinert & Grun 1990; Love & Brownlee 1993) can be explained qualitatively (e.g., Grun et al.
1985; see also §3.1.3).
Analysis of the collision rates of objects in the Kuiper belt (Stern 1995) shows that a
collisional cascade should exist here too; there is also evidence to suggest that the Kuiper belt was
once more massive than it is today (Jewitt 1999), meaning that in the past collisions would have
played a much larger role in determining its structure than they do today, maybe even causing
the supposed mass loss (Stern & Colwell 1997). The size distribution of the observed Kuiper
belt objects appears to be slightly steeper than that in the inner solar system (q > 11/6, Jewitt
1999), while observations have been unable, as yet, to determine its dust distribution (Backman,
Dasgupta, & Stencel 1995; Gurnett et al. 1997),
3.1.3. Radiation Forces, β
For most of the collisional cascade, gravity can be considered to be the only significant
force acting on disk particles. The smallest particles, however, are significantly affected by their
interaction with the photons from the star.
Radiation Pressure Radiation pressure is the component of the radiation force that points
radially away from the star. It is inversely proportional to the square of a particle's distance from
the star, and is defined for different particles by its ratio to the gravitational force of equation (11)
(Gustafson 1994):
β(D) = Frad/Fgrav = Cr(σ/m)hQpriT⋆ (L⋆/L⊙)(M⊙/M⋆),
(17)
where Cr = 7.65 × 10−4 kg/m2, σ/m is the ratio of the particle's cross-sectional area to its mass
(e.g., σ/m = 1.5/ρD for spherical particles of density ρ), and hQpriT⋆ = R Qpr(D, λ)Fλdλ/R Fλdλ
-- 10 --
is the particle's radiation pressure efficiency1 averaged over the stellar spectrum, Fλ.
An approximation for large particles is that hQpriT⋆ ≈ 1; thus, large spherical particles have:
(18)
β(D) ≈ (1150/ρD)(L⋆/L⊙)(M⊙/M⋆),
where ρ is measured in kg/m3, and D in µm. This approximation is valid for particles in the solar
system with D > 20 µm (Gustafson 1994). For particles in exosolar systems, this limit scales with
the wavelength at which the star emits most of its energy, λ⋆ ∝ 1/T⋆; i.e., equation (18) is valid
for spherical particles with D > 20(T⊙/T⋆) µm, where T⊙ = 5785 K is the effective temperature
of the Sun. Since β ∝ 1/D, this means that the smaller a particle, the larger its β; this holds
down to micron-sized particles, smaller than which β decreases to a level that is independent of
the particle's size (Gustafson 1994).
The effect of radiation pressure is equivalent to reducing the mass of the star by a factor
1 − β. This means that a particle for which β 6= 0 moves slower around the same orbit by a factor
of √1 − β than one for which β = 0 (eq. [14]). It also means that daughter fragments created
by the break-up of a parent body move on orbits that can differ substantially from that of the
parent. The reason for this, is that while the positions and velocities of a parent and its daughter
fragments are the same at the moment of break-up (apart from a small velocity dispersion), their
β are different, and so the daughter fragments move in effective potentials that are different from
that the parent moved in. Daughter fragments created in the break-up of a parent particle that
had β = 0, and for which orbital elements at the time of the collision were a, e, I, Ω, ω, and f , move
in the same orbital plane as the parent, I
,
eccentricities, e
= Ω, but on orbits with semimajor axes, a
, that are given by (Burns, Lamy, & Soter 1979):
, and pericenter orientations, ω
′
′
= I and Ω
′
′
′
′
a
′
e
= a(1 − β)/h1 − 2β(1 + e cos f )/(1 − e2)i ,
= (1 − β)−1qe2 + 2βe cos f + β2,
= arctan [β sin f /(β cos f + e)].
′
′
ω
− ω = f − f
(19)
(20)
(21)
Analysis of equations (19)-(21) shows that the orbits of the largest fragments, those for which
β < 0.1, are similar to that of the parent. On the other hand, the smallest fragments, those for
which β > 0.5(1 − e2)/(1 + e cos f ), have hyperbolic orbits (e
> 1); these particles are known
as "β meteoroids"2. Since β meteoroids are lost from the system on the timescale of the orbital
period of the parent (eq. [12]), the diameter of particle for which β > 0.5 essentially defines
the lower end of the collisional cascade. However, there may also be a population of submicron
′
1A particle's radiation pressure efficiency is related to its absorption and scattering efficiencies by Qpr =
Qabs + Qsca(1 − hcos θi), where hcos θi accounts for the asymmetry of the scattered radiation.
2Note that particles with β > 1 are β meteoroids even if they were not created collisionally, since they "see" a
negative mass star.
-- 11 --
particles that have β < 0.5 (Gustafson 1994). The intermediate-sized fragments, those for which
0.1 < β < 0.5, that we call "β critical" particles, have orbits that differ substantially from that
of the parent. However, the point of closest approach to the star of the orbits of all daughter
fragments, irrespective of their size, is the same as that of the parent: combining equations (19)
and (20) gives the pericenter distance of daughter fragments, r
), as
′
′
p = a
′
(1 − e
r
′
p/rp = 1 + e(1 − cos f ) + O(e2).
(22)
The orbits of collisional fragments with different β from a parent particle that was on a circular
orbit are shown in Fig. 1.
Poynting-Robertson (P-R) Light Drag The component of the radiation force tangential to
a particle's orbit is called the P-R drag force. This force is also proportional to β. It results in an
evolutionary decrease in both the semimajor axis and eccentricity of the particle's orbit (Burns et
al. 1979):
apr = −(α/a)(2 + 3e2)/(1 − e2)3/2 = −2α/a + O(e2)
epr = −(α/a2)2.5e/(1 − e2)1/2 = −2.5αe/a2 + O(e2),
(23)
(24)
Ipr = Ωpr = 0; neither does it affect the orientation of the particle's pericenter,
where α = 6.24 × 10−4(M⋆/M⊙)β AU2/year. P-R drag does not change the plane of the particle's
ωpr = 0.
orbit,
For a particle with zero eccentricity, equation (23) can be solved to find the time it takes for the
particle to spiral in from a radial distance of r1 to r2:
tpr = 400(M⊙/M⋆)[(r1/a⊕)2 − (r2/a⊕)2]/β,
(25)
where tpr is given in years.
Consider the daughter fragments created in the break-up of a parent body that was on an
orbit at a distance r from the star. The largest fragments are broken up by collisions before their
orbits have suffered any significant P-R drag evolution, while the smaller fragments, for which
the P-R drag evolution is faster, can reach the star without having encountered another particle
(at which point they evaporate). Particles for which P-R drag significantly affects their orbits in
their lifetime can be estimated as those for which their collisional lifetime (eq. [16]; the use of this
equation is justified in the next paragraph) is longer than their P-R drag lifetime (eq. [25] with
r2 = 0), i.e., those for which β > βpr, where
for large spherical particles, this is also those for which D < Dpr, where
βpr = 5000τef f (r)q(M⊙/M⋆)(r/a⊕);
ρ is measured in kg/m3, and Dpr in µm.
Dpr = [0.23/ρτef f (r)](L⋆/L⊙)q(M⊙/M⋆)(a⊕/r),
(26)
(27)
-- 12 --
Consider the daughter fragments created in the break-up of an endless supply of parent
bodies that are on orbits with the same semimajor axis, as. Ignoring collisional processes, the
fragments with orbits that are affected by P-R drag, those with β > βpr, have their semimajor
axes distributed from a = as to a = 0 according to (eq. [23]):
n(a) ∝ 1/ apr ∝ a;
(28)
this corresponds to a volume density distribution that is roughly inversely proportional to distance
from the star3. If the collisional processes leading to the size distribution of the parent bodies,
ns(D), still holds for the production of the P-R drag affected particles, then their size distribution
is given by:
n(D) ∝ ns(D)/ apr ∝ ns(D)D.
(29)
If ns(D) can be given by equation (15) with q = 11/6, the cross-sectional area of a disk's P-R
drag affected particles is concentrated in the largest of these particles, while that of its unaffected
particles is concentrated in the smallest of these particles; i.e., most of a disk's cross-sectional area
is expected to be concentrated in particles with Dtyp ≈ Dpr, justifying the use of equation (16) for
the collisional lifetime of these particles.
Observations of the zodiacal cloud at 1 AU show that its effective optical depth here is
τef f = O(10−7). Since these particles originated in the asteroid belt at ∼ 3 AU, arriving at 1
AU due to the P-R drag evolution of their orbits, the zodiacal cloud's volume density should
vary ∝ 1/r, and its effective optical depth at 3 AU should be similar to that at 1 AU. Assuming
zodiacal cloud particles to have a density ∼ 2500 kg/m3 (Leinert & Grun 1990), the cross-sectional
area of material in the asteroid belt should be concentrated in particles with Dpr = O(500 µm)
(eq. [27]), for which both the collisional lifetime, and the P-R drag lifetime, is ∼ 4 Myr. The
cross-sectional area of material at 1 AU is expected to be concentrated in particles smaller than
that in the asteroid belt, since many of the larger particles should have been broken up by
collisions before they reach the inner solar system; this is in agreement with observations that
show the cross-sectional area distribution at 1 AU to peak for particles with D = 100 − 200
µm (Leinert & Grun 1990; Love & Brownlee 1993). Also, equation (A15) with Dtyp = 500 µm,
Dcc(D)/D = (10−4)1/3, and q = 11/6, predicts that the collisional lifetime of large bodies in the
asteroid belt should be:
(30)
where tcoll is given in years, and D in km. Since the solar system is ∼ 4.5 × 109 years old, this
implies that asteroids larger than ∼ 20 km should be primordial asteroids; this is in agreement
with more accurate models of the observed size distribution of these asteroids (Durda et al. 1998).
tcoll ≈ 109√D,
3If the particles had circular orbits, equation (28) means a spherical shell of width dr, the volume of which is
∝ r2dr, would contain a number of particles that is ∝ rdr (see e.g., Gorkavyi et al. 1997).
-- 13 --
3.2. Division of a Disk into Particle Categories
Disk particles of different sizes can be categorized according to the dominant physical
processes affecting their evolution. Particles in the different categories have different lives; i.e.,
the way they are created, their dynamical evolution, and the way they are eventually destroyed,
are all different. Each of a disk's categories has a different spatial distribution; which of these
categories dominates the disk's observable structure depends on the relative contribution of each
to the disk's cross-sectional area (see §2.3).
3.2.1. Category Definitions
A disk's largest particles, those with β < 0.1, have orbital elements that are initially the
same as, or at least very similar to, those of their parents. Of these large particles, only those
with β < βpr suffer no significant P-R drag evolution to their orbits in their lifetime. These truly
large particles continue on the same orbits as those of their ancestors until they collide with a
particle large enough to cause a catastrophic collision; the resulting collisional fragments populate
the collisional cascade. The spatial distribution of these "large" particles in a disk is its base
distribution. The spatial distributions of a disk's smaller particles can only be understood in terms
of how they differ from the disk's base distribution.
Disk particles with βpr < β < 0.1, spiral in from their parent's orbits due to P-R drag, so that
they are closer to the star than their parents by the time of their demise (which could be caused
either by collisions or by evaporation close to the star); the spatial distribution of these "P-R drag
affected" particles in a disk differs from the disk's base distribution in that it extends closer in to
the star. The orbits of particles with 0.1 < β < 0.5 also undergo significant P-R drag evolution
before their demise, but their original orbits are already different from those of their parents; the
spatial distribution of these "β critical" particles in a disk extends both further out, and further
in, from the disk's base distribution. Particles with β > 0.5 leave their parents on hyperbolic
orbits, and so are quickly lost from the system; the spatial distribution of these "β meteoroids" in
a disk extends further out, but not further in, from the disk's base distribution.
Thus, a disk comprises four particle categories, each of which has a different spatial
distribution, although all are inextricably linked to that of the large particles through the
collisional cascade. For disks that have βpr > 0.5, however, i.e., those with
τef f (r) > 10−4q(M⋆/M⊙)(a⊕/r),
(31)
there is no significant P-R drag evolution of any of its constituent particles. Such disks comprise
just three categories, since the P-R drag affected category is empty.
-- 14 --
3.2.2. Category Cross-Sectional Area
The size distributions of a disk's large particles, and its P-R drag affected particles, were
discussed in §§3.1.2 and 3.1.3. These discussions imply that the cross-sectional area of a disk in
which there is a population of P-R drag affected particles (see eq. [31]) is dominated by particles
with D ≈ Dpr (eq. [27]). As a first-cut approximation, the size distribution of a disk in which there
are no P-R drag affected particles follows equation (15) from Dmax down to Dmin = D(β = 0.5);
however, since a particle's catastrophic collision rate is affected by the size distribution of particles
smaller than itself (Durda & Dermott 1997; eq. [A4]), this distribution cannot be expected to
hold all the way down to D(β = 0.5). This means that the cross-sectional area of such a disk is
concentrated in its smallest particles, and the contribution of β critical particles to the disk's total
cross-sectional area is given by
dσ/σtot = [D5−3q]D(β=0.5)
D(β=0.1)/[D5−3q]D(β=0.5)
Dmax
;
(32)
e.g., if q = 11/6, and β ∝ 1/D, this means that half of the disk's cross-sectional area comes from
its β critical particles.
Since β meteoroids have hyperbolic orbits, they are expected to contribute little to a disk's
cross-sectional area unless they are produced at a high enough rate to replenish their rapid loss
from the system. This could be the case if the disk was very dense, since material would pass
quickly through the cascade; such a disk would undergo considerable mass loss. An estimate of
how dense the disk would have to be for this to be the case depends on the assumptions made
about the physics of collisions between small particles. For heuristic purposes, it is assumed here
that the total cross-sectional area of β meteoroids created by the collisional break-up of a parent
body is comparable to that of the parent itself; this is probably an underestimate if the collision is
destructive, but an overestimate if the collision is erosive. If this were the case, then the disk's β
meteoroids would dominate a disk's cross-sectional area only if their lifetime, which is of the order
of the orbital period of their parents, is longer than the lifetime of these parents, which can be
approximated by equation (16), i.e., only if
τef f (r) > 0.1.
(33)
In conclusion, from a theoretical stand-point, there are few solid assumptions that can be
made about a disk's size distribution, other than that the denser a disk is, the smaller the diameter
of particles that its cross-sectional area is concentrated in.
3.3. The Perturbed Dynamic Disk
In addition to the physical processes described in §3.1, the particles of the dynamic disk
described in this section are affected by a number of perturbing processes; these produce subtle,
-- 15 --
but perhaps observable, changes in a disk's structure. The dominant perturbing processes in the
zodiacal cloud are the secular and resonant gravitational perturbations of the planets. Secular
perturbations are discussed in §4. Resonant perturbations give rise to planetary resonant rings:
a planet's resonant ring is an asymmetric circumstellar ring of material that co-orbits with the
planet as a result of the resonant trapping of particles into the planet's exterior mean motion
resonances (not to be confused with circumplanetary dust rings). There is both observational and
theoretical evidence for the existence of the Earth's resonant ring (Dermott et al. 1994; Reach et
al. 1995); many of the observed Kuiper belt objects are trapped in 2:3 resonance with Neptune
(Jewitt 1999), thus forming Neptune's resonant ring (Malhotra 1995).
Other possible perturbing processes include: stellar wind forces, that, at least for dust in the
solar system, effectively increase the value of β for P-R drag (e.g., Leinert & Grun 1990); Lorentz
forces acting on charged particles, that, while negligible for particles in the inner solar system,
are increasingly important for particles at distances further from the Sun (e.g., Kimura & Mann
1998); interactions with dust from the interstellar medium (e.g., Artymowicz & Clampin 1997);
the sublimation of icy dust grains, which is one of the mechanisms that has been suggested as the
cause of the inner hole in the HR 4796 disk (Jura et al. 1998); and the self gravity of a massive
disk, which could have played an important role in determining the evolution of the primordial
Kuiper belt (Ward & Hahn 1998).
4. Structure of a Secularly Perturbed Disk
The orbit of a particle in a circumstellar disk that is in a system in which there are one or
more massive perturbers is inevitably affected by the gravitational perturbations of these bodies.
The consequent evolution of the particle's orbit can be used to obtain a quantitative understanding
of the effect of these perturbations on the structure of the disk.
4.1. Secular Perturbation Theory
4.1.1. Perturbation Equations
The gravitational forces from a planetary system that act to perturb the orbit of a
particle in the system can be decomposed into the sum of many terms that are described by
the particle's disturbing function, R. The long-term average of these forces are the system's
secular perturbations, and the terms of the disturbing function that contribute to these secular
perturbations, Rsec, can be identified as those that do not depend on the mean longitudes of either
the planets or the particle (the other forces having periodic variations).
Consider a particle that is orbiting a star of mass M⋆, that also has Npl massive, perturbing,
bodies orbiting it. This particle has a radiation pressure force acting on it represented by β,
-- 16 --
and its orbit is described by the elements a, e, I, Ω and ω. To second order in eccentricities
and inclinations, the secular terms in the particle's disturbing function are given by (Brouwer &
Clemence 1961; Dermott et al. 1985; Dermott & Nicholson 1986; Murray & Dermott 1999):
Rsec = na2
1
2
A(e2 − I 2) +
Npl
Xj=1
[Ajeejcos(ω − ωj) + BjIIjcos(Ω − Ωj)]
,
(34)
where n = (2π/tyear)p(M⋆/M⊙)(1 − β)(a⊕/a)3 is the mean motion of the particle in rad/s,
tyear = 2π/qGM⊙/a3
⊕ = 3.156 × 107 s is one year measured in seconds, and
A = +
Aj = −
Bj = +
n
n
Npl
n
M⋆(cid:19) αjαjb1
Xj=1(cid:18) Mj
4(1 − β)
4(1 − β) (cid:18) Mj
M⋆(cid:19) αjαjb2
4(1 − β) (cid:18) Mj
M⋆(cid:19) αjαjb1
3/2(αj),
3/2(αj),
3/2(αj),
(35a)
(35b)
(35c)
where αj = aj/a and αj = 1 for aj < a, and αj = αj = a/aj for aj > a, and
bs
and Bj are in units of rad/s, and Rsec is in units of m2/s2.
j )−3/2 cos sψdψ are the Laplace coefficients (s = 1, 2). A, Aj
0 (1 − 2αj cos ψ + α2
3/2(αj) = (π)−1R 2π
The effect of these perturbations on the orbital elements of the particle can be found using
Lagrange's planetary equations (Brouwer & Clemence 1961; Murray & Dermott 1999). The
semimajor axis of the particle remains constant, asec = 0, while the variations of its eccentricity
and inclination are best described when coupled with the variations of its longitude of pericenter
and ascending node using the variables defined by its complex eccentricity, z, and complex
inclination, y:
z = e ∗ exp iω,
y = I ∗ exp iΩ,
(36a)
(36b)
where i2 = −1. Using these variables Lagrange's planetary equations give the orbital element
variations due to secular perturbations as:
zsec = +iAz + i
ysec = −iAy + i
Npl
Npl
Xj=1
Xj=1
Ajzj,
Bjyj,
(37a)
(37b)
where zj and yj are the complex eccentricities and inclinations of the perturbers, which have a
slow temporal variation due to the secular perturbations of the perturbers on each other (Brouwer
& Clemence 1961; Murray & Dermott 1999):
-- 17 --
zj(t) =
yj(t) =
Npl
Npl
Xk=1
Xk=1
ejk ∗ exp i(gkt + βk),
Ijk ∗ exp i(fkt + γk),
(38a)
(38b)
where gk and fk are the eigenfrequencies of the perturber system, the coefficients ejk and Ijk are
the corresponding eigenvectors, and βk and γk are constants found from the initial conditions of
the perturber system.
4.1.2. Solution to Perturbation Equations
Ignoring the evolution of a particle's orbital elements due to P-R drag, equations (37a)
and (37b) can be solved to give the secular evolution of the particle's instantaneous complex
eccentricity and inclination (a.k.a. the particle's osculating elements). This secular evolution is
decomposed into two distinct time-varying elements -- the "forced", subscript f , and "proper",
subscript p, elements -- that are added vectorially in the complex planes (see e.g., Fig. 2a):
z(t) = zf (t) + zp(t) =
y(t) = yf (t) + yp(t) =
Npl
Npl
Xk=1
Xk=1
P
P
Npl
j=1 Ajejk
gk − A
fk + A
Npl
j=1 BjIjk
∗ exp i(gkt + βk) + ep ∗ exp i(+At + β0),
∗ exp i(fkt + γk) + Ip ∗ exp i(−At + γ0),
(39a)
(39b)
where ep, β0, and Ip, γ0 are determined by the particle's initial conditions.
These equations have simple physical and geometrical interpretations. A particle's forced
elements, zf and yf , depend only on the orbits of the perturbers in the system (that have a slow
secular evolution, eqs. [38a] and [38b]), as well as on the particle's semimajor axis (which has no
secular evolution). Thus, at a time t0, a particle that is on an orbit with a semimajor axis a, has
forced elements imposed on its orbit by the perturbers in the system that are defined by zf (a, t0)
and yf (a, t0). The contribution of the particle's proper elements to its osculating elements, z(t0)
and y(t0), is then given by: zp(t0) = z(t0) − zf (a, t0) and yp(t0) = y(t0) − yf (a, t0); thus defining
the particle's proper eccentricity, ep, and proper inclination, Ip, which are its fundamental orbital
elements (i.e., those that the particle would have if there were no perturbers in the system), as
well as the orientation parameters β0 and γ0. Since both the forced elements, and the osculating
elements, of collisional fragments are the same as those of their parent (apart from fragments with
β > 0.1), particles from the same family have the same proper elements, ep and Ip.
The evolution of a particle's proper elements is straight-forward -- they precess around circles
of fixed radius, ep and Ip, at a constant rate, A, counterclockwise for zp, clockwise for yp. The
-- 18 --
secular precession timescale depends only on the semimajor axis of the particle's orbit:
tsec = 2π/Atyear,
(40)
where tsec is given in years, and A is given in equation (35a); secular perturbations produce long
period variations in a particle's orbital elements (e.g., tsec = O(0.1 Myrs) in the asteroid belt).
The centers of the circles that the proper elements precess around are the forced elements (see e.g.,
Fig. 2a). Actually the forced elements vary on timescales that are comparable to the precession
timescale (eq.[40]); thus, it might appear ambitious to talk of the precession of a particle's
osculating elements around circles when its real evolutionary track in the complex eccentricity
and complex inclination planes may not be circular at all. The reason it is presented as such is
that at any given time, all of the particles at the same semimajor axis precess (at the same rate)
around the same forced elements on circles of different radii, and this has consequences for the
global distribution of orbital elements (see §4.2.1).
There are two things that are worth mentioning now about a particle's forced elements. If
there is just one perturber in the system, Npl = 1, its complex eccentricity and complex inclination
do not undergo any secular evolution, and the forced elements imposed on a particle in the system
are not only constant in time, but also independent of the mass of the perturber:
zf = hb2
3/2(αj)/b1
yf = Ij ∗ exp iΩj.
3/2(αj)i ej ∗ exp iωj,
(41a)
(41b)
This implies that a body of low mass, such as an asteroid, has as much impact on a particle's
orbit as a body of high mass, such as a Jupiter mass planet. The perturbations from a smaller
perturber, however, produce longer secular precession timescales (eq. [40]):
tsec = 4[αj αjb1
3/2(αj)(a⊕/a)3/2(Mj/M⋆)qM⋆/M⊙]−1;
(42)
they would also be similar in magnitude to those of the disk's self-gravity, which in that case could
no longer be ignored. If there is more than one perturber in the system, Npl > 1, then particles
on orbits for which their precession rate equals one of the system's eigenfrequencies (A = gk, or
−A = fk) have infinite forced elements imposed on their orbits, and so are quickly ejected from
such a "secular resonance" region.
The solution given by equations (39a) and (39b) accounts for the fact that small particles see
a less massive star due to the action of radiation pressure, but not for the P-R drag evolution of
their orbits; the solution for these particles is discussed in Appendix B. Also, the perturbation
theory of §4.1.1 is only valid for particles with small eccentricities; i.e., it is not valid for the
evolution of a disk's β critical particles, or its β meteoroids. However, if the evolution of a disk's β
critical particles is affected by secular perturbations (i.e., if their lifetime is longer than the secular
timescale), then it is probably also affected by P-R drag (i.e., their lifetime is probably also longer
than the P-R drag timescale), in which case the disk's β critical particles do not contribute much
to its observable structure (§§3.1.3 and 3.2.2). There is no secular evolution to the orbits of β
meteoroids because of their short lifetimes.
-- 19 --
4.2. Offset and Warp
The effect of secular perturbations on the structure of a disk can be understood by considering
the effect of the secular evolution of the constituent particles' orbits on the distribution of their
orbital elements. The perturbation equations of §4.1.1 show that secular pertubations affect
only the distribution of disk particles' complex eccentricities, n(z), and complex inclinations,
n(y), while having no effect on their size distribution (and hence the division of the disk into its
particle categories), or on their semimajor axis distribution (and hence the disk's large-scale radial
distribution).
4.2.1. Offset and Plane of Symmetry of Family Material
Consider the family of collisional fragments originating from a primordial body, the orbital
elements of which were described by a, ep, and Ip. Here we consider only fragments that are
unaffected by P-R drag (those that are affected by P-R drag are discussed in Appendix B).
Large (β < 0.1) Fragments The orbital elements of the largest fragments, those with β < 0.1,
created in the break-up of the primordial body are initially very close to those of the primordial
body; they do not have identical orbits due to the velocity dispersion imparted to the fragments in
the collision. The forced elements imposed on the orbits of all of these collisional fragments are the
same as those imposed on the primordial body. The secular evolution of their osculating complex
eccentricities (eq. [39a]) and complex inclinations (eq. [39b]), is to precess about the forced
elements (which are also varying with time), but at slightly different rates (due to their slightly
different semimajor axes). A similar argument applies for all particles created by the collisional
break-up of these fragments. Thus, after a few precession timescales, the complex eccentricities
and complex inclinations of the collisional fragments of this family lie evenly distributed around
circles that are centered on zf (a, t) and yf (a, t), and that have radii of ep and Ip (e.g., their
complex eccentricities lie on the circle shown in Fig. 2a), while their semimajor axes are all still
close to a. This is seen to be the case in the asteroid belt: there are families of asteroids that
have similar a, ep, and Ip, that are the collisional fragments resulting from the break-up of a much
larger asteroid (Hirayama 1918).
The distribution of the complex eccentricities, n(z), of these particles, has a distribution
of pericenters that is biased towards the orientation in the disk that is defined by ωf . The
consequence of this biased orbital element distribution on the spatial distribution of this family
material is best described with the help of Fig. 2b. This shows a face-on view (i.e., perpendicular
to the plane of symmetry), of the family material in orbit around a star S. The resulting disk
is made up of particles on orbits that have the same a, ef , ωf , and ep, but random ωp. The
contribution of each particle to the spatial distribution of material in the disk can be described by
the elliptical ring of material coincident with the particle's orbit (see §3.1.1). In Fig. 2b, these
-- 20 --
elliptical rings have been represented by uniform circles of radius a, with centers that are offset
by ae in a direction opposite to the pericenter direction, ω (this is a valid approximation to first
order in the particles' eccentricities); a heavy line is used to highlight the orbital ring with a
pericenter located at P , and a displaced circle center located at D, where DP = a. The vector
SD can be decomposed into its forced and proper components; this is shown by the triangle SCD,
where SD = ae, SC = aef , and CD = aep (there is a similar triangle in Fig. 2a). Given that the
distribution of ωp is random, it follows that the distribution of the rings' centers, D, for the family
disk are distributed on a circle of radius aep and center C. Thus, the family forms a uniform torus
of inner radius a(1 − ep) and outer radius a(1 + ep) centered on a point C displaced from the
star S by a distance aef in a direction away from the forced pericenter, ωf (Dermott et al. 1985;
Dermott et al. 1998).
The distribution of the complex inclinations, n(y), of these particles, is also the distribution
of their orbital planes. Changing the reference plane relative to which the particles' orbital
inclinations are defined to that described by yf , shows that the secular complex inclination
distribution of this family material leads to a disk that is symmetrical about the yf plane; the
flaring of this disk is described by Ip.
Small (β > 0.1) Fragments Since the small particles in this family originate from the larger
particles, the orbital elements of the parents of the small particles have the same a, ep and Ip
(they also have the same zf and yf ), but random ωp and Ωp. Consider the β critical particles that
are produced at the same time from the population of family particles that have the same ωp and
Ωp at that time; this parent population is spread out along the orbit defined by the elements a,
e, and ω, which could be one of the rings shown in Fig. 2b. The average pericenter orientation of
these β critical particles, hω
i, is the same as that of the orbit of the larger particles, ω (obvious
because of the symmetry of eq. [21] with respect to f ); their pericenter locations are also the same
(eq. [22]). Thus, the ring shown in Fig. 2b defines the inner edge of the disk of β critical particles
created in the break-up of large particles on this ring; i.e., their disk is offset by an amount ae
in the ω direction. Consequently, the inner edge of the disk of β critical particles created in the
break-up of all large particles in this family is offset by an amount aef in the ωf direction. This
disk has the same plane of symmetry as the families' large particles, since all particle categories
from the same family have the same distribution of orbital planes, n(y). Similar arguments apply
for the families' β meteoroids.
′
4.2.2. Offset and Warp of Whole Disk
The disks of material from all of the families that have the same semimajor axis, or
equivalently that are at the same distance from the star, have the same offset inner edge, and
the same plane of symmetry. This is because their large particles have the same forced elements
imposed on their orbits. The complex eccentricities and complex inclinations of the large particles
-- 21 --
of all of these families lie evenly distributed around circles with the same centers, zf and yf , but
with a distribution of radii, n(ep) and n(Ip), that are the distributions of the proper elements of
these families (defining the width and flaring of the torus consisting of these families' material).
The whole disk is made up of families with a range of semimajor axes; the families that are at
different semimajor axes can have different forced elements imposed on their orbits, and in view
of the proper element distributions in the asteroid belt and in the Kuiper belt, they can also have
different distributions of proper elements.
If the forced eccentricity imposed on the disk is non-zero, which it is if there is at least one
perturber in the system that is on a non-circular orbit (eqs. [39a] and [41a]), then the disk's center
of symmetry is offset from the star. If the forced inclination imposed on the disk is different for
families at different semimajor axes, which it is if there are two or more perturbers in the system
that are moving on orbits that are not co-planar (eqs. [39b] and [41b]), then the disk's plane of
symmetry varies with distance from the star; i.e., the disk is warped.
4.2.3. Physical Understanding of Offset and Warp
There is also a physical explanation for the secular perturbation asymmetries. The secular
perturbations of a massive body are equivalent to the gravitational perturbations of the elliptical
ring that contains the mass of the perturber spread out along its orbit, the line density of which
varies inversely with the speed of the perturber in its orbit (Gauss' averaging method, Brouwer
& Clemence 1961; Murray & Dermott 1999). The ring's elliptical shape, as well as its higher
line density at the perturber's apocenter, mean that the center of mass of the "star-ring" system
is shifted from the star towards the perturber's apocenter. The focus of the orbits of particles
in such a system is offset from the star; i.e., the center of symmetry of a disk in this system is
offset from the star. The gravitational perturbations of the ring also point to the plane coincident
with the perturber's orbital plane. In systems with two or more perturbers, the system's plane of
symmetry (that in which the perturbing forces out of this plane cancel) varies with distance from
the star; i.e., a disk in such a system is warped.
4.3. Observational Evidence of Offset and Warp in the Zodiacal Cloud
Mid-IR geocentric satellite observations (such as the IRAS, COBE, and ISO observations)
are dominated by the thermal emission of the zodiacal cloud's P-R drag affected particles in all
directions except that of the galactic plane (Leinert & Grun 1990). Such observations contain
detailed information about the spatial structure of the zodiacal cloud, especially since their
observing geometry changes throughout the year as the Earth moves around its orbit. Since
there are 9 massive perturbers in the solar system, the resulting secular perturbation asymmetries
should be apparent in the IRAS, COBE, and ISO data-sets (see Appendix B for a discussion of
-- 22 --
why the distribution of a disk's P-R drag affected particles should also contain the signatures of
the system's secular perturbations).
Fig. 3a shows COBE observations of the sum of the brightnesses in the 25 µm waveband
at the north and south ecliptic poles, (N + S)/2 (Dermott et al. 1999), where there is no
contamination from the galactic plane. If the zodiacal cloud was rotationally symmetric with
the Sun at the center, then the cross-sectional area density of particles in the near Earth region
would vary according to σ(r, θ, φ) ∝ r−νf (φ), where ν is a constant. Because the Earth's orbit is
eccentric, geocentric observations sample the zodiacal cloud at different radial distances from the
Sun. Thus, the minimum of the (N + S)/2 observation is expected to occur either at the Earth's
aphelion, λ⊕ = 282.9◦, or perihelion, λ⊕ = 102.9◦, depending on whether ν > 1 or ν < 1, which
is determined by the collisional evolution of particles in the near-Earth region (e.g., Leinert &
Grun 1990 discuss the observational evidence and conclude that ν ≈ 1.3 as found by the Helios
zodiacal light experiment). However, the minimum in the 25 µm waveband observations occurs at
λ⊕ = 224◦, and a similar result is found in the 12 µm waveband. This is expected only if the Sun
is not at the center of symmetry of the zodiacal cloud. Parametric models of the zodiacal cloud
have also shown the need for an offset to explain the observations (e.g., Kelsall et al. 1998).
Fig. 3b shows the variation of the brightnesses of the ecliptic poles with ecliptic longitude
of the Earth (Dermott et al. 1999). The north and south polar brightnesses are equal when the
Earth is at either the ascending or descending node of the local (at 1 AU) plane of symmetry
of the cloud, giving an ascending node of Ωasc = 70.7 ± 0.4◦. However, COBE observations of
the latitudes of the peak brightnesses of the zodiacal cloud measured in the directions leading
and trailing the Earth's orbital motion give Ωasc = 58.4◦ (Dermott et al. 1996). Since such
observations sample the cloud external to 1 AU, this implies that the plane of symmetry of the
zodiacal cloud varies with heliocentric distance, i.e., that the zodiacal cloud is warped.
To observe the zodiacal cloud's offset and warp asymmetries, an observer outside the solar
system, would, at the very least, need an observational resolution greater than the magnitude
of these asymmetries; e.g., to observe the offset asymmetry, the observer needs a resolution of
> (aef /a⊕)/R⊙ arcseconds, where the distance from the observer to the Sun, R⊙, is measured in
pc. An offset would be more observable in a disk such as HR 4796 due to its central cavity, since
this causes a brightness asymmetry in the emission from the inner edge of the disk.
5. The Secularly Perturbed HR 4796 Disk Model
This section describes our model of the HR 4796 disk that accounts for the brightness
distribution seen in the IHW18 waveband observation of T99 (Fig. 8a). There are three
components of the observation that had to be included in the model (§2): the disk's structure,
σ(r, θ, φ), defined in §2.2; the combination of the optical properties and the size distribution of the
disk's particles given by P (λ, r), defined in §2.3; and the disk's orientation. While the modeling
-- 23 --
techniques used here are new to the study of circumstellar disks, they have already been widely
used to study the observed structure of the zodiacal cloud (see, e.g., Dermott et al. 1994; Grogan
et al. 1997), and so can be used with a certain degree of confidence.
5.1. Model of Offset Disk Structure, σ(r, θ, φ)
The disk's structure was modeled as that of a secularly perturbed dynamic disk. A
combination of the theory of §§3 and 4, and inferences from the observation (Fig. 8a), was used to
parameterize the distribution of the orbital elements of the disk's large particles (§5.1.1). This was
used to create parameterized models of the spatial distribution of these large particles (§5.1.2),
that could then be compared with the observed spatial distribution. Since the observed spatial
distribution is that of the "emitting" particles, the underlying assumption is that these emitting
particles either are the disk's large particles, or that their spatial distribution is the same as that
of the disk's large particles (whether they are large or not). The implications of this assumption
are discussed in the interpretation of the model (§6).
5.1.1. Distribution of Orbital Elements, σ(a, e, I, Ω, ω)
The quintessentially secular part of the distribution of the orbital elements of the large
particles in a secularly perturbed disk is the distributions of their complex eccentricities, n(z), and
complex inclinations, n(y). For a particle with an orbit of a given semimajor axis, a, its complex
eccentricity, z, and complex inclination, y, are the addition of forced elements, zf (a) and yf (a),
to proper elements that have ωp and Ωp chosen at random, while ep and Ip are chosen from the
distributions n(ep) and n(Ip).
Since there is insufficient information available to determine the variation of the forced and
proper elements with semimajor axis in this disk, they were assumed to be constant across the
disk. The forced elements were left as model variables: the forced eccentricity, ef , defines the
magnitude of the offset asymmetry in the disk model; the forced pericenter orientation, ωf , defines
the orientation of this asymmetry; and the forced inclination, yf , defines the plane of symmetry
of the disk model. When creating a disk model, both ωf and yf were set to zero; these were
incorporated later into the description of the disk's orientation to our line of sight (see §5.3). The
distributions of the proper eccentricities, n(ep), and proper inclinations, n(Ip), of particles in the
disk model were taken to be like those of the main-belt asteroids with absolute magnitudes H < 11
(Bowell 1996). These large asteroids constitute a bias-free set (Bowell 1996) and have mean proper
eccentricities and proper inclinations of hepi = 0.130 and hIpi = 10.2◦. Not enough Kuiper belt
objects have been discovered yet to infer a bias-free distribution for their orbital elements.
The distribution of the semimajor axes, n(a), of particles in the disk defines its radial
distribution. There is no way of guessing this distribution from theoretical considerations, since it
-- 24 --
depends on the outcome of the system's planetary formation process, which varies from system to
system (compare the distribution of the solar system's planets, and its disk material, with those
found in exosolar systems, e.g., Backman & Paresce 1993; Marcy & Butler 1998). Thus, it had to
be deduced purely observationally. Fig. 8a shows that the disk has an inner edge, inside of which
there is a negligible amount of dust; this was modeled as a sharp cut-off in the distribution of
semimajor axes at amin, a model variable. The observation also shows that the disk has an outer
edge at ∼ 130 AU; this was modeled as a sharp cut-off in the distribution of semimajor axes at
amax = 130 AU (this is a non-critical parameter, since particles near the outer edge of the disk
contribute little to the observation, see §5.2.3). The distribution between amin and 130 AU was
taken as n(a) ∝ aγ, where n(a)da is the number of particles on orbits with semimajor axes in
the range a ± da/2, and γ is a model variable. To get an idea of the radial distribution resulting
from this semimajor axis distribution, consider that if the particles had zero eccentricity, this
distribution would result in a volume density (number of particles per unit volume) distribution
that is ∝ rγ−2 (since the number of particles in a spherical shell of width dr, the volume of which
is ∝ r2dr, would contain a number of particles that is ∝ rγdr).
5.1.2. Conversion to Spatial Distribution, σ(r, θ, φ)
Disk models were created from the orbital element distribution of §5.1.1 using the "SIMUL"
program; SIMUL was developed by the solar system dynamics group at the University of Florida
(Dermott et al. 1992). A disk model is a large three-dimensional array, σ(r, θ, φ), that describes
the spatial distribution of the cross-sectional area of material in the disk model per unit volume
binned in: r, the radial distance from the star; θ, the longitude relative to an arbitrary direction
(set here as the forced pericenter direction, ωf ); and φ, the latitude relative to an arbitrary plane
(set here as the forced inclination, yf , or symmetry, plane). SIMUL creates a disk model by taking
the total cross-sectional area of material in the disk (specified by the model variable σtot), and
dividing it equally among a large number of orbits (5 million in this case), the elements of each of
which are chosen randomly from the specified distribution (§5.1.1). The disk model is populated
by considering the contribution of each orbit to the cross-sectional area density in each of the cells
it crosses.
The spatial distribution of material in one of our models of the HR 4796 disk can be described
by the three variables amin, γ, and ef ; σtot simply scales the amount of material in the model,
and ωf and yf describe the orientation of the disk to our line of sight. Fig. 4 is a plot of the
surface density of material in a disk model with amin = 62 AU, ef = 0.02, and γ = −2 (this is
our final model of §5.4). This illustrates how the specified distribution of orbital elements affects
the spatial distribution of material in the disk model: the sharp cut-off in semimajor axes at
amin determines the radial location of the inner hole, which has a sloping cut-off in r due to
the particles' eccentricities; as predicted in §4.2.1, particles at the inner edge of the disk in the
forced pericenter direction are closer to the star than those in the forced apocenter direction by
-- 25 --
∼ 2aminef ; the distribution of semimajor axes has produced a surface density distribution that is
∝ rγ−1, but only exterior to 70 AU.
5.2. Model of P (λ, r)
5.2.1. Optical Properties of Disk Particles
The optical properties of the disk particles were found assuming the particles to be made
of astronomical silicate (Draine & Lee 1984; Laor & Draine 1993), a common component of
interplanetary dust found in both the zodiacal cloud (Leinert & Grun 1990) and exosolar systems
(e.g., Telesco & Knacke 1991; Fajardo-Acosta, Telesco, & Knacke 1993; Sitko et al. 1999); such
an assumption can be tested at a later date using spectroscopy to look for silicate features in the
HR 4796 disk emission. Furthermore, the particles were assumed to be solid, spherical, and have
a density of ρ = 2500kg/m3. Their optical properties were calculated using Mie theory, assuming
that HR 4796A has a luminosity and temperature of L⋆ = 21L⊙, T⋆ = 9500◦K (Jura et al. 1998),
and using for its spectrum, that of the A0V star Vega (Cohen 1999). In all calculations, the mass
of HR 4796A was assumed to be M⋆ = 2.5M⊙ (Jayawardhana et al. 1998).
The properties of particles of different sizes, and at different distances from HR 4796A, are
shown in Fig. 5. The temperatures of the particles are plotted in Figs. 5a and 5b. The form of
Fig. 5a can be understood by consideration of equation (1), and the wavelengths at which the
star and a particle, if it was a black body, emit most of their energy: λ⋆ ≈ 2898/T⋆ µm, and
λbb ≈ 2898/Tbb = 10pr/a⊕(L⊙/L⋆)0.25 µm. As a crude approximation, a particle with diameter D
has Qabs ≈ 1 for λ ≪ πD and Qabs → 0 for λ ≫ πD. Thus, in terms of their thermal properties,
disk particles can be divided into four categories: the largest particles, D ≫ λbb/π, are efficient
absorbers and emitters at all relevant wavelengths and so achieve nearly black body temperatures,
Tbb; particles with D ≪ λbb/π, are inefficient emitters at their black body temperature, and so
need temperatures higher than Tbb to re-radiate all of the incident energy; the smallest particles,
D ≪ λ⋆/π, are also inefficient absorbers at the stellar temperature, and so do not need as high
temperatures as slightly larger particles to re-radiate the absorbed energy; and particles with
D ≈ 20 µm have temperatures below that of a black body -- this is because these particles are
super-efficient emitters at their black body temperatures (due to silicate resonances, Qabs can go
up as high as 2), and so need lower temperatures to re-radiate the incident energy. The form of
Fig. 5b can be understood in the same way. The fall-off of a large (e.g., D = 1000 µm) particle's
temperature with distance from HR 4796A is like that of a black body, i.e., T ∝ 1/√r, while the
fall-off for smaller particles is not that steep because these particles emit less efficiently the further
they are from the star (due to their lower temperatures, and consequently higher λbb); e.g., the
fall-off for D = 2.5 µm particles is T ∝ 1/r0.34, which is close to the 1/r1/3 fall-off expected for
particles with an emission efficiency that decreases ∝ 1/λ2 (e.g., Backman & Paresce 1993).
More important observationally is the variation of the particles' Qabs(λ, D)Bν [λ, T (D, r)],
-- 26 --
since this determines the contribution of a particle's thermal emission to the flux density received
at the Earth (eqs. [3],[7] and [8]). This is plotted for λ = 18.2 µm in Figs. 5c and 5d, and for
λ = 10.8 µm in Figs. 5e and 5f. The form of Figs. 5c and 5e can be explained in the same way
that Fig. 5a was explained: all three figures have similar forms, which is to be expected since the
particles' temperature also appears in Bν; Figs. 5c and 5e are, however, attenuated for D ≪ λ/π,
since these particles are inefficient emitters at that wavelength. The fall-off with distance of the
different particles shown in Figs. 5d and 5f is due solely to their different temperature fall-offs
(Fig. 5b); e.g., for λ = 18.2 µm, the approximate fall-off for D = 1000 µm particles is ∝ 1/r5.4,
while that for D = 2.5 µm particles is ∝ 1/r2.6.
5.2.2. Cross-sectional Area Distribution
The definition of P (λ, r) (eqs. [8] and [9]) shows that it is the convolution of QabsBν (Fig. 5c -
5f) with the cross-sectional area distribution, ¯σ(D, r). Since theoretical arguments cannot supply
an accurate size distribution (§3.2.2), we use the assumption of equation (10), which is that P (λ, r)
is equal to the QabsBν of particles in the disk with characteristic size, Dtyp. We further assume
that this Dtyp is constant across the disk in the IHW18 waveband; i.e., the brightness of a disk
model observation in the IHW18 waveband is calculated using P (λ, r) from the line on Fig. 5c
corresponding to Dtyp, where Dtyp is a model variable. A qualitative understanding of Dtyp comes
from Fig. 5c: since QabsBν is fairly flat for particles larger than ∼ 8 µm, an IHW18 waveband
observation is dominated by those particles with the most cross-sectional area, unless there are
a significant amount of particles smaller than 8 µm; Fig. 5c also shows that the observation is
unlikely to be dominated by particles smaller than O(0.01 µm), unless their contribution to the
disk's cross-sectional area is much higher than that of larger particles.
In addition to the IHW18 waveband, T99 observed HR 4796 in the N (λ = 10.2 µm)
waveband. The N band observations also show the double-lobed feature, but the inaccuracy of
the subtraction of the image of HR 4796A from the observations means that they cannot be used
to constrain the disk's structure. The N band observations can, however, be used to constrain the
disk's brightness in this waveband; the brightness of a disk in different wavebands differs only in
the factor P (λ, r) (eq. [7]), and so the N band observation can be used to obtain information about
the disk's size distribution. Since Figs. 5c and 5e have similar forms, the two observations should
be dominated by the emission of similarly sized particles; i.e., they should have the same Dtyp,
and show similar structures. The same characteristic particle size, Dtyp, was used to calculate
the brightness of a disk model in both the IHW18 and N wavebands, using P (λ, r) from the
appropriate lines on the plots of Figs. 5d and 5f. There is just one Dtyp that simultaneously
matches the disk's observed brightnesses in both wavebands.
We made an initial estimate for Dtyp based on the flux densities observed in the two
wavebands: the flux densities of the disk in the IHW18 and N wavebands are 857 and 40 mJy,
respectively (T99); i.e., the observed flux density ratio (N/IHW18) is O(0.05). The expected flux
-- 27 --
density ratio of the two wavebands (the ratio of P (λ, r) for different Dtyp) is plotted in Fig. 6a.
Assuming the emission to arise mostly from particles near the inner edge of the disk, r = 60-80
AU, Fig. 6a shows that the observed emission can be fairly well-constrained to come from particles
with Dtyp = 2 − 3 µm.
5.2.3. Pericenter Glow
Fig. 7a shows a contour plot of an unsmoothed IHW18 waveband observation of the disk
model of Fig. 4 viewed face-on (i.e., perpendicular to the disk's plane of symmetry, yf ). This
shows the observational consequence of the offset center of symmetry of the disk model. Because
the particles at the inner edge of the disk (those that contribute most to the disk's brightness) are
closer to the star in the forced pericenter direction, ωf , than those in the forced apocenter direction
(Fig. 4), they are hotter and so contribute more to the disk's thermal emission than those at the
forced apocenter. This is the "pericenter glow" phenomenon, which leads to the horseshoe-shaped
highest contour line (the filled-in 1.02 mJy/pixel line), which is pointed in the ωf direction. This
asymmetry is a consequence of amin and ef only, and its magnitude is determined by ef only. In
particular, if there is a gradient of ef across the disk, then it is ef at the inner edge of the disk
that controls the magnitude of the asymmetry. The outermost contour plotted on Fig. 7a, which
is an offset circle with a radius of 95 AU, is that corresponding to 0.17 mJy/pixel. Thus, there is
little emission from the outer edge of the disk, justifying the arbitrary use of amax = 130 AU in
the modeling.
5.3. Disk Model Orientation
The two variables that define the orientation of the HR 4796 disk to our line of sight are
ωf and Iobs; how they define this orientation is best explained using Fig. 7a. Imagine that the
disk starts face-on with the forced pericenter direction pointing to the left. It is then rotated
clockwise by ωf (this is shown in Fig. 7a, where ωf = 26◦), and then tilted by 90◦ − Iobs about the
dotted line on Fig. 7a. The direction of this tilt, whether the top or bottom of the disk ends up
closer to the observer, is not constrained in the modeling, since no account was made for either
the extinction of the disk's emission by the disk itself, or for the disk's scattered light (e.g., an
observer would see forward-scattered starlight from the closest part of the disk and back-scattered
starlight from the farthest part, a phenomenon that could produce an apparent asymmetry in a
symmetric disk, Kalas & Jewitt 1995). If the resulting inclination of the disk's symmetry plane to
our line of sight, Iobs, is small, then the resulting nearly edge-on observation shows two lobes, one
either side of the star. Since the hotter, brighter, pericenter glow material is predominantly in one
of the lobes (unless ωf = 90◦), the lobes have asymmetric brightnesses.
-- 28 --
5.4. Modeling Process and Results
Pseudo-observations of disk models in the IHW18 and N wavebands were produced that
mimicked the real OSCIR (the University of Florida mid-IR imager) observations in both pixel
size, 1 pixel = 0.′′0616 = 4.133 AU at 67 pc (Jura et al. 1998), and smoothing, using the observed
PSFs (which are asymmetric and slightly fatter than diffraction limited, T99), and including the
post-observational gaussian smoothing of FWHM = 3 pixels. The model variables: amin, γ, Iobs,
ef , ωf , and σtot, were optimized so that the modeled IHW18 observation correctly predicts the
observed IHW18 brightness distribution; at the same time, the variable, Dtyp, was optimized so
that the modeled N band observation correctly predicts the observed N band brightness.
The model observations were compared with the real observations using the following
diagnostics: the lobe brightnesses, Fne,sw, and their projected radial offsets from HR 4796A,
Rne,sw, that were found by fitting a quintic polynomial surface to a 10 × 10 pixel region around
each of the lobes with 0.1 pixel resolution (the location of HR 4796A in the real observations was
also found in this way); and line-cuts through the disk both parallel (e.g., Figs. 8d and 9), and
perpendicular (e.g., Figs. 8c and 8e) to the line joining the two lobes in the IHW18 observations.
An understanding of how the different model variables affect the different diagnostics allowed
the modeling process to be decoupled into solving for: the disk's symmetrical structure, defined
by Iobs, amin, and γ; the particle size Dtyp; and the disk's asymmetrical structure, defined by ef
and ωf . Throughout the modeling, the amount of material in a model, σtot, was scaled so that
the model observation predicted the correct observed mean brightness of the lobes in the IHW18
waveband, Fmean = (Fne + Fsw)/2 = 1.40 ± 0.02 mJy/pixel; its final value, σtot = 2.03 × 1024 m2,
was calculated once the other variables had been constrained.
5.4.1. Symmetrical Disk Structure
Since ef and ωf only pertain to the disk's asymmetrical structure, then for a given Dtyp, the
variables pertaining to the disk's symmetrical structure, Iobs, amin, and γ, could be solved using
a model with ef = 0; such a model is axisymmetric, and so the variable ωf is redundant. The
inclination of the disk's plane of symmetry to the line of sight, Iobs = 13 ± 1◦, was constrained
to give the best fit to the line-cuts perpendicular to the lobes (Figs. 8c and 8e); this is in
agreement with that found by previous models of HR 4796 disk observations (Koerner et al. 1998;
Schneider et al. 1999). Since our model is that of a "fat" disk, a different proper inclination
distribution would lead to a different Iobs; e.g., if the disk is actually thinner than modeled here,
hIpi < 10.2◦, then the inferred Iobs is an underestimate, and vice versa. The inner edge of the
disk, amin = 62 ± 2 AU, was constrained such that the model observation reproduces the observed
mean radial offset of the lobes from HR 4796A, Rmean = (Rne + Rsw)/2 = 58.1 ± 1.3 AU. The
semimajor axis distribution, γ = −2 ± 1, was constrained to give the best fit to the cut along the
line joining the two lobes (Fig. 8d).
-- 29 --
5.4.2. Particle Size
Since Dtyp was already estimated to be about 2-3 µm (§5.2.2), the modeling of the disk's
symmetrical structure was repeated for Dtyp = 2, 2.5, and 3 µm. Adjusting Dtyp by such a small
amount did not affect the inferred symmetrical structure parameters. This was expected, since
the P (18.2 µm, r) for each of these Dtyp are very similar (Fig. 5d). Remembering that the model
is always normalized to predict the observed mean IHW18 lobe brightnesses, the predicted N
band lobe brightnesses are compared for the three values of Dtyp in Fig. 9. This shows that the
particle size can be constrained to be Dtyp = 2.5 ± 0.5 µm; i.e., the crude method of calculating
the particle size of §5.2.2 gives a very good estimate of this size. This particle size means that the
total mass of emitting particles in the disk model is ∼ 1.4 × 10−3M⊕, where M⊕ = 3 × 10−6M⊙ is
the mass of the Earth. However, this is not a useful constraint on the disk's mass, since the disk's
mass is expected to be concentrated in its largest particles. The best estimate of the mass of the
HR 4796 disk, from submillimeter observations, is that it is between 0.1M⊕ and 1.0M⊕ (Jura et
al. 1995; Jura et al. 1998), which, as expected, is well above the mass of our disk model.
5.4.3. Asymmetrical Disk Structure
The disk's observed asymmetries are defined by: the lobe brightness asymmetry,
(Fne− Fsw)/Fmean = 5.1± 3.2%, and the radial offset asymmetry, (Rsw− Rne)/Rmean = 6.4± 4.6%.
These asymmetries are also apparent in the disk model, and their magnitudes are determined by
both ef and ωf . The lobe brightness asymmetry was used to constrain ef and ωf , and it was
found that the ef necessary to cause the 5.1 ± 3.2% asymmetry depends on the geometry of the
observation according to the relation shown in Fig. 7b. Thus, for the majority of the geometries,
a forced eccentricity of between 0.02 and 0.03 is sufficient to cause the observed brightness
asymmetry. In the context of this modeling, the observed brightness asymmetry implies a radial
offset asymmetry of ∼ 5%, which is within the limits of the observation. Our final model shown in
Figs. 4 and 7-9 assumes a modest value of ef = 0.02, which corresponds to an ωf orientation of 26◦
(Fig. 7b). Whether this rotation puts the pericenter glow material above or below the horizontal,
ωf = 26◦ or −26◦, is not constrained here, since it has a minimal effect on the observation: in the
model observation of Fig. 8, the top of the disk is brighter than the bottom of the disk by a fraction
that would be undetectable in the observation due to noise and the disk's unknown residual
structure. The line-cuts of Fig. 8 show how well the model fits all aspects of the observation --
the vertical structure, the horizontal structure, and the lobe location and asymmetry.
5.4.4. Statistical Significance
The standard deviations of the OSCIR lobe observations quoted in this section were found
using model observations that mimicked the noise present in the OSCIR observations. The
-- 30 --
background sky noise in the IHW18 observation was found to be approximately gaussian with
zero mean and a 1σ noise per pixel of 0.15 mJy (T99); this was included after the PSF smoothing,
but before the post-observational smoothing. Observations of the model were repeated for 50,000
different noise fields to obtain the quoted standard deviations. Since the observed PSF was
asymmetric (T99), this introduces an apparent lobe asymmetry of −0.8% in an observation of
a symmetric disk (one with ef = 0), and so the observed lobe asymmetry is 5.9 ± 3.2% from
the mean, and its statistical significance is 1.8σ. While this is small, it does show that the
pericenter glow phenomenon is observable with current technology: HR 4796 was observed with
the infrared imager OSCIR for only one hour on Keck II (T99); in the background-limited regime,
the significance level of any asymmetry increases at a rate ∝ √t; thus, one good night on a 10
meter telescope should be enough to get a definitive observation of the HR 4796 lobe asymmetry.
The real significance of the HR 4796 asymmetry may also be higher than quoted above, since it
also seems to be apparent in other observations of this disk (Koerner et al. 1998; Schneider et al.
1999). Even if subsequent observations happen to disprove the existence of the asymmetry, this is
still significant, since, as we show in §6.2, an asymmetry is to be expected if the companion star,
HR 4796B, is on an eccentric orbit.
6.
Interpretation of the HR 4796 Disk Model
The interpretation of our HR 4796 disk model is broken down into sections that cover
discussions of: the dynamics of the disk particles, §6.1; the lobe asymmetry, §6.2; the emitting
particle category, §6.3; the origin of the inner hole, §6.4; and the residual structure of the
observation once the model has been subtracted, §6.5.
6.1. The Dynamic HR 4796 Disk
Our interpretation of the observed structure of the HR 4796 disk starts with a discussion of the
dynamics of the particles in the disk, and where the emitting particles fit into our understanding
of the dynamic disk (§3).
6.1.1. Radiation Forces, β
The radiation forces, defined by β (eq. [17]), acting on particles in the HR 4796 disk can be
found from their optical properties. Fig. 6b shows the β of particles with the optical properties
assumed in our model (§5.2.1). Thus, particles in the disk with D < 8 µm are β meteoroids (even
the submicron particles have β > 0.5), and a good approximation for the non-β meteoroids is:
β ≈ 4/D,
(43)
-- 31 --
where D is measured in µm (in agreement with eq. [18]). Fig. 9 shows that the emitting particles
have Dtyp ≈ 2.5 µm; these particles have β = 1.7. So, the modeling implies that the emitting
particles are β meteoroids, i.e., that they are blown out of the system on hyperbolic orbits. Since
the lifetime of β meteoroids is O(370 years) for those created at ∼ 70 AU, which is much shorter
than the age of the HR 4796 system, any β meteoroids that are currently in the disk cannot be
primordial particles, rather they must be continuously created from a reservoir of larger particles;
i.e., the existence of β meteoroids implies the existence of a dynamically stable population of
larger particles.
If the emitting particles are on hyperbolic orbits, we can use the mass of the disk's emitting
particles, ∼ 1.4× 10−3M⊕, and the emitting lifetime of these particles, ∼ 370 years, to estimate the
mass loss rate of the disk to be ∼ 4 × 10−6M⊕/year. If this mass loss rate has been sustained over
the age of the system, the original disk must have been ∼ 40M⊕ more massive than it is today.
In fact, a more massive disk would have had a higher mass loss rate, since this rate increases
proportionally with the total cross-sectional area in the disk (i.e., ∝ m2/3
tot ). This means that if the
current disk has a mass 1.0M⊕, the original disk must have had a mass ∼ 7 × 104M⊕; i.e., the HR
4796 disk may provide evidence for the type of collisional mass loss that may have happened in
the early Kuiper belt (Stern & Colwell 1997).
6.1.2. Collisional Processes
The collisional lifetime of the disk's emitting particles can be calculated directly from the
disk model using equations (A7) and (A12): tcoll(Dtyp) < 104 years across most of the disk (55-85
AU), with a minimum at ∼ 70 AU of ∼ 4500 years. This collisional lifetime is much less than
the age of the HR 4796 system. Thus, the emitting particles cannot be primordial particles
(irrespective of whether they are β meteoroids). The collisional lifetime of the emitting particles
can be corroborated using equation (16). The model's effective optical depth (eq. [A8]) is its
surface density, plotted in Fig. 4, multiplied by the cross-sectional area of a particle in the model,
typ/4. This peaks at 70 AU, where τef f (70 AU) = 5 × 10−3, and tper = O(370 years),
σ = πD2
giving a collisional lifetime that has a minimum of tcoll(Dtyp) ≈ 6000 years. The disk's effective
optical depth at 70 AU, τef f (70 AU), can also be corroborated directly from the observation
using equation (A10). The observed edge-on, smoothed lobe brightness is ∼ 1.40 mJy/pixel;
this can be scaled to the unsmoothed face-on brightness by the factor 1.10/1.40 (see Figs. 7a
and 8b), which, since each pixel subtends (0.0616 ∗ π/648000)2 sr, gives a brightness of Fν (18.2
µm,70 AU)/Ωobs = 12 × 109 Jy/sr. For 2.5 µm particles, Figs. 5c and 5d give P (18.2 µm,70
AU) ≈ 2.2 × 1012 Jy/sr, thus confirming that τef f (70 AU) ≈ 5 × 10−3 (this is also in agreement
with τ ≈ 5 × 10−3 found by Jura et al. 1995). In fact, assuming that the disk's total IHW18
flux density, 857 mJy (T99), comes from particles between 70 ± 15 AU, the unsmoothed face-on
brightness of the disk can also be corroborated; equation (A11) with R⋆ = 67 pc gives Fν(18.2
µm, 70 AU)/Ωobs = O(1010 Jy/sr).
-- 32 --
Since particles are only broken up by collisions with particles that have diameters more than
a tenth of their own (eq. [A2]), the collisional lifetime of the disk's large particles must be longer
than that of the smaller emitting particles. Assuming the cross-sectional area distribution to follow
equation (15) with q = 11/6 down to particles of size Dtyp, the collisional lifetime of particles with
D ≫ Dtyp can be estimated as (eq. [A15]):
tcoll ≈ 4 × 107√D,
(44)
where tcoll is measured in years, and D in km. Since particles for which tcoll < tsys cannot be
primordial, this implies that particles currently in the HR 4796 disk that are smaller than 60
m cannot be original, rather they must have been created in the break-up of larger particles;
i.e., these disk particles form a collisional cascade, through which the spatial distributions of the
smaller particles are related to that of the larger particles. The disk's particles that are larger
than 60 m are its primordial particles.
6.1.3. P-R Drag
The disk's high effective optical depth means that βpr = 132 at 70 AU (eq. [26]), which in
turn means that there is not expected to be significant P-R drag evolution for any of the disk's
particles. Analysis of the P-R drag evolution of disk particles (eq. [25]) shows that even the most
affected particles, those with β = 0.5, have a P-R drag lifetime of tpr = O(1.6 Myr) at 70 AU,
and that these particles would only make it to ∼ 69.8 AU before they are broken up by collisions
and blown out of the system by radiation pressure. Appendix B, which gives a discussion of
lobe asymmetries in disks for which P-R drag is important, gives an analysis that shows that the
cumulative effect of P-R drag on the orbits of all of a disk particle's ancestors is also insignificant.
6.1.4. The Dynamic HR 4796 Disk
Since P-R drag is not an important process in the HR 4796 disk's evolution, there are just
three categories of particles in the disk: large particles, β critical particles, and β meteoroids. If
we are to believe the modeled emitting particle size (discussed further in §6.3), the particles that
are seen in both the IHW18 and the N band observations are the disk's β meteoroids. Since we
modeled the disk as if the emitting particles are large particles, this inconsistency needs to be
borne in mind in our interpretation of the model.
While the modeling used a distribution of orbital elements that is only appropriate for the
disk's large particles, it was the spatial distribution of the disk's emitting material, σ(r, θ, φ),
that was constrained by the modeling, not the distribution of orbital elements, σ(a, e, I, Ω, ω).
Therefore, the inferred distribution, σ(r, θ, φ), which is that shown in Fig. 4, is an accurate
description of the spatial distribution of the disk's emitting particles, whatever their size. Indeed,
-- 33 --
it is in excellent agreement with that inferred from other observations of the HR 4796 disk
(Jayawardhana et al. 1998; Koerner et al. 1998; Schneider et al. 1999). If the emitting particles
are the disk's large particles, then the inferred model variables have physical interpretations
for the distribution of the orbits of the disk's large particles. If, as appears to be the case, the
emitting particles are β meteoroids, then further modeling of σ(r, θ, φ) needs to be done to infer
the distribution of the orbits of these particles.
However, since the disk's β critical and β meteoroid particles are created from its large
particles, the spatial distributions of all of these particles share a great deal in common (see §§3.2
and 4.2): they all have the same plane of symmetry, the same flaring, and the same offset and warp
asymmetries, but the radial distributions of the smaller particles are more extended than that
of the large particles. This means that if we are seeing the disk's β meteoroids, then the spatial
distribution of its large particles has a plane of symmetry that is defined by Iobs, and an inner
edge that is at the same radial location, and that is offset by the same amount and in the same
direction, as that shown in Fig. 4; their radial distribution, however, would not be as extended
as that of Fig. 4. Thus, the model parameters amin, ef , and ωf have physical interpretations for
the distribution of the orbits of the disk's large particles, irrespective of the size of the emitting
particles.
6.2.
Interpretation of Lobe Asymmetry: HR 4796's Secular Perturbations
The lobe asymmetry in the model of §5 is due solely to the offset inner edge of the disk. The
model shows that secular perturbations amounting to a comparatively small forced eccentricity,
ef = 0.02, imposed on the orbits of large particles at the inner edge of the disk, a = 62 AU, would
cause the disk's inner edge to be offset by a sufficient amount to cause the observed 5% lobe
asymmetry. This section considers what kind of a perturber system would impose such a forced
eccentricity on the inner edge of the disk, and whether such a system is physically realistic. A
discussion of the system's secular perturbations also allows interpretation of the disk's inferred
orientation, defined by the parameters ωf and Iobs.
If HR 4796A's binary companion, HR 4796B, is on an eccentric orbit, it would have imposed
a forced eccentricity on the disk particles. However, a forced eccentricity could also have been
imposed on the disk by an unseen planet close to the inner edge of the disk, a planet which could
be responsible for clearing the inner region (e.g., Roques et al. 1994). The secular perturbations
imposed on the HR 4796 disk by a perturber system that includes HR 4796B and a putative
planet located at the inner edge of the disk are shown in Fig. 10 for the four cases: Mpl = 0,
Mpl = 0.1MJ , Mpl = 10MJ , and just planet. The parameters of the two perturbers are assumed
to be:
HR 4796B MB = 0.38M⊙(Jayawardhana et al. 1998); the orbit of HR 4796B is unknown
at present (Jura et al. 1993), so the semimajor axis of its orbit is arbitrarily taken as its
-- 34 --
projected distance, aB = 517 AU (Jura et al. 1998; note that we are not assuming that this
is the semimajor axis of the ellipse that the star's orbit traces on the sky) -- this gives an
orbital period of ∼ 7000 years (eq. [12]), and a timescale for secular perturbations from HR
4796B to have built up at 62 AU of O(1 Myr) (eq. [42]); eB = 0.13, the eccentricity necessary
to cause ef = 0.02 at a = 62 AU if there were no unseen perturbers (eq. [41a]); IB = 0◦,
defining the reference plane for the analysis.
Planet Mpl is a variable measured in Jupiter masses, where MJ = 10−3M⊙ (current observations
have limited the size of a planet in the system to Mpl < 20MJ , Jura et al. 1998); apl = 47
AU (see §6.4); epl = 0.023, the eccentricity necessary to cause ef = 0.02 at a = 62 AU if the
planet was the only perturber (eq. [41a]); Ipl = 5◦, an arbitrary choice that represents the
fact that the orbital plane of the planet is not necessarily be aligned with that of HR 4796B.
6.2.1. Just HR 4796B
For the cases when there is just one perturber in the system, the forced elements in the
system can be found from equations (41a) and (41b): the forced eccentricity, ef , is determined by
the ratio of the semimajor axes of the perturber and the particle, and by the eccentricity of the
perturber's orbit, but is independent of the perturber's mass; the forced pericenter, ωf , is aligned
with the pericenter of the perturber; and the plane of symmetry of the disk, yf , is constant across
the disk, and is the orbital plane of the perturber.
So, if the only perturber is HR 4796B, then to impose ef = 0.02 at a = 62 AU, the eccentricity
of its orbit would have to be eB = 0.13; the consequent forced eccentricity imposed on the disk
is plotted in Fig. 10a. This also means that if eB > 0.1, then a brightness asymmetry in this
disk of > 5% would be expected unless adverse geometrical conditions prevented it. The position
angle from north of HR 4796B relative to HR 4796A is 225◦ (Jayawardhana et al. 1998), while
that of the SW lobe (i.e., the least bright lobe) is 206◦ (T99). For the lobe asymmetry to be
the consequence of perturbations from HR 4796B only, HR 4796B must currently be close to
its apastron, and its orbital plane must be the plane of symmetry of the disk, i.e., inclined at
13◦ to the line of sight. All of these conclusions are consistent with our initial estimate that the
semimajor axis of the star's orbit is equal to its observed projected distance, 517 AU.
6.2.2. HR 4796B and a Planet
Consider the effect of adding a planet at the inner edge of the disk into the HR 4796 system4.
If there are two perturbers in the system, then the forced element variation with semimajor
4The orbital elements of a low-mass planet in the system would, just like the disk particles, have forced and proper
components; a high-mass planet would perturb the orbit of HR 4796B.
-- 35 --
axis depends both on the masses of the perturbers and on the orientations of their orbits. In
the plot of Fig. 10b, ωpl = ωB + 180◦ was chosen so that the forced eccentricity (and hence
the lobe asymmetry) is aligned with the planet's pericenter for a < acrit, and aligned with HR
4796B's pericenter for a > acrit, where acrit is the semimajor axis for which ef = 0. Since the
two perturbers were also chosen to have different orbital planes, a similar change in the disk's
alignment is seen in the plot of If (Fig. 10d): the disk's plane of symmetry is aligned with the
planet's orbital plane at its inner edge, and with the orbital plane of HR 4796B at its outer edge;
this could cause an image of the disk to appear warped. Such a warp could be modeled using the
same modeling techniques that are described in this paper, and would provide further constraints
on the perturbers in the system (even if no warp was observed).
So, it is possible that the brightness asymmetry, and the symmetry plane of the lobes, are
determined by a planet close to the edge of the disk that has Mpl > 0.1MJ , rather than by HR
4796B. Using such an analysis, the pericenter glow phenomenon could be used to test for the
existence of a planet in the HR 4796 system, but only after the orbit of HR 4796B has been
determined; e.g., if ωB, eB or the plane of HR 4796B's orbit contradicted the observed asymmetry
orientation, brightness asymmetry magnitude, or the plane of symmetry of the lobes, then the
existence of a planet at the inner edge of the disk with Mpl > 0.1MJ could be inferred.
6.2.3. Just Planet
A double-lobed disk structure could also be observed in a system with no observable
companion. The only possible perturber in such a system is an unseen planet, the secular
perturbations of which warrant the same kind of discussion as for the case when HR 4796B was
the only perturber (§6.2.1). Depending on the planet's mass, radial location and eccentricity, it
too could give rise to a detectable pericenter glow. The only constraint on the planet's mass is
that the disk must be old enough for its secular perturbations to have affected the distribution of
orbital elements of the disk particles over the age of the system. Since it takes of the order of one
precession timescale to distribute the complex eccentricities of collisional fragments around the
circle centered on the forced eccentricity, the constraint on the planet's mass can be approximated
as that for which is that the age of the system, tsys, is greater than the secular timescale,
tsec ∝ 1/Mpl (eq. [42]); for the HR 4796 system this limit is Mpl > 10M⊕. The constraint on the
planet's eccentricity is even less stringent than for the binary companion because the planet is
closer to the edge of the disk: a planet in the HR 4796 system would only need an eccentricity of
epl > 0.02 to produce the observed 5% lobe asymmetry, and the forced eccentricity imposed on the
disk by a planet with epl = 0.023 is plotted in Fig. 10c. So, the signature of even a low-mass planet
would not escape detection and symmetrical double-lobed features are unlikely to be observed in
systems that contain planets.
-- 36 --
6.2.4. Other Considerations
If the disk itself is massive enough to cause significant gravitational perturbations to the
orbits of the disk particles, then the mass of the disk should be incorporated into the analysis of
the secular perturbations in the system. A massive disk could dampen the eccentricity of a planet
at the inner edge of the disk (Ward & Hahn 1998), thus reducing the offset asymmetry.
6.2.5. Fomalhaut
The Fomalhaut disk lobes may have asymmetric brightnesses (Holland et al. 1998), but the
statistical significance of this asymmetry is low. Fomalhaut is a wide visual binary system. Gliese
879 is Fomalhaut's common proper motion companion (Barrado Y Navascu´es et al. 1997); the
two stars are separated by ∼ 2◦, which corresponds to a projected separation of O(55,000 AU)
at 7.7 pc. At such a distance, the forced eccentricity imposed on the disk by the binary star is
insignificant (eq. [41a]). A secular perturbation offset asymmetry in this disk would be expected
only if there is a planet in the disk that has a non-circular orbit.
6.3. Discussion of Emitting Particle Category
The emitting lifetime of the disk's β meteoroids, O(370 years), is less than the emitting
lifetime of their parents, O(104 years); equivalently, τef f < 0.1. Thus, our understanding of
the dynamic disk implies that the disk's cross-sectional area distribution should not contain a
significant amount of β meteoroids. Rather, since there are no P-R drag affected particles in the
disk, the disk's emission is expected to come from its β critical particles, and its smallest large
particles (§3.2.2). Not all disk particles have the same composition and morphology; even if this
were a close approximation, there is, as yet, no evidence to suggest whether the particle properties
chosen in our modeling (§5.2.1) are correct. Are we indeed seeing the disk's β meteoroid particles,
or did the assumptions of the modeling lead us to this conclusion?
Consider the initial crude estimate of the particle size (§5.2.2); this proved to be an accurate
method for estimating the particle size (§5.4.2). If different assumptions had been made about the
particles' properties (e.g., if the particles had been assumed to be made of ice) or morphologies
(e.g., if the particles had been assumed to be like the "bird's nest" structures of Gustafson 1994),
both Figs. 6a and 6b would be different, and different conclusions might have been drawn about
the β of the emitting particles. If a size distribution had been included in the modeling, this
would also have affected our conclusions. These are considerations that should be modeled before
any firm conclusions about the dynamics of the emitting particles can be reached. However,
since irrespective of their assumed properties large particles have black body temperatures and
brightness ratios similar to those of the D > 100 µm particles in Fig. 6b, a flexible interpretation
-- 37 --
of the observed brightness ratio is that the emitting particles must have temperatures that are
hotter than black body. Thus, the emitting particles are either small (e.g., the simple analysis
of §5.2.1 implies that D < 10 µm), in which case they are likely to be β meteoroids (e.g., 10
µm particles have β > 0.5 unless they have densities > 3000 kg/m3, eq. [18]), or they are large
particles that are made up of smaller hotter particles.
The assumptions about the particle properties in the model could also have affected our
conclusions about the collisional lifetime of the emitting particles. Consider the estimate of τef f (70
AU) derived from the IHW18 lobe brightness (§6.1.2). Changing the properties of particles in the
model would change the estimate of τef f because of the resultant changes in P (18.2 µm, r); e.g., if
we had modeled the disk using 30-50 µm particles, we would have had to put more cross-sectional
area in the model for it to give the observed lobe brightness. Fig. 5c shows that for astronomical
silicate Mie spheres that are larger than 0.01 µm in diameter, P (18.2 µm,70 AU) must lie in the
range 0.034− 2.3× 1012 Jy/sr, depending on whether the disk's cross-sectional area is concentrated
in its 30-50 µm particles that emit at cool temperatures, or in its 2-3 µm particles that are small
and hot, but large enough to emit efficiently at 18.2 µm. For particles with different optical
properties, P (18.2 µm,70 AU) could be below 0.034× 1012 Jy/sr if the particles have temperatures
well below black body (or if they have low emission efficiencies); equally, it could be higher than
2.3 × 1012 Jy/sr if the particles are hotter than the 2-3 µm astronomical silicate Mie spheres.
Taking 0.034 × 1012 Jy/sr as a lower limit for P (18.2 µm,70 AU) implies that τef f (70 AU) < 0.35,
giving a collisional lifetime for the emitting particles that could be as low as 85 years. However,
given the temperature of the emitting particles inferred from the observed brightness ratio, O(104
years) remains the best estimate for their collisional lifetime. A further note of caution is necessary
about the inferred collisional lifetime: Appendix A assumes collisions between disk particles to
be either catastrophic or irrelevant. While this may be appropriate for the disk's larger particles,
since these are likely to be solid bodies, collisions between its smaller particles, which may have
fluffy "bird's nest" structures (Gustafson 1994), could be more erosive than destructive, and could
lead to significant grain growth.
In conclusion, neither observational, nor theoretical considerations can provide a definitive
answer as to the dynamics of the emitting particles. However, the confirmation of the emitting
particles' collisional lifetimes means that we can be sure that these particles are not primordial,
and that there are no P-R drag affected particles in the disk. β meteoroids remain the most likely
candidate for the emitting particles. Mid-IR emitting particles that are on hyperbolic orbits have
been inferred from observations of the disks around both β Pictoris (Telesco et al. 1988) and HD
141569 (Fisher et al. 1999).
6.4. Origin of the Inner Hole
Whatever the size of the emitting particles, analysis of the optical depth of the disk's inner
region (T99) shows that it is a few hundred times less than that of the outer disk, and so there
-- 38 --
are very few emitting particles in this region. Because this central hole is necessary for the secular
perturbation offset asymmetry to be observed (without the hole only the radial offset could be
observed), its physical origin requires attention. Since the existence of small emitting particles in
the disk requires the existence of large particles, the question to answer is why there are so many
large particles in the outer disk, but so few in the inner disk? Either the physical conditions were
such that they were able to form in the outer region, but not in the inner region, or they formed
across the whole disk, but those formed in the inner region have since been removed. Rather
than discussing the planetary formation process and the stage of the system's evolution (although
these are of utmost importance in determining the physics of the disk), this section offers possible
dynamical explanations for the removal of the large particles.
If a planet (or planets) exists interior to the inner edge of the disk, then resonance overlap
removes all material from a region of radial width ∼ 1.3apl(Mpl/M⋆)2/7 about the planet's orbit
within about 1000 orbital periods (Wisdom 1980; Duncan, Quinn, & Tremaine 1989). Material
is also removed from the secular resonance regions (Lecar & Franklin 1997) -- the origin of
these regions, which cover the range of semimajor axes on the plots of Figs. 10b and 10d where
ef , If → ∞, was discussed in §4.1.2. The radial distribution of material would also be affected by
planetary radial migration (Malhotra 1995; Trilling et al. 1998). However, such mechanisms can
only explain the total lack of large particles in the inner region by invoking a system with either
many planets, or just one planet that is either very large, or on a very eccentric orbit. Since all of
these mechanisms take longer than a particle's orbital period to take effect, they would cause an
inner cut-off in the disk particles' semimajor axes (as opposed to a cut-off in radial distance from
the star).
To estimate the orbit of a putative planet at the inner edge of the HR 4796 disk that is causing
the cut-off, consider the inner edge of the Kuiper belt. There is almost no Kuiper belt material
on orbits with semimajor axes interior to that of Pluto (Jewitt 1999), which is in 2:3 resonance
with Neptune. This is supposed to be the result of resonance sweeping that occurred as Neptune's
orbit expanded early in the history of the solar system due to the clearing of planetesimal debris
from the inner solar system and the formation of the Oort cloud (Malhotra 1995). By analogy,
assuming that the inner cut-off of the disk's large particles occurs at the planet's 2:3 resonance
location, and that this cut-off can be described by amin = 62 AU, we can estimate the orbit of the
planet to have a semimajor axis of apl = amin[2/3]2/3 = 47 AU, giving an orbital period of ∼ 200
years (eq. [12]).
6.5.
Interpretation of the Residual Structure
So far, no explanation has been offered for the structure of the residuals (what is left after
subtracting the model from the observation, see T99). Analogy with the zodiacal cloud implies
that there could be a population of warmer dust in the inner region that may be unrelated to the
dust in the outer disk. Depending on the perturbers in the inner region, such dust could contain
-- 39 --
considerable structure. Analysis of emission from such regions would reveal a great deal about the
system's perturbers, however, this would not be easy, since the resolution required to map such
small-scale structure is at the limit of current technological capabilities. In addition, such emission
is masked by that of both the stellar photosphere and the outer disk, the accurate subtractions of
which are difficult.
There may also be residual structure associated with the outer regions of the disk. If there is a
planet orbiting HR 4796A close to the inner edge of the disk, then the distribution of large particles
in the outer disk would contain structure associated with the planet's gravitational perturbations
in addition to the secular perturbations already discussed (Dermott et al. 1994; Malhotra 1996;
Dermott et al. 1998; Ward & Hahn 1998). Some of the emitting particles might be trapped in its
resonant ring. Such a ring could be responsible for some of the observed lobe asymmetry. The
existence of such a ring would give the inner edge of the disk structure that co-orbits with the
planet; i.e., observations of this structure would vary on timescales of ∼ 200 years, offering a
method of distinguishing between this structure and the large-scale background structure, which
would vary on secular timescales of O(1 Myr). The asymmetric structure of the Earth's resonant
ring includes a 0.23 AU3 cloud of dust located permanently in the Earth's wake with a number
density ∼ 10% above the background (Dermott et al. 1994). Observations of such an asymmetric
structure in an exosolar disk could be modeled using the same techniques that were used to model
the Earth's resonant ring (Dermott et al. 1994; Jayaraman & Dermott 1999), possibly allowing us
to determine the presence, location, and even the mass of the perturbing planet (Dermott et al.
1998). However, the evidence suggests that such observations may not be possible with currently
available technology. In fact, calculations show that when viewed from a distant point in space
normal to the ecliptic plane, the Earth's "wake" would only have an IR signal O(0.1) times that
of the Earth (Backman 1998). So, regardless of the resolution requirements, if one were to observe
the solar system from outside, it would be easier to detect the Earth directly than to infer its
existence from the structure of the zodiacal cloud.
The Earth's resonant ring is a result of the trapping of particles that are evolving into the
inner solar system due to P-R drag. Another method of forming a resonant ring is for a planet to
undergo radial migration of its orbit, trapping all particles exterior to its orbit into its strongest
resonances; such an interaction is supposed to have happened between Neptune and the Kuiper
belt (Malhotra 1995). The amount of disk material that is trapped in a Kuiper belt ring depends
on how much radial migration has taken place to the planets orbit. A ring with all of its particles
trapped in the 1:2 and 2:3 resonances with the planet (these are the strongest resonances) would
have three lobes, with the planet residing in an "empty" fourth lobe. Could this be the cause of
the tri-lobed structure observed in ǫ Eridani (Greaves et al. 1998)?
-- 40 --
7. Conclusions
The primary intent of this paper was to show how the long-term effect of the gravitational
perturbations, i.e., the secular perturbations, of a massive perturber could be the cause of the
∼ 5% brightness asymmetry of the double-lobed feature seen in observations of the HR 4796 disk
(T99):
1. We showed how the secular perturbations of a massive perturber in a disk impose a forced
eccentricity on the orbits of particles in the disk, thus causing the disk's center of symmetry
to be offset from the star towards the perturber's apastron. We also showed how the same
perturbations impose a forced inclination on the particles' orbits, which, if there is more
than one perturber in the disk, could cause the disk to be warped.
2. We produced a model of the HR 4796 disk that accurately maps the 18.2 µm brightness
distribution observed by T99; this model is based on a consideration of the dynamics of the
particles in the disk. The model shows how the brightness of a disk that has an inner clear
region, that also has an offset center of symmetry caused by a forced eccentricity imposed
on the disk particles' orbits, would be asymmetric, since the inner edge of one side of the
disk is closer to the star, and so is hotter and brighter, than the other side. We showed that
a forced eccentricity as small as 0.02 is all that is necessary to cause a 5% lobe brightness
asymmetry in the HR 4796 disk.
3. If the eccentricity of orbit of the companion star, HR 4796B, is larger than 0.13, then a
forced eccentricity of 0.02 is to be expected. However, if there is a planet of mass > 0.1MJ
located close to the inner edge of the disk, then the forced eccentricity, and hence the
asymmetry, imposed on material in the disk's lobes is controlled by the planet rather than
the binary companion; this could also cause the disk to be warped. If a forced eccentricity is
indeed the cause of the observed lobe asymmetry then observations that constrain the orbit
of HR 4796B would help to clarify whether such a planet exists. If the HR 4796 system had
no binary companion, a forced eccentricity of 0.02 could have been imposed on the disk by a
lone planet with a mass of > 10M⊕, and an eccentricity of > 0.02.
4. The statistical significance of the HR 4796 disk's lobe asymmetry in the observations of T99
is only at the 1.8σ level, however, it is also apparent in the observations of other authors
(Koerner et al. 1998; Schneider et al. 1999). It would take one good night on a 10 meter
telescope to get a clear observation of the HR 4796 asymmetry. Thus, the indirect detection
of planets, even small planets, hiding in circumstellar disks is clearly within reach using these
dynamic modeling techniques. This is particularly important, since the direct detection of
planets around even nearby stars is well beyond current capabilities (Backman 1998), and
indirect detection techniques such as radial velocity (Marcy & Butler 1998) or astrometric
(Gatewood 1996) techniques, permit detection only of very massive planets that are close to
the star.
-- 41 --
5. If there is a planet close to the inner edge of the disk, many of the disk's particles could
be trapped in resonance with that planet, thus forming a resonant ring. Such a ring would
give the inner edge of the disk structure that rotates on the timescale of the orbital period
of the planet, ∼ 200 years. This structure could be contributing to the observed lobe
asymmetry, and may also be present in the residuals of the observation. This possibility
could be explored with further observations. Resonant rings may be the predominantly
observable structures of some exosolar systems, such as that recently observed around ǫ
Eridani (Greaves et al. 1998).
The HR 4796 disk modeling also revealed important information about the large-scale
symmetrical structure of the disk, as well as about the dynamic properties of its emitting particles:
7. The spatial distribution of material in the disk inferred from our modeling matches that
inferred by other authors (Jayawardhana et al. 1998; Koerner et al. 1998; Schneider et al.
1999). The surface density of cross-sectional area in the disk peaks at ∼ 70 AU from HR
4796A, falling off ∝ r−3 outside this radius, dropping by a factor of ∼ 2 between 70 and 60
AU, and falling to zero by 45 AU. This soft inner edge to the disk is to be expected if the
disk particles' orbits are eccentric.
8. Assuming the particles to be astronomical silicate Mie spheres, the diameter of the emitting
particles was estimated to be Dtyp = 2 − 3 µm. Particles this small have radiation forces
that are characterized by β ≈ 2, and so are blown out of the system on hyperbolic orbits
on timescales of ∼ 370 years. The HR 4796 disk is very dense; the collisional lifetime of its
emitting particles is ∼ 104 years. Thus, the emitting particles cannot be primordial particles,
rather they must be continuously created from a reservoir of larger particles. The collisional
lifetimes of all of the disk's particles are shorter than their P-R drag lifetimes; i.e., none of
the disk's particles are affected by P-R drag. Further investigation of the particles' properties
needs to be done before any firm conclusions can be reached about whether the disk is dense
enough to support a population of particles on hyperbolic orbits that is large enough to
dominate the disk's emission. If the emitting particles are on hyperbolic orbits, the modeling
implies a mass loss rate for the disk ∼ 4 × 10−6M⊕/year, and further modeling would have
to be done to ascertain the spatial distribution of the dynamically stable population of large
particles from which these emitting particles originated.
A. Collisional Lifetime of Disk Particles
Consider a collision between two disk particles, the larger of which is denoted by the subscript
1, and the smaller by the subscript 2. For this collision to be "catastrophic", that is, for it to
result in the break-up of the larger particle, the impact energy of the collision must be large
enough both to overcome the tensile strength of the larger particle, and to impart enough energy
-- 42 --
to the collisional fragments to overcome its gravitational binding energy. In the asteroid belt this
limit means that a collision is only catastrophic if m2/m1 ≥ 10−4 (Dohnanyi 1969). Since the
impact energy of a collision is ∝ m2v2
rel, assuming that exosolar disk particles have similar tensile
strengths to the solar system's asteroids, this limit can be scaled to exosolar disks by the square of
the ratio of the mean relative velocity of collisions in the asteroid belt (at ∼ 3 AU), vrel ≈ 5 km/s
(Vedder 1998), to that of collisions in the exosolar disk. The mean relative velocity of collisions in
exosolar disks can be described by:
vrel(r) = f (e, I)v(r),
(A1)
where f (e, I) is some function of the disk particles' eccentricities and inclinations, and
v(r) = 30p(M⋆/M⊙)(a⊕/r) km/s is the average velocity of particles at r (eq. [14] with a replaced
by r). Thus, assuming f (e, I) ≈ 0.3 as for collisions in the asteroid belt, an exosolar disk particle of
diameter D ∝ m1/3, would only suffer a catastrophic collision if the other particle in the collision
had a diameter ≥ Dcc(D), where
Dcc(D) = 0.03[(M⊙/M⋆)(r/a⊕)]1/3D.
(A2)
The "collisional lifetime", i.e., the mean time between catastrophic collisions, of a particle
of diameter D, at a location in a disk denoted by r, θ, and φ, is the inverse of its catastrophic
collision rate (Kessler 1981):
where
tcoll(D, r, θ, φ) = [Rcoll(D, r, θ, φ)]−1,
Rcoll(D, r, θ, φ) = σcc(D, r, θ, φ)vrel(r),
(A3)
(A4)
σcc(D, r, θ, φ) is the catastrophic collision cross-section seen by the particle, and vrel(r) is the mean
encounter velocity of disk particles at r (eq. [A1]). Using the definition of a disk's structure given
by equation (4), this catastrophic collision cross-section is:
where
σcc(D, r, θ, φ) = fcc(D, r)σ(r, θ, φ),
fcc(D, r) = Z Dmax
Dcc(D)
(1 + D/D
′
′
)2 ¯σ(D
, r)dD
′
,
(A5)
(A6)
and Dcc(D) is the smallest particle with which a catastrophic collision could occur (eq. [A2]).
However, unless tcoll ≪ tper, the particle's orbit takes it through a range of θ and φ before
a collision occurs (there is also a variation of r along the particle's orbit due to the eccentricity
of its orbit). Thus, it is more appropriate to calculate the particle's collisional lifetime using
the mean catastrophic collision rate of the particles in the size range D ± dD/2 that are in the
spherical shell of radius, r, and width dr. Consider an element of this shell that has a volume,
dV = r2drdθ cos φdφ. The number of particles in the diameter range D ± dD/2 in this element is
-- 43 --
given by n(D, r, θ, φ)dDdV , and each of these particles has a catastrophic collision rate given by
equation (A4). Integrating over the whole shell gives:
tcoll(D, r) =
−Imax R 2π
R +Imax
0 σ(r, θ, φ)dθ cos φdφ
−Imax R 2π
R +Imax
0 [σ(r, θ, φ)]2dθ cos φdφ ∗ fcc(D, r)vrel(r)
,
(A7)
where Imax is the maximum inclination of the disk particles' orbits to the reference plane.
Equation (A7) can be simplified by considering a cylindrical shell, defined by r, θ, and z,
rather than a spherical one. An element of the cylindrical shell has a volume dV = rdrdθdz, and
the corresponding collisional lifetime of a particle in the shell is given by equation (A7), but with
φ, cos φdφ and ±Imax, replaced by z, dz, and ±h, where h = r sin Imax. Here we introduce the
parameter τef f , the disk's face-on effective optical depth:
τef f (r) = Z +h
−h
σ(r, θ, z)dz,
(A8)
where the dependence on θ has been dropped since orbits sample the full range of θ. This is not a
true optical depth, since that would include a consideration of the particles' extinction coefficients
(Qext = Qabs + Qsca); rather, it is the disk's face-on surface density of cross-sectional area, which is
equal to its true optical depth if its particles had Qext = 1. Assuming that σ(r, θ, z) is independent
of z, so that R +h
−h [σ(r, θ, z)]2dz = 0.5τef f (r)2/h, and that the encounter speed is determined by the
vertical motion of particles in the disk, so that f (e, I) ≈ sin Imax, equation (A7) can be simplified
to:
tcoll(D, r) =
,
(A9)
tper(r)
πfcc(D, r)τef f (r)
where tper(r) is the average orbital period of particles at r (eq. [12] with a replaced by r).
A disk's effective optical depth, τef f , can be estimated observationally from equation (6):
τef f (r) ≈ (Fν (λ, r)/Ωobs)/P (λ, r),
(A10)
where Fν /Ωobs is the disk's face-on unsmoothed brightness. The disk's face-on unsmoothed
brightness can be calculated either from the observed brightness, making corrections to account
for both the disk's orientation, as well as for the PSF smoothing, or from the disk's total flux
density, Fν (λ), and assuming the disk material to be evenly distributed between r ± dr/2:
Fν(λ, r)/Ωobs = Fν (λ)Cf R2
⋆/rdr,
(A11)
where Cf = 6.8 × 109 AU2/pc2/sr, and R⋆ is the distance of the star from the observer.
A.1. Collisional Lifetime of Particles with the Most Cross-sectional Area
Consider the particles in a disk that make up most of the disk's cross-sectional area, i.e.,
those with diameters close to Dtyp. By definition, these particles are most likely to collide with
-- 44 --
each other (a collision that would definitely be catastrophic), and so their collisional lifetime can
be found from equation (A7) using the approximation:
Applying this approximation to equation (A9) gives equation (16), which is also the collisional
lifetime determined by Artymowicz (1997) for particles in β Pictoris.
fcc(Dtyp, r) ≈ 4.
(A12)
A.2. Collisional Lifetime of a Disk's Large Particles
The collisional lifetime of particles of different sizes in a disk differ only by the factor fcc(D, r).
This factor can be ascertained by making assumptions about the disk particles' size distribution.
Assuming that the size distribution of equation (15) holds for disk particles between Dmin and
Dmax, the normalized cross-sectional area distribution is given by:
Substituting into equation (A6) gives:
¯σ(D) = (3q − 5)D4−3q/D5−3q
min .
fcc(D) = (cid:18) XD
Dmin(cid:19)5−3q(cid:20)1 +
6q − 10
(3q − 4)X
+
3q − 5
(3q − 3)X 2(cid:21) ,
(A13)
(A14)
where X = Dcc(D)/D for Dcc(D) > Dmin, and X = Dmin/D for Dcc(D) ≤ Dmin; the collisional
lifetime of particles in a disk with this distribution is a minimum for particles for which
Dcc(D) = Dmin. The size distribution of particles in a real disk is more complicated than equation
(A13), however, equation (A14) can be used to give a crude approximation for the collisional
lifetime of a disk's large particles:
tcoll(D, r) ≈ tcoll(Dtyp, r) ∗ (Dcc(D)/Dtyp)3q−5.
(A15)
A.3. Other Considerations for a Particle's Collisional Lifetime
If the P-R drag lifetime, tpr (eq. [25]), is comparable to, or shorter than tcoll, then the
effect of P-R drag on the collisional lifetime must be accounted for; e.g., on average, particles
dr = 1. Also, if
800(M⊙/M⋆)r
from a parent at rparent, survive until they reach rcoll where R rcoll
there is a significant change in r along a particle's orbit due to the eccentricity of its orbit, a
similar analysis can be done to take this into account when calculating its collisional lifetime.
By considering the washer-like disk of particles on orbits with a, e, and random ω (Sykes 1990):
tcoll(r)β
rparent
tcoll(a, e) = π[R q′
and q′ = a(1 + e) is its outer edge.
q (r/a)/[tcoll(r)p(r − q)(q′ − r)]dr]−1, where q = a(1 − e) is this disk's inner edge,
-- 45 --
B. Lobe Asymmetries in Disks where P-R Drag is Important
B.1. Offset and Warp of a Disk's P-R Drag Affected Particles
To find the secular evolution of the orbital elements of a particle that is affected by P-R drag,
i.e., one with βpr < β < 0.1, the equations governing the evolution of its complex eccentricity,
z = zsec − 2.5(α/a2)z (eqs. [24] and [37a]), and its complex inclination, y = ysec (eq. [37b]), must
both be solved in conjunction with the P-R drag evolution of its semimajor axis (eq. [23]). While
the solution given by equations (39a) and (39b) is no longer applicable, the decomposition of the
particle's complex eccentricity and complex inclination into forced and proper elements, and the
physical meaning of these elements, is still valid; however, each of these elements now depends on
the particle's dynamical history.
Consider the P-R drag affected particles created by the break-up of the asteroid family group
described in §4.2.1. Immediately after they are created, the orbital elements of these particles
are the same as those of the rest of the family; i.e., they have semimajor axes a, and complex
eccentricities and complex inclinations that are uniformly distributed in these planes around
circles of radii equal to the proper elements of the family, ep and Ip. The dynamical evolution
of a wave of these particles, i.e., those that were created at the same time, can be followed by
numerical integration to ascertain how the orbital elements of the particles in the wave vary as
their semimajor axes decrease due to P-R drag; this is the "particles in a circle" method (Dermott
et al. 1992). It was found that the complex eccentricities and complex inclinations of a wave of
particles originating in the asteroid belt remain on circles, and that as the wave's semimajor axis,
awave, decreases: its effective proper eccentricity (the radius of the wave's circle in the complex
eccentricity plane) decreases ∝ ep ∗ (awave/a)5/4; its effective proper inclination (the radius of the
wave's circle in the complex inclination plane) remains constant at Ip; the distributions of the
particles' ωp and Ωp remain random; while its effective forced elements (the centers of the circles
in the complex eccentricity and complex inclination planes) have a more complicated variation
(Dermott et al. 1992; Liou 1993).
Thus, the orbital element distributions, n(z) and n(y), of P-R drag affected particles are
like that of the large particles, in that they are the addition of zf and yf to symmetrical proper
element distributions; however, their forced and proper elements are different for particles from
different families, as well as being different for particles of different sizes and with different orbital
semimajor axes. This means that their spatial distribution is subject to offset and warp secular
perturbation asymmetries.
B.2. Origin of Inner Hole
For disks in which P-R drag is a significant physical process, the existence of an inner hole
implies not only that there must be no large particles in the inner region, but also that there
-- 46 --
must be some mechanism that prevents the particles that are in the outer disk from evolving into
the inner disk by P-R drag. In HR 4796, and other disks in which P-R drag is insignificant, this
mechanism is collisions; i.e., particles created in the outer disk are broken up by collisions before
they reach the inner region. This section presents dynamical explanations of how a planet located
at the inner edge of the disk would help to prevent the emitting particles from reaching the inner
disk.
This is not just important for disks in which its individual particles are not affected
significantly by P-R drag in their lifetime, since while the inner edge of a disk's largest particles
may be at amin, the cumulative effect of P-R drag on all of the stages of the collisional cascade
that a primordial particle goes through before the fragments are small enough to be blown out
of the system by radiation pressure could mean that the inner edge of these blow-out particles is
further in than amin. To assess whether the cumulative effect of P-R drag has an impact on the
inner edge of the emitting particles in the HR 4796 disk, consider its size distribution to follow
that assumed in equation (A13), where Dmin = Dtyp = 2.5 µm, and q = 11/6. The lifetime of an
intermediate particle can be found using equations (A9) and (A14), and the amount of P-R drag
evolution in its lifetime is (eqs. [23] and [43]):
da ≈ −0.0125tcoll(D, r)/Da,
(B1)
where da and a are measured in AU, D in µm, and tcoll(D, r) in years. Assuming that the largest
fragment created in a collision has a diameter half that of the original particle, a β meteoroid
particle (e.g., a 2.5 µm particle) at 62 AU, assuming it to have originated from a gravitationally
bound particle (i.e., one with β < 0.5), can at the very most be removed by 27 generations from
its primordial ancestor, which would at the very most have been originally at 62.1 AU.
One consequence of a planet at the inner edge of the disk is that it accretes some of the
particles that pass it on their way into the inner region. A simple estimate for the proportion
of particles lost in such a way can be obtained by considering the P-R drag evolution of a
torus of particles with orbital elements a, e, I, and random Ω, ω and λ; the volume of this
torus is Vtor = 8πa3e sin I (Sykes 1990). In the time it takes for the torus to pass the planet,
∆t = (1602/β)(M⊙/M⋆)(apl/a⊕)2e yrs, the planet accretes a volume of dust given by (Kessler
1981): Vacc = σcapvrel∆t, where σcap = πR2
rel) is the capture cross-sectional area, Rpl
pl(1 + v2
e /v2
is the radius of the planet, ve = q2GMpl/Rpl is the escape velocity of the planet, and vrel is the
mean relative velocity of encounter between the planet and the particles. Thus, the proportion of
dust accreted onto the planet, P = Vacc/Vtor, is given by:
P ≈
12
β√1 − β(cid:18) M⊙
M⋆(cid:19)3/2(cid:18) Mpl
M⊙(cid:19)4/3 a⊕
apl!1/2 ρJ
ρpl!1/3
g(e, I),
(B2)
where ρJ = 1330 kg/m3 is the mean density of Jupiter, g(e, I) = [(vrel/v) sin I]−1, and v is the
velocity of the particle; e.g., if g(e, I) = 100, then a Jupiter-like planet at the inner edge of the
disk would need Mpl > 30MJ to accrete all of the particles passing it (i.e., P = 1 for β < 0.5).
-- 47 --
Another consequence of a planet at the inner edge of the disk is that it traps some of the
disk particles into its exterior mean motion resonances. The resulting resonant ring has three
consequences that may aid with the formation of a clear inner region. Firstly, the calculation of the
probability of accretion onto the planet given in equation (B2) does not apply to the particles that
are trapped in the ring. It is thought that resonance trapping helps the accretion process, since
trapped particles may leave the resonance upon a close encounter with the planet (Kortenkamp
& Dermott 1998). This means that the constraint on the mass of the planet derived in the last
paragraph is an upper limit for the accretion process to be the sole removal mechanism. Secondly,
the ring increases the number of particles that are lost by collisional break-up, since it both
decreases the collisional lifetime of disk particles (by increasing the number density of particles in
the ring region) and increases the amount of time it takes for particles that are trapped in the
ring to reach the inner region. Finally, if the trapping timescales, tres, are longer than the age of
the system, tsys, then the trapped particles cannot have reached the inner disk yet; i.e., the ring
causes a bottle-neck in the flow of particles to the inner disk.
-- 48 --
REFERENCES
Artymowicz, P. 1997, Annu. Rev. Earth Planet. Sci., 25, 175
Artymowicz, P., & Clampin, M. 1997, ApJ, 490, 863
Backman, D. E., & Paresce, F. 1993, in Protostars and Planets III, 1253
Backman, D. E., Dasgupta, A., & Stencel, R. E. 1995, ApJ, 450, L35
Backman, D. E. 1998, in Exozodiacal Dust Workshop Conference Proceedings, eds. D. E. Backman
et al. (NASA/CP-1998-10155), 13
Barrado Y Navascu´es, D., Stauffer, J. R., Hartmann, L., & Balachandran, S. C. 1997, ApJ, 475,
313
Bowell, E. 1996, Asteroid Orbital Elements Database (Lowell Observatory)
Brouwer, D., & Clemence, G. M. 1961, Methods of Celestial Mechanics (New York: Academic
Press)
Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icarus, 40, 1
Cohen, M. 1999, private communication
Dermott, S. F., Nicholson, P. D., Burns, J. A., & Houck, J. R. 1984, Nature, 312, 505
Dermott, S. F., Nicholson, P. D., Burns, J. A., & Houck, J. R. 1985, in Properties and Interactions
of Interplanetary Dust, eds. R. H. Giese, & P. Lamy (Dordrecht: D. Reidel), 395
Dermott, S. F., & Nicholson, P. D. 1986, Nature, 319, 115
Dermott, S. F., Gomes, R. S., Durda, D. D., Gustafson, B. A. S., Jayaraman, S., Xu, Y. L., &
Nicholson, P. D. 1992, in Chaos, Resonance and Collective Dynamical Phenomena in the
Solar System, ed. S. Ferraz-Mello (IAU), 333
Dermott, S. F., Jayaraman, S., Xu, Y. L., Gustafson, B. A. S., & Liou, J. C. 1994, Nature, 369,
719
Dermott, S. F., Jayaraman, S., Xu, Y. L., Grogan, K., & Gustafson, B. A. S. 1996, in Unveiling
the Cosmic Infrared Background, ed. E. Dwek (New York: AIP), 25
Dermott, S. F., Grogan, K., Holmes, E. K., & Wyatt, M. C. 1998, in Exozodiacal Dust Workshop
Conference Proceedings, eds. D. E. Backman et al. (NASA/CP-1998-10155), 59
Dermott, S. F., Grogan, K., Holmes, E. K., & Kortenkamp, S. 1999, in Formation and Evolution
of Solids in Space, ed. J. M. Greenberg (Dordrecht: Kluwer), in press
Dohnanyi, J. S. 1969, J. Geophys. Res., 74, 2531
-- 49 --
Draine, B. T., & Lee, H. M. 1984, ApJ, 285, 89
Duncan, M., Quinn, T., & Tremaine, S. 1989, Icarus, 82, 402
Durda, D. D., & Dermott, S. F. 1997, Icarus, 130, 140
Durda, D. D., Greenberg, R., & Jedicke, R. 1998, Icarus, 135, 431
Fajardo-Acosta, S. B., Telesco, C. M., & Knacke, R. F. 1993, ApJ, 417, L33
Fisher, R. S., Telesco, C. M., Pina, R. K., Knacke, R. F., & Wyatt, M. C. 1999, ApJ, submitted
Gatewood, G. 1996, BAAS, 188, 4011
Gorkavyi, N. N., Ozernoy, L. M., Mather, J. C., & Taidakova, T. 1997, ApJ, 488, 268
Greaves, J. S., et al. 1998, ApJ, 506, L133
Grogan, K., Dermott, S. F., Jayaraman, S., & Xu, Y. L. 1997, Planet. Space Sci., 45, 1657
Grun, E., Zook, H. A., Fechtig, H., & Giese, R. H. 1985, Icarus, 62, 244
Gurnett, D. A., Ansher, J. A., Kurth, W. S., & Granroth, L. J. 1997, Geo. Res. Lett., 24, 3125
Gustafson, B. A. S. 1994, Annu. Rev. Earth Planet. Sci., 22, 553
Hirayama, K. 1918, AJ, 31, 185
Holland, W. S., et al. 1998, Nature, 392, 788
Jayaraman, S., & Dermott, S. F. 1999, Icarus, submitted
Jayawardhana, R., Fisher, R. S., Hartmann, L., Telesco, C. M., Pina, R. K., & Fazio, G. 1998,
ApJ, 503, L79
Jewitt, D. C. 1999, AREPS, in press
Jura, M., Zuckerman, B., Becklin, E. E., & Smith, R. C. 1993, ApJ, 418, L37
Jura, M., Ghez, A. M., White, R. J., McCarthy, D. W., Smith, R. C., & Martin, P. G. 1995, ApJ,
445, 451
Jura, M., Malkan, M., White, R., Telesco, C. M., Fisher, R. S., & Pina, R. K. 1998, ApJ, 505, 897
Kalas, P., & Jewitt, D. C. 1995, AJ, 110, 794
Kalas, P. 1998, Science, 281, 182
Kelsall, T., et al. 1998, ApJ, 508, 44
-- 50 --
Kessler, D. J. 1981, Icarus, 48, 39
Kimura, H., & Mann, I. 1998, ApJ, 499, 454
Koerner, D. W., Ressler, M. E., Werner, M. W., & Backman, D. E. 1998, ApJ, 503, L83
Kortenkamp, S. J., & Dermott, S. F. 1998, Icarus, 135, 469
Laor, A., & Draine, B. T. 1993, ApJ, 402, 441
Lecar, M., & Franklin, F. 1997, Icarus, 129, 134
Leinert, C., & Grun, E. 1990, in Space and Solar Physics, Vol. 20, Physics and Chemistry in Space:
Physics of the Inner Heliosphere I, eds. R. Schween, & E. Marsch (Berlin: Springer), 207
Liou, J. C. 1993, Ph.D. Thesis, Univ. of Florida
Lissauer, J. J. 1993, ARA&A, 31, 129
Love, S. G., & Brownlee, D. E. 1993, Science, 262, 550
Low, F. J., et al. 1984, ApJ, 278, L19
Malhotra, R. 1995, AJ, 110, 420
Malhotra, R. 1996, AJ, 111, 504
Marcy, G. W., & Butler, R. P. 1998, ARA&A, 36, 57
Murray, C. D., & Dermott, S. F. 1999, Solar System Dynamics, (Cambridge: Cambridge University
Press)
Reach, W. T., et al. 1995, Nature, 374, 521
Roques, F., Scholl, H., Sicardy, B., & Smith, B. A. 1994, Icarus, 108, 37
Schneider, G., et al. 1999, ApJ, 513, L127
Sitko, M. L., Grady, C. A., Lynch, D. K., Russell, R. W., & Hanner, M. S. 1999, ApJ, 510, 408
Stauffer, J. R., Hartmann, L. W., & Barrado Y Navascu´es, B. 1995, ApJ, 454, 910
Stern, S. A. 1995, AJ, 110, 856
Stern, S. A., & Colwell, J. E. 1997, ApJ, 490, 879
Sykes, M. V., Lebofsky, L. A., Hunten, D. M., & Low, F. J. 1986, Science, 232, 1115
Sykes, M. V. 1990, Icarus, 84, 267
-- 51 --
Telesco, C. M., Decher, R., Becklin, E. E., & Wolstencroft, R. D. 1988, Nature, 335, 51
Telesco, C. M., & Knacke, R. F. 1991, ApJ, 372, L29
Telesco, C. M., et al. 1999, ApJ, submitted
Trilling, D. E., Benz, W., Guillot, T., Lunin, J. I., Hubbard, W. B., & Burrows, A. 1998, ApJ,
500, 428
Vedder, J. D. 1998, Icarus, 131, 283
Ward, W. R., & Hahn, J. M. 1998, ApJ, 116, 489
Wisdom, J. 1980, AJ, 85, 1122
This preprint was prepared with the AAS LATEX macros v4.0.
-- 52 --
Fig. 1. -- Figure showing the new orbits of the fragments of a collision in which a large parent
particle "P", that was on a circular orbit around a star "S", was broken up. Fragments of different
sizes have different radiation pressure forces, characterized by a particle's β, acting on them, and
so have different orbits: those with β < 0.1, the "large" particles, have orbits that are close to
that of the parent; those with 0.1 < β < 0.5, the "β critical" particles, have orbits that have the
same pericenter distance as the parent, but larger apocenter distances; and those with β > 0.5,
the "β meteoroids", have hyperbolic orbits. Since when they are created, the velocity vector of all
fragments is perpendicular to the stellar direction, this is the point of their orbit's closest approach
to the star. The orbits of fragments with β =0, 0.1, 0.3, 0.4, 0.51, 1.0, and 2.2 are shown here. The
thick circular line denotes both the orbit of the parent particle and that of collisional fragments
with β = 0. All particles are orbiting the star counterclockwise.
-- 53 --
Fig. 2. -- (a) The osculating (instantaneous) complex eccentricity, z = e∗ exp iω = SD, of the orbit
of a particle in a system with one or more massive perturbers can be resolved vectorially into two
components: a forced eccentricity, zf = ef ∗ exp iωf = SC, that is imposed on the particle's orbit
by the perturbers; and a proper eccentricity, zp = ep ∗ exp iωp = CD, that is the particle's intrinsic
eccentricity. The secular evolution of its complex eccentricity is to precess counterclockwise around
the circle in (a), although the forced eccentricity may also vary with time. Initially, the orbital
elements of a family of collisional fragments created in the break-up of one large asteroid are the
same as those of the original asteroid. After a few precession timescales, their complex eccentricities
are evenly distributed around the circle in (a); this is because each fragment has a slightly different
precession timescale. Thus, these fragments have the same a, ef , ωf , ep, but random ωp, and so
their orbits have different eccentricities and orientations. The spatial distribution of these collisional
fragments is shown in (b). Their elliptical orbits are represented here by circles of radius a with
centers that are offset from the star, S, by ae in a direction opposite to the pericenter direction,
ω. A heavy line is used to highlight the orbit of a fragment with a pericenter P , and displaced
circle center D, where DP = a and SD = ae; the triangle SCD, where SC = aef and CD = aep,
corresponds to a similar one in (a). Since the distribution of ωp is random, it follows that the
points D for all the fragments are distributed on a circle of radius aep and center C. Thus, the
fragments form a uniform torus of inner radius a(1 − ep), and outer radius a(1 + ep), centered on a
point C displaced from the star by a distance aef in a direction away from the forced pericenter,
ωf (Dermott et al. 1985).
-- 54 --
(a)
(b)
Fig. 3. -- COBE observations in the 25 µm waveband of the variation of brightnesses at the North,
N , and South, S, ecliptic poles with ecliptic longitude of the Earth, λ⊕ (Dermott et al. 1999). (a)
N + S is at a minimum at λ⊕ = 224 ± 3◦. The displacement of this minimum from the Earth's
aphelion at λ⊕ = 282.9◦ implies that the center of symmetry of the zodiacal cloud is displaced from
the Sun. (b) N and S are equal when the Earth is at either the ascending or the descending node
of the plane of symmetry of the cloud at 1 AU, giving an ascending node of 70.7± 0.4◦. This plane
of symmetry is different from the plane of symmetry of the cloud at distances > 1 AU from the
Sun, implying that the cloud is warped.
-- 55 --
Fig. 4. -- The surface number density of the 2.5 µm dust grains in the HR 4796 disk derived from
the 18.2 µm brightness distribution (T99). The solid curve is the azimuthal average of the surface
number density, and the dotted and dashed lines indicate the density through the disk towards and
away from the forced pericenter direction of the model, respectively. The offset is a result of the
forced eccentricity imposed on the disk model; the inner edge of each side of the disk is offset by
∼ aminef ≈ 1 AU. The disk's surface density peaks at ∼ 1.02 × 109 m−2 at ∼ 70 AU. Interior to
this, the surface density falls to zero by 45 AU; the sloping cut-off is due to the eccentricities of the
disk model particles' orbits. Exterior to 70 AU, the surface density falls off ∝ r−3; this is due to
the distribution of the disk model particles' semimajor axes, n(a) ∝ a−2.
-- 56 --
Fig. 5. -- The thermal properties of astronomical silicate Mie spheres in the HR 4796 disk, plotted
for particles of different sizes at 40, 60, 80, and 100 AU from HR 4797A (a), (c), and (e), and
for 0.01, 2, 3, and 1000 µm diameter particles at different distances from HR 4796A (b), (d),
and (f ). The temperatures that these particles attain is plotted in (a) and (b). The contribution
of a particle's thermal emission to the flux density received at the Earth per solid angle that its
cross-sectional area subtends there, Qabs(D, λ)Bν [T (D, r), λ] (eq. [3]), is plotted for observations
in the IHW18, 18.2 µm, (c) and (d), and N, 10.8 µm, (e) and (f ) wavebands. The brightnesses of
disk models in these two wavebands were calculated by taking P (λ, r) from the lines on (d) and
(f ) corresponding to particles of diameter Dtyp.
-- 57 --
Fig. 6. -- (a) The ratio of the thermal emission in the N, 10.8 µm, and IHW18, 18.2 µm, wavebands,
of astronomical silicate Mie spheres of different sizes at 40, 60, 80, and 100 AU from HR 4796A
(i.e., Fig. 5e divided by Fig. 5c). Assuming the disk's flux densities in the two wavebands to be
dominated by the emission of particles at 60-80 AU, the observed ratio of flux densities, O(0.05)
(T99; §5.2.2), can be used to estimate that the disk's emitting particles have Dtyp = 2 − 3 µm.
(b) The ratio, β, of the radiation pressure force to the gravitational force acting on astronomical
silicate Mie spheres of different sizes in the HR 4796 disk.
-- 58 --
Fig. 7. -- (a) Contour plot of an unsmoothed face-on view of the HR 4796 disk model (shown also
in Fig. 4) seen in the IHW18 (18.2 µm) waveband. The contours are spaced linearly at 0.17, 0.34,
0.51, 0.68, 0.85, and 1.02 mJy/pixel. The disk's offset causes particles in the forced pericenter
direction, located at a position angle of 90◦ − ωf , where ωf = 26◦, as measured from North in a
counterclockwise direction, to be hotter, and hence brighter, than those in the forced apocenter
direction; this is the "pericenter glow" phenomenon, evident in this figure by the shape of the
brightest (filled-in) contour. The geometry of the observation is defined such that the disk as it is
shown here is rotated by a further 90◦ − Iobs about the dotted line, where Iobs = 13◦. (b) Relation
of forced eccentricity, ef , to the orientation of the longitude of forced pericenter, ωf , in the model
to achieve the observed lobe brightness asymmetry of 5.1± 3.2% (a similar relationship is necessary
to achieve the observed radial offset). A forced eccentricity as small as 0.02 would suffice to achieve
the observed asymmetry, but a higher forced eccentricity could be necessary if the forced pericenter
is aligned in an unfavorable direction. Our final model, shown in Figs. 4, 7-9, has ef = 0.02 and
ωf = 26◦, and this point is shown with an asterisk on this plot.
-- 59 --
s
d
n
o
c
e
s
c
r
A
1.0
0.5
0.0
-0.5
-1.0
-2
(a)
-1
0
Arcseconds
1
2
-2
-1
0
Arcseconds
1
2
(b)
Fig. 8. -- Top -- False color images of HR 4796 in the IHW18 (18.2 µm) waveband. Both the
observation (a), on the left, and the model (b), on the right, have been rotated to horizontal with
the NE lobe on the left. The contours are spaced linearly at 0.22, 0.35, 0.49, 0.62, 0.75, 0.89, 1.02,
1.15, 1.29, and 1.42 mJy/pixel. The observation has had the photospheric emission of HR 4796A
subtracted, and a 3 pixel FWHM gaussian smoothing applied. The image of the model mimics the
observation both in pixel size (1 pixel = 0.′′0616 = 4.133 AU) and smoothing (using an observed
PSF and including the 3 pixel post-observational smoothing). Bottom -- Line-cuts in the vertical
direction through the NE (c) and SW (e) lobes, and in the horizontal direction through the center
of both lobes (d). The observations are shown with a solid line and the model with a dotted line.
-- 60 --
Fig. 9. -- Horizontal line-cuts along the plane of the lobes in the N (10.8 µm) band. The observation
is shown with a solid line and models with particle diameters of Dtyp = 2, 2.5, and 3 µm are shown
with dotted, dashed, and dash-dot lines. The total amount of cross-sectional area in the models,
σtot, has been scaled to fit the observed mean brightness of the lobes in the IHW18 (18.2 µm)
waveband; the model with Dtyp = 2.5 µm gives the best fit to the observed lobe brightnesses in the
N band. The observed N band flux density is not well constrained within 0.′′8 of HR 4796A due to
imperfect subtraction of the stellar photosphere from the image (T99), and so it is not shown here.
-- 61 --
Fig. 10. -- Plots of the forced eccentricities imposed on the orbits of particles in the HR 4796 system
as a function of their semimajor axes, assuming different combinations of perturbers in the system:
(a) HR 4796B only; (b) HR 4796B and a planet at the inner edge of the disk; (c) a planet only.
The forced inclinations imposed by the two perturber system are shown in (d). The inner edge of
the disk is at a = 62 AU. HR 4796B and the planet are assumed to have: MB = 0.38M⊙, aB = 517
AU, eB = 0.13, and IB = 0; Mpl = 0.1 and 10 MJ , where MJ is the mass of Jupiter, apl = 47
AU, epl = 0.023, and Ipl = 5◦. The orbital elements of the planet are marked by an asterisk on the
forced elements plots. In a one perturber system the forced eccentricity is independent of the mass
of the perturber, and the forced inclination is the plane of the perturber's orbit. In a two perturber
system, the shapes of the forced element plots depend on the mass of the planet, and the forced
eccentricity also depends on the orientations of the perturbers' orbits. The forced eccentricity is
plotted in (b) assuming that ωpl = ωB + 180◦. This means that ωf = ωpl for a < acrit, and ωf = ωB
for a > acrit, where acrit is the semimajor axis of a particle's orbit for which ef = 0; a similar
alignment with the perturbers' orbital planes seen in the plot of the particles' forced inclinations.
Thus, the lobes have both their asymmetries and their plane of symmetry aligned with the orbit of
the planet if Mpl > 0.1MJ , and with the orbit of HR 4796B if Mpl < 0.1MJ .
|
astro-ph/0309021 | 1 | 0309 | 2003-09-01T08:46:31 | Proton capture elements in the globular cluster NGC 2808 I. First detection of large variations in Sodium abundances along the Red Giant Branch | [
"astro-ph"
] | We have used spectra obtained as part of the Science Verification program of the FLAMES multi-object spectrograph mounted on Kueyen (VLT-UT2) to perform an abundance analysis of stars along the giant branch (RGB) in the globular cluster NGC 2808. Sodium abundances are derived from Na D lines for a sample of 81 cluster stars spanning a range of about 2 magnitudes from the tip of the RGB. Our results show that a large star-to-star scatter does exist at all positions along the RGB, suggesting large variations in the abundance of proton capture elements down to luminosities comparable to the red horizontal branch (HB). The distribution of Na abundances along the RGB seems to point out that in this cluster most of the observed spread has a primordial origin. Overimposed evolutionary effects, if any, must be only at ``noise'' level, at odds with results from a similar analysis in M 13 (Pilachowski et al. 1996). This study is a first step towards ascertaining if a link exists between the distribution of chemical anomalies in light elements along the RGB and the global properties of globular clusters, in particular the HB morphology. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. carretta
(DOI: will be inserted by hand later)
November 3, 2018
3
0
0
2
p
e
S
1
1
v
1
2
0
9
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Proton capture elements in the globular cluster NGC 2808 ⋆
I. First detection of large variations in Sodium abundances along the Red
Giant Branch
E. Carretta1, A. Bragaglia2, C. Cacciari2 , E. Rossetti2
(1) INAF - Osservatorio Astronomico di Padova, vicolo dell'Osservatorio 5, 35122 Padova, Italy
(2) INAF - Osservatorio Astronomico di Bologna, via Ranzani 1, 40127 Bologna, Italy
Received ; accepted
Abstract. We have used spectra obtained as part of the Science Verification program of the FLAMES multi-object
spectrograph mounted on Kueyen (VLT-UT2) to perform an abundance analysis of stars along the giant branch
(RGB) in the globular cluster NGC 2808. Sodium abundances are derived from Na D lines for a sample of 81
cluster stars spanning a range of about 2 magnitudes from the tip of the RGB. Our results show that a large
star-to-star scatter does exist at all positions along the RGB, suggesting large variations in the abundance of
proton capture elements down to luminosities comparable to the red horizontal branch (HB). The distribution
of Na abundances along the RGB seems to point out that in this cluster most of the observed spread has a
primordial origin. Overimposed evolutionary effects, if any, must be only at "noise" level, at odds with results
from a similar analysis in M 13 (Pilachowski et al. 1996). This study is a first step towards ascertaining if a link
exists between the distribution of chemical anomalies in light elements along the RGB and the global properties
of globular clusters, in particular the HB morphology.
Key words. Stars: abundances -- Stars: evolution -- Stars: Population II -- Galaxy: globular clusters: general
Galaxy: globular clusters: individual: NGC 2808
1. INTRODUCTION
For a long time the globular clusters (GC) have been con-
sidered the best approximation in the Universe of Simple
Stellar Populations (SSP), i.e. associations of single, co-
eval stars sharing the same initial chemical composition.
These assumptions have been questioned in the last few
years. Apart from the relevant presence of binary stars
in GC's (showing up as Blue Stragglers or eclipsing bina-
ries or peculiar objects, and strongly needed by dynami-
cal considerations as a crucial ingredient in the dynamical
evolution of clusters), also the other two properties that
define a SSP have been seriously challenged on the basis
of modern, accurate studies of the chemical composition
of cluster stars.
Leaving aside the very peculiar case of ω Centauri, it
has been evident for more than 30 years (Osborn 1971)
that the hypothesis of monometallicity holds for GC only
as far as "heavy" elements (those belonging to the Fe
group) are concerned. Early studies (see the comprehen-
Send offprint requests to: E. Carretta, [email protected]
⋆ Based on data collected at
the European Southern
Observatory, Chile, during the FLAMES Science Verification
sive reviews by Smith 1987 and Kraft 1994) based on
indexes from photometry or low dispersion spectroscopy
clearly demonstrated that the lighter elements (in primis
Carbon and Nitrogen) showed marked differences along
the RGB in several GC's. Moreover, striking variations in
the CH band strenghts, anticorrelated with CN (and NH,
when accessible) strenghts were observed in several nearby
clusters as far down as to main sequence and turn-off stars
(see e.g. Cannon et al. 1998 and references therein)
Three fundamental steps forward have been made,
thanks to the continuous progress in the building of ef-
ficient spectrographs: (i) since the early '90s, the high res-
olution studies by the Lick/Texas group (see Ivans et al.
2001 for updated references) well assessed that Na and
O abundances are anti-correlated among the first ascent
red giant branch (RGB) stars in almost all the clusters
surveyed, for at least 1 magnitude below the RGB tip;
(ii) the anomalies seem to be restricted to cluster stars,
while the abundance pattern of field stars is perfectly ex-
plained by the classical first dredge-up and a second mix-
ing episode after the level of the RGB-bump (Gratton et
al. 2000); (iii) the Na-O anticorrelation was recently found
for the first time among turn-off stars in NGC 6752 by
2
E. Carretta et al.: Proton capture elements in NGC 2808
Gratton et al. (2001). These seminal works point out that
not only the CN-cycle (easily detectable even with low
resolution analysis), but also the ON-cycle of the com-
plete CNO H-burning cycle is at work, producing the ob-
served pattern. In particular, Na can be produced from
22Ne by the proton-capture fusion mechanism that is at
work in the high temperature (inner) regions where O is
transformed into N (see Denisenkov& Denisenkova 1990;
Langer, Hoffman & Sneden 1993).
The Na-O anticorrelation is the most convincing evi-
dence of this process that, however, is restricted only to
high density environments (such as those in the globular
clusters), since Gratton et al. (2000) showed that Na and
O abundances are hardly involved in the normal evolution
along the RGB of isolated field stars. In a still unknown
way, the large number of interactions between stars in GCs
has to be involved in producing the observed pattern of
chemical abundances.
Finally, the recent analysis of high resolution high
quality UVES spectra of turn-off and early subgiant stars
in NGC 6397 and NGC 6752 (Gratton et al. 2001) and 47
Tuc (Carretta et al. 2003a, in preparation) indicated that
groups of both Na-rich/O-poor and Na-poor/O-rich un-
evolved stars are present in all these clusters. In turn, this
must be due to pre-existing abundance variations, since
these stars do not have the requirements (central high
temperatures and an extended envelope able to bring the
products of proton fusion to the photosphere) to produce
the observed chemical pattern as a consequence of internal
mixing during their evolution.
Primordial variations, viable to explain the observed
anomalies, have been suggested both in the primordial
gas of the proto-cluster (see Cottrell & Da Costa 1981)
in the very early phases of star/cluster formation, and as
Na-rich, O-poor material ejected from a first generation of
(now extinct) intermediate mass AGB stars that polluted
the gas from which a second generation of stars (that are
presently observed) formed (D'antona, Gratton & Chieffi
1983; Ventura et al. 2001). Notice that in this latter case
the assumption of strictly coeval stars in a GC must be
dropped.
On the other hand, the increasingly high N abun-
dances, accompanied by a progressive decline in C abun-
dance and 12C/13C ratio, observed at increasing luminos-
ity along the RGB in many clusters, testifies that changes
are going on during the evolution of stars, such as e.g.
very deep mixing in RGB stars, that may be triggered by
enhanced core rotation possibly due to the dense cluster
environment.
Nowadays, the most likely explanation of the overall
chemical pattern observed in GC stars is that a contribu-
tion of both aspects (primordial and evolutionary mixing)
is required. The debate seems presently shifted rather on
how to disentangle these two contributions and to prop-
erly ascertain their relative weights within a cluster and in
clusters of different physical properties (Horizontal Branch
(HB) morphology, density, age, metallicity, etc.).
To this purpose large and homogeneous sets of cluster
stars must be observed, possibly in different evolutionary
stages. This has become possible only recently, with the
advent of efficient multiobject spectrographs on large tele-
scopes, such as Hydra at the KPNO 4-m and WIYN 3.5 m
telescopes used by Pilachowski et al. (1996b) and Sneden,
Pilachowski & Kraft (2000) to survey the RGB in M13,
or in M15 and M92 respectively.
FLAMES mounted at the 8m Kueyen VLT-UT2 tele-
scope is the ideal instrument for this type of studies (see
Pasquini et al. 2002 for a technical description).
NGC 2808 is by itself a very interesting object, since
its HB morphology (a red stubby HB and a population
of blue HB stars, about 1/4 of the total HB stars) has
often been labelled as the second parameter effect working
within an individual cluster. However, the overwhelming
majority of previous works was photometric, to investigate
the morphology of the color-magnitude diagram (CMD).
The only spectroscopic study is that of Gratton (1982),
who analyzed a single giant star obtaining a metallicity of
[Fe/H]= −1.1 dex1.
This situation dramatically changed at the end of
January 2003, when ∼ 130 RGB stars in NGC 2808
were observed during the program "Mass loss in Globular
Cluster Red Giant Stars" (proposed by C. Cacciari and
A. Bragaglia), that was part of the FLAMES Science
Verification program. The results of this study are pre-
sented in a separate paper (Cacciari et al. 2003, in prepa-
ration). In the present paper we exploit part of this mate-
rial for the analysis of the Sodium D lines in order to get
insights on the distribution of this proton capture element
along the RGB.
In the next section we present the star selection (that
was performed for studying the mass loss process and was
not optimized for the present purpose) and the observa-
tions. In Sect. 3 and 4, respectively, we explain the atmo-
spheric parameters adopted and the abundance analysis
for Na, whose results are discussed in Sect. 5. Comparison
with results for other clusters are made in Sect. 6, and a
short summary is given in Sect. 7.
2. SAMPLE SELECTION AND OBSERVATIONS
Our program stars were observed in 2003, January 24-25
during the FLAMES Science Verification program at the
ESO Paranal Observatory. The data were released to the
public on 2003, March 3. The night was not photomet-
ric, the sky conditions were thin cirrus and the seeing was
about 0.9 arcsec. 86 stars were observed with the grat-
ing HR 12 of GIRAFFE, giving an useful spectral range
from 5820 to 6134 A, with a resolution of about 15,000 at
the centre of spectra. During all GIRAFFE exposures, the
1 We use the usual spectroscopic notation: log n(A) is the
abundance (by number) of the element A in the usual scale
where log n(H)=12; notation [A/H] is the logarithmic ratio of
the abundances of elements A and H in the star, minus the
same quantity in the Sun.
E. Carretta et al.: Proton capture elements in NGC 2808
3
Fig. 1. Left panel: V , B − V colour-magnitude diagram
of NGC 2808 (Piotto et al. 2003, in preparation). Stars
observed with the GIRAFFE HR 12 grating in the Medusa
configuration are indicated by open circles, while filled
circles are stars observed with fibers feeding the UVES
Red Arm. Right panel: average S/N in 0.5 mag bins along
the RGB.
available fibers (8 for each setup, including one fiber on the
sky) feeding the Red Arm of the high resolution spectro-
graph UVES were centered on other RGB stars, for a to-
tal of 20 objects observed at high resolution (R∼ 45, 000).
Only 4 stars were observed by both GIRAFFE and UVES.
Details of these UVES observations and their analysis are
deferred to a forthcoming paper (Carretta et al. 2003b, in
preparation). Here, we concentrate only on the GIRAFFE
sample in the Na D region.
Fig. 1 indicates the position of our program stars in
the V , B − V color-magnitude diagram of NGC 2808. We
remind here that the main purpose of this program was
to study the mass loss along the RGB, so the observations
are concentrated in the region close to the RGB tip. Cool
and (moderately) metal-rich stars are not the best targets
for abundance analysis due to the line crowding, affecting
the continuum positioning, and to concerns on existing
model atmospheres that are unable to well reproduce the
cooler red giants.
In Fig. 2 typical spectra are displayed for stars span-
ning all the range of parameters sampled along the RGB.
The sample goes from the tip of the RGB (about
MV = −3, adopting a distance modulus (m-M)V =15.59
from Harris 1996) to just below the level of the red HB
(about MV = 1). Only two stars have been observed be-
low V = 16, the bulk of program stars lie within 2 mag
from the RGB tip. The right panel in Fig. 1 shows how
the average S/N ratio varies as a function of the V mag-
Fig. 2. Typical spectra for bright (upper panels), inter-
mediate (middle panels) and faint (lower panels) objects
in our sample with different Na abundances. The small
arrows indicate the position of the stellar Na I D lines;
notice how well they can be separated from the interstel-
lar lines, here shifted to the blue. We also indicate for each
star the values of Teff, log g, and [Na/Fe]. S/N (per pixel)
varies from about 100, to 85, to 45 for the three magnitude
levels.
nitude along the RGB. The scatter around a mean value
at fixed magnitude is likely due to a combination of small
mis-positioning errors of the star in the individual fibers
and different throughtputs in transmission from different
fibers.
Cluster membership was mostly secured by choosing
stars near the mean ridge line of the giant branch and
checked a posteriori by measuring radial velocities from
spectral lines. The radial velocity (RV) for each star was
derived by fitting Gaussian profiles to about 30 lines (with
the IRAF2 routine rvidlines).
Fig. 3 shows the distribution of radial velocities. A
small zero point offset between fibers is possible, due to
the reduction procedure not optimized for precise radial
velocities measurements, that could reach ± 10 kms−1,
according to the ESO relevant web pages. We have not
done any detailed check on radial velocities, since it was
outside our main purpose; we only note that the 4 stars
that were observed also with UVES at high resolution -
and more reliable wavelength calibration - show a differ-
ence between derived velocities from 1 to -- 8 kms−1, with
an average of -- 3.5 kms−1, and measurement of the telluric
Na I line at 5889.884 A in star #56136 indicates a velocity
2 IRAF is distributed by the NOAO, which are operated by
AURA, under contract with NSF
4
E. Carretta et al.: Proton capture elements in NGC 2808
preparation) have done a similar survey of the same stars
we are analysing here, and detected emissions only for one
such star (#10341): we will still retain it in our sample,
cautioning that it could well not be a true RGB star.
3. ATMOSPHERIC PARAMETERS
(2003),
As described in Cacciari et al.
the effec-
tive temperatures Teff and bolometric corrections B.C.
for our stars were derived from the V − K colors
in the 2MASS All-Sky Data Release (accessible at
www.ipac.caltech.edu/2mass/releases/allsky/ and released
on March 25, 2003), and using the calibrating equations
by Alonso et al. (1999, with the erratum of Alonso et
al. 2001). We have adopted AK = 0.353E(B − V ) and
AV = 3.1E(B − V ) from Cardelli et al. (1989), and
E(B − V )=0.22 from Harris (1996).
The 2MASS photometry was transformed to the TCS
system and an input metallicity of [Fe/H]= −1.25 was
adopted3.
Surface gravities log g were obtained from effective
temperatures and bolometric corrections, using the dis-
tance modulus (m − M )V = 15.59, and assuming that the
stars have masses of 0.85 M⊙. The adopted bolometric
magnitude of the Sun is M (bol)⊙ = 4.75.
For the overall metallicity, we adopted the average
metal abundance as derived by the analysis of the high
resolution UVES red spectra (Carretta et al. 2003b, in
preparation): [A/H]= −1.14 ± 0.01 dex, σ = 0.06 dex
(19 stars)4. This choice stems from the small number of
Fe I lines available in our spectral range (about 17 lines,
on average, are measurable per star) and from the rather
low resolution. As discussed, e.g., in Carretta & Gratton
(1997), at lower resolution line blends are more frequent,
leading to overestimate the equivalent widths (EW) in a
complex way that is a function of metallicity, S/N and
temperature (all affecting the line crowding), as well as
resolution. Since the RGB magnitude range monitored by
GIRAFFE observations was covered also by UVES obser-
vations, we believe that the use of UVES-based metallic-
ities is justified and does not bias the results we obtain
from GIRAFFE data.
Finally, based on the same considerations, we adopted
the microturbulent velocity vt derived from the high res-
olution UVES spectra, using a large number of Fe I lines
of different strenghts. The resulting relation is well con-
strained, so we adopted vt=1.73 kms−1 for Teff < 4100 K
and vt = −1.3 · 10−4 × Teff + 7.10 for Teff ≥ 4100.
Star designations and derived atmospheric parameters
are listed in the first four column of Table 1.
3 This value is intermediate among several previous deter-
minations (cf. Walker 1999), and is slightly different from the
metallicity we adopt in the present analysis; however, the de-
pendence of (V − K) on [Fe/H] is so weak that temperatures
are almost unaffected by this difference
4 This in turn is based on a solar Iron abundance of logn(Fe
I)⊙ = 7.54, see Gratton et al. (2001).
Fig. 3. Upper panel: measured radial velocities for pro-
gram stars plotted as a function of the fiber numbers.
Filled circles are stars rejected from our analysis as non-
members. Lower panel: histogram of the radial velocities,
where the rejected stars are indicated by the filled his-
togram.
zero point of 3.5 kms−1. So the derived RV's are in general
precise enough to discriminate between cluster and field
stars. Using the RV distribution, we disregarded 4 stars
as sure field interlopers. Another star was rejected from
the sample after the abundance analysis (see below, Sect.
4). The average heliocentric velocity is RV = 102.1 kms−1
(σ = 10.9), after eliminating these 5 non-members.
The final sample consists therefore in 81 red giant
stars. What is the physical evolutionary status of these
giants? Are they truly first ascent giant stars or is the
sample contaminated by stars in a more evolved stage, i.e.
Asymptotic Giant Branch (AGB) stars? For the bright-
est objects there is no way to discriminate between RGB
and AGB stars either photometrically or spectroscopi-
cally: their position in the CMD is almost identical, and
small differences in colours may well be due to photomet-
ric errors and/or variability, and in both cases emissions
on the Hα wings can be present. Anyway, from robust
evolutionary theory expectations, AGB stars should rep-
resent only a small fraction (∼ 10%) of the brightest ob-
jects. When we consider lower luminosity stars, photomet-
ric distinction is easier, and we may also hope to detect
AGB stars via Hα emission wings (as done by Sneden et
al. 2000): this emission, which is rather common also in
RGB stars near the tip (see e.g. Lyons et al. 1996, and
the vast literature cited there), is not present in fainter
< -- 1 or fainter present
objects. Hence, if stars with MV ∼
Hα emission wings, they may be taken as -- or at least
suspected to be -- AGB stars. Cacciari et al. (2003, in
E. Carretta et al.: Proton capture elements in NGC 2808
5
Table 1. Atmospheric parameters and derived Na abun-
dances, both in LTE and in NLTE. Star designations are
taken from Piotto et al. (2003)
Star
Teff
log g
vt
[Na/Fe]
(K)
4245
4691
4459
4290
4211
4564
4132
4243
4455
4717
4283
4399
4252
4569
4826
4733
4666
4731
4379
4788
4525
4572
4538
4312
4547
4547
4317
4588
4659
4717
4357
4364
4684
4409
4377
4268
4906
4813
4791
4769
4035
4463
4431
4411
4310
4405
4425
4220
4568
4383
4474
4463
4051
4288
4375
3969
4368
7536
7558
7788
8603
8739
8826
9230
10012
10105
10201
10265
10341
10571
13575
13983
30523
30900
31851
32469
32685
32862
32909
32924
33918
35265
37496
37998
38228
38244
38967
39060
40983
41008
41828
41969
42073
42165
42789
42886
42996
43041
43247
43794
43822
44665
44984
45443
45840
46041
46367
46663
46868
47145
47421
48011
48060
48128
dex
1.18
1.87
1.53
1.24
1.12
1.66
0.96
1.12
1.44
2.02
1.18
1.40
1.19
1.80
2.17
1.94
1.74
1.83
1.14
2.03
1.69
1.66
1.57
1.21
1.76
1.58
1.18
1.65
1.78
1.83
1.28
1.26
1.79
1.36
1.30
1.17
2.39
1.93
2.14
1.93
0.70
1.50
1.38
1.39
1.18
1.36
1.39
1.06
1.63
1.46
1.41
1.50
0.74
1.14
1.28
0.63
1.31
kms−1
1.58
1.00
1.30
1.52
1.63
1.17
1.73
1.58
1.31
0.97
1.53
1.38
1.57
1.16
0.83
0.95
1.03
0.95
1.41
0.88
1.22
1.16
1.20
1.49
1.19
1.19
1.49
1.14
1.04
0.97
1.44
1.43
1.01
1.37
1.41
1.55
0.72
0.84
0.87
0.90
1.73
1.30
1.34
1.37
1.50
1.37
1.35
1.61
1.16
1.40
1.28
1.30
1.73
1.53
1.41
1.73
1.42
LTE
dex
−0.16
+0.17
+0.12
−0.24
−0.08
+0.42
+0.04
+0.23
−0.05
+0.42
+0.24
+0.44
−0.01
−0.23
+0.35
+0.26
+0.21
+0.77
+0.44
+0.37
+0.15
+0.29
+0.54
+0.06
+0.07
+0.26
+0.12
+0.38
+0.23
+0.41
+0.21
+0.29
+0.67
+0.11
+0.07
−0.02
−0.09
+0.78
+0.11
+0.29
−0.23
+0.29
+0.05
+0.16
+0.17
−0.00
+0.05
−0.14
+0.26
+0.21
+0.47
+0.05
+0.02
+0.21
+0.63
−0.35
+0.48
[Na/Fe]
NLTE
dex
+0.06
+0.28
+0.29
−0.02
+0.14
+0.57
+0.29
+0.45
+0.13
+0.52
+0.45
+0.63
+0.21
−0.10
+0.42
+0.37
+0.34
+0.89
+0.65
+0.47
+0.30
+0.44
+0.70
+0.27
+0.21
+0.42
+0.34
+0.52
+0.36
+0.53
+0.41
+0.49
+0.80
+0.30
+0.26
+0.20
−0.06
+0.89
+0.19
+0.39
+0.05
+0.46
+0.24
+0.34
+0.39
+0.19
+0.23
+0.10
+0.41
+0.38
+0.65
+0.22
+0.29
+0.43
+0.83
−0.05
+0.68
σ
dex
0.04
0.09
0.15
0.13
0.15
0.16
0.02
0.04
0.01
0.08
0.04
0.04
0.08
0.01
0.16
0.05
0.11
0.03
0.11
0.01
0.10
0.09
0.15
0.09
0.10
0.03
0.08
0.01
0.04
0.08
0.08
0.01
0.03
0.22
0.01
0.13
0.04
0.09
0.08
0.09
0.04
0.09
0.13
0.03
0.16
0.13
0.13
0.03
0.04
0.03
0.04
0.02
0.05
0.11
0.01
0.04
0.08
Table 1. (continue)
Star
Teff
log g
vt
[Na/Fe]
(K)
4164
3952
4665
4023
4054
4039
4432
4577
4546
4154
4886
3970
4242
4397
4668
4327
4236
4355
3995
4440
4430
4459
4905
4643
dex
0.94
0.62
1.77
0.75
0.82
0.69
1.42
1.64
1.61
1.05
2.01
0.62
1.11
1.27
1.84
1.26
1.08
1.21
0.67
1.44
1.46
1.41
2.38
1.86
49509
49680
49743
49942
50371
50861
50866
50910
51416
51515
51646
51871
51963
52006
54264
54284
54733
54789
55031
55437
55609
55627
56136
56710
kms−1
1.69
1.73
1.04
1.73
1.73
1.73
1.34
1.15
1.19
1.70
0.75
1.73
1.59
1.38
1.03
1.47
1.59
1.44
1.73
1.33
1.34
1.30
0.72
1.06
LTE
dex
+0.12
−0.37
+0.46
−0.15
−0.23
+0.13
+0.40
+0.25
+0.39
+0.04
+0.88
−0.02
+0.06
+0.62
+0.15
+0.14
+0.25
+0.66
−0.35
+0.27
+0.23
+0.52
+0.04
+0.04
[Na/Fe]
NLTE
dex
+0.37
−0.07
+0.60
+0.13
+0.04
+0.41
+0.58
+0.40
+0.54
+0.28
+0.98
+0.27
+0.28
+0.81
+0.27
+0.34
+0.48
+0.87
−0.05
+0.45
+0.41
+0.71
+0.09
+0.16
σ
dex
0.08
0.05
0.06
0.02
0.07
0.06
0.09
0.02
0.09
0.06
0.08
0.10
0.04
0.09
0.09
0.11
0.05
0.01
0.05
0.05
0.04
0.13
0.09
0.09
4. ANALYSIS: SODIUM ABUNDANCES
Equivalent width (EW) of Na D lines were measured fol-
lowing the method outlined in Bragaglia et al. (2001): we
refer to that paper for further details. The fraction of the
(highest) spectrum points to be used by the automatic
program in order to set a local continuum centered on the
feature of interest is the crucial point of the procedure. We
checked this parameter by inspecting by eye the contin-
uum location as defined by the program for some lines of
typical spectra distributed all over the temperature range.
After this test, we set this fraction to 1/3 for stars with
Teff ≤ 4200 K and to 1/2 for warmer stars.
Using this value and the atmospheric parameters de-
rived in the previous section, we measured the EWs of
the Na D lines at 5889.97 and 5894.94 A for all 81 pro-
gram stars, and derived the Na abundances interpolating
in the Kurucz (1995) grid of model atmospheres with the
overshooting option set on.
However, a Gaussian function is a poor approximation
for fitting these very strong lines and may give misleading
results, missing the relevant contribution of the damping
wings of the lines. For these reasons we decided to adopt
the Na abundances obtained from comparison with syn-
thetic D1 and D2 lines: the procedure is much more time
consuming but gives more reliable results than a Gaussian
fit or even a direct integration. Moreover, concerns related
to the positioning of the continuum are much reduced.
6
E. Carretta et al.: Proton capture elements in NGC 2808
Hence, we computed a grid of synthetic spectra for
each of the Na D lines. Each grid includes 21 spectra with
Teff ranging from 3950 to 4950 K (this range brackets
all our program stars) in steps of 50 K. Using the single
value [A/H]= −1.14 dex, vt for each synthetic spectrum
was computed using the relation obtained from the UVES
spectra, as described in the previous section; the gravity
was derived from a Teff-log g relation obtained with a lin-
ear fit from our program stars.
For each of the 21 spectra (i.e. values of Teff ) in our
grid we computed 10 synthetic spectra of the Na line by
varying the [Na/Fe] ratio from -- 0.4 to +1.4 dex, in steps
of 0.2 dex. Thus we could associate to each Na line a
set of 210 synthetic spectra spanning the relevant range
both in temperature and Na abundances. These synthetic
spectra were convolved with a Gaussian function to match
the resolution of the GIRAFFE spectra. We found that a
smoothing FWHM of 0.3 A is appropriate for our program
stars.
Starting with input Na abundances determined by line
analysis, we chose the nearest value of temperature for
each star, and compared the observed spectrum to the 3
synthetic spectra bracketing the Na abundance from EW.
Visual inspection was used to obtain the Na abundances
from each line. For the star Na abundance we then took
simply the average from the two Na D lines, adopting log
n(Na)=6.23 as the solar Na abundance.
In Fig. 4 we show two examples of the fitting of syn-
thetic spectra to the observed Na D lines for two stars
in different position along the RGB. In the upper pan-
els, the star 41008 has Teff = 4684 K and was compared
to spectra computed at 4700 K, while in the lower pan-
els the program star has Teff = 3969 and Na abundances
are obtained through comparison with synthetic spectra
computed at 3950 K.
This Figure shows that despite the difference of ∼ 1
dex in Na abundance and ∼ 800 K in temperature, the
Na features are well reproduced.
A comparison of Na abundance derived by line anal-
ysis (EWs) and by spectrum synthesis is highly instruc-
tive. The average abundance difference (in the sense syn-
thesis minus line) is ∆[Na/Fe]=+0.30±0.01 dex, with
an r.m.s.=0.09 dex (81 stars5). This test illustrates how
severely the Gaussian fitting approximation may underes-
timate the measure of EWs for these strong lines, resulting
in spurious Na abundances.
Since the observations were made for other purposes
and did not aim at deriving chemical abundances, rapidly
rotating early type stars were not observed. Lacking a suit-
able template to subtract telluric lines, we evaluated their
5 We exclude from the sample and from the following dis-
cussion star 37781, that is probably not-member of the clus-
ter even if its radial velocity is compatible with member-
ship. The unusually high Na abundance obtained for this stars
([Na/Fe]≃ +1.5) is likely due to an erroneous estimate of at-
mospheric parameters due to the initial assumption that this
star is member of NGC 2808, or because it might be an AGB
star.
Fig. 4. Upper left panel: spectrum synthesis of the Na D2
line at 5889.97 A in star 41008. Open squares are the ob-
served spectrum, while lines represent the synthetic spec-
tra computed at 4700 K for 3 different Na abundances:
[Na/Fe]=0.4, 0.6 and 0.8 dex from top to bottom, respec-
tively. Upper right panel: the same, but for the Na D1
line at 5895.94 A. Lower left panel: spectrum synthesis
of the Na D2 line in star 48060, near the RGB tip. The
plotted synthetic spectra are computed at 3950 K and for
abundance ratios of [Na/Fe]= -0.4, -0.2 and 0.0 from top
to bottom. Lower right panel: the same, but for the Na
D1 line. All synthetic spectra have been convolved with
a Gaussian having FWHM=0.3 A to take into account
the instrumental profile of the observed spectra. All abun-
dances are given in the LTE assumption.
impact on derived abundances by estimating their contri-
bution to the EW of the Na lines within the region where
the fitting of synthetic spectra was performed. We veri-
fied that the possible effect of telluric lines on the derived
abundances amounts to no more than a few hundredths
of a dex and was therefore considered negligible.
Finally, we neglected the hyperfine splitting for Na D
lines. Explorative computations with atomic parameters
for the components of each line taken from McWilliam et
al. (1995) allowed us to conclude that the effect on our Na
abundances is negligible, likely because in very strong lines
the effect itself is reduced since all hyperfine components
in the line core are saturated (see McWilliam et al. 1995).
4.1. NLTE corrections
Statistical equilibrium computations (e.g. Bruls et al.
1992, Gratton et al. 1999) show that the NLTE abundance
corrections for Na D lines follow a rather complicated pat-
tern, due to the complex relative interplay between pho-
E. Carretta et al.: Proton capture elements in NGC 2808
7
Fig. 5. Corrections to the Na abundances for departure
from LTE as a function of effective temperature, derived
for our program stars.
toionization and the so-called photo suction effect (Bruls
et al. 1992) that consists in a recombination cascade to
the ground level, where the D lines originate. Moreover,
the strong Na D lines form high in the atmosphere, where
departures from LTE are larger. In order to assess the
true Na distribution, a careful evaluation of these effects
is then required.
We derived appropriate corrections for departures from
the LTE following the prescriptions of Gratton et al.
(1999). Since these corrections are a function of both stel-
lar physical status (temperature and gravity) and line
strength, we measured the equivalent widths of synthetic
Na D2 and D1 lines for all (Teff , [Na/Fe]) pairs. These
EWs were used to construct interpolating relations (one
for each line) as a function of temperature and Na abun-
dance.
Next, we entered in these relations with stellar temper-
atures and previously derived LTE Na abundances from
spectrum synthesis for all our program stars, deriving
[Na/Fe] ratios properly corrected for departures from LTE
following Gratton et al. (1999). These corrected abun-
dances are listed in Table 1 (labeled as [Na/Fe]N LT E) and
represent our final set of abundances on which the follow-
ing considerations and discussions are based. As reference
for the reader, in this Table are also listed abundances
derived under the LTE assumption. The errors listed in
Table 1 are the r.m.s. of the straight average between the
abundances as given by Na D2 and D1 lines.
In Fig. 5 we show the NLTE corrections to [Na/Fe]
LTE abundances as a function of effective temperature.
As one can see, the corrections are progressively larger
going from lower luminosity giants toward giants near the
Fig. 6. Upper panel: abundance ratios [Na/Fe] for stars
along the RGB of NGC 2808, corrected for the effects of
departures from LTE, as a function of the effective tem-
peratures. At the top of the figure the approximate lu-
minosity levels for the RGB-tip, the RGB-bump and the
average luminosity of the red extreme of the HB are in-
dicated. At the bottom of the figure we show the typical
error bars over the whole range of temperatures spanned
by our sample, due to random uncertainties in the adopted
atmospheric parameters and measurement errors.
Bottom panel: the solid line indicates the difference in
[Na/Fe] derived by using Bell et al. (1976) model atmo-
spheres with respect to Kurucz (1995) models. The dashed
line indicates the effect of NLTE vs LTE approximations
on the derived [Na/Fe] abundances.
tip of the RGB (where they can reach as much as 0.3 dex).
Neglecting these corrections would then result in a spuri-
ous trend as a function of Teff , with tip giants having
0.2-0.3 dex lower [Na/Fe] abundances than less evolved
RGB stars.
5. RESULTS AND DISCUSSION
Our final [Na/Fe] abundances (including NLTE correc-
tions) are plotted as a function of temperature in Fig. 6.
The first clear evidence that can be derived from this
plot is the large star-to-star scatter, at every luminos-
ity/temperature along the RGB. Also, the average Na
abundance is generally quite high. Neither feature is sur-
prising in itself, as both have been commonly found in
several other clusters; now they are found for the first
time also in this cluster, that had never been studied in
detail from a spectroscopic point of view before.
When compared to abundances of field stars of sim-
ilar metallicity (e.g. Gratton et al. 2000, Pilachowski et
8
E. Carretta et al.: Proton capture elements in NGC 2808
al. 1996a), these results confirm once more that globular
cluster stars present a pattern of abundances suggestive
of processes involving elements produced somewhere deep
in the star, near the H-burning shell, by proton-capture
nucleosynthesis that operates at high temperatures, and
likely producing Na through the NeNa cycle.
In fact, while field stars have generally low Na abun-
dances (almost coincident with the lower envelope of the
cluster abundances), with small scatter around the aver-
age value (comparable with observational errors), in clus-
ter stars the standard evolutionary paradigm is broken.
The spread in Na abundances in NGC 2808 reaches as
much as 1 dex and does not decrease with temperature
much below 0.5 dex, in [Na/Fe].
Apart from the well known Na-O anticorrelation (also
present among RGB stars of NGC 2808 at a level that
could be at least comparable with that existing in M 13,
Carretta et al. 2003b), our results from Na abundances
alone allow us to conclude that the standard first dredge-
up and an additional mixing mechanism at luminosities
above the RGB-bump are not enough to explain the ob-
served pattern, at odds with what happens in field stars.
Na abundances so large and scattered from star to star
do require that a variable amount of Na-enriched matter
polluted the stars in the past, or that the efficiency of
some further (deep) mixing mechanism varies along the
RGB modifying pre-existing stellar abundances. Thus, the
question is now: is the observed scatter a cosmic scatter
intrinsic to the studied sample, or is it an artifact of the
analysis due to conjuring effects of uncertainties on de-
rived abundances?
In general, we may consider three sources of errors on
abundance determinations, some of which may affect the
scatter, and some others only the zero point. First, the
adopted temperature or oscillator strength scales may be
wrong6. However, this would result only in a zero point
shift, affecting rigidly the entire metallicity distribution
but leaving almost unchanged the star-to-star spread.
Second, other effects exist, that may result in spurious
systematic trends on the results if not taken into account.
A clear example is provided by the corrections for de-
parture from LTE considered in Sect. 4.1. In the bottom
panel of Fig. 6 we have shown (dashed line) the changes
in [Na/Fe] abundances that would result by neglecting the
non-LTE corrections. Since, as shown by Fig. 5 and dis-
cussed in Gratton et al. (1999), the amount of the correc-
tion increases at lower temperatures and gravities, had we
neglected this effect a marked trend would appear in the
6 Note however that a systematic error of about 100 K in
the temperature scale would give a difference of about 0.2 dex
in abundances from neutral and singly ionized Fe lines, while
the average difference found by Carretta et al. (2003b, from
whom we adopted the average metallicity for NGC 2808) is
only −0.01 dex, implying a rather small systematic error on
the temperatures. Moreover, the atomic parameters for the
strong Na D lines are well known, as well as the laboratory
oscillator strengths for many of the Fe lines used to derive the
abundances from the UVES spectra.
results, with Na abundances decreasing as the stars ap-
proach the RGB-tip. Notice that this effect should be in
principle larger at low metallicity. This should be borne
in mind when comparing the pattern found in different
clusters, such as the survey of Na abundances in giants of
M 15 and M 92 by Sneden et al. (2000) where Na D lines
were also used, but no correction for departures from LTE
was applied.
Another effect of some impact may be the choice of
the grid of atmospheric models used in the analysis. In
the present study we adopted the Kurucz (1995; K95)
models with the overshooting option. We then repeated
the analysis by computing the Na abundances using the
Bell et al. (1976; hereinafter BEGN) model atmospheres,
that are maybe a bit better at reproducing the atmo-
spheres of cooler stars such as those at the RGB-tip. Fig. 6
shows (bottom panel solid line) that at Teff about 3900
K the BEGN models provide [Na/Fe] ratios about 0.04
dex larger than those derived with K95 models. At lower
luminosity, around Teff ∼ 4900 K, the trend is inverted,
with abundances from BEGN models being about 0.08
dex lower than those from K95. The net effect is again a
(small) trend; if one used BEGN models and no correc-
tions for non-LTE, the two opposite trends would nearly
compensate each other. Anyway, also this effect would
only tilt the distribution function of Na abundances along
the RGB, leaving the rms scatter untouched.
Finally, random errors due to uncertainties in the
adopted atmospheric parameters and to measurement er-
rors may potentially add a random noise to the observed
abundances. Are they enough to explain the observed scat-
ter? In order to test this, we evaluated the sensitivity of
abundances to errors in the adopted atmospheric parame-
ters in the usual way, i.e. we repeated the analysis chang-
ing by a given amount one parameter at a time (Teff ,
log g, [A/H] and vt). We performed this test on three typ-
ical stars spaced all along the RGB, the results of this
exercise are summarized in Table 2.
To correctly use this Table, one has to notice that the
changes in the parameters are arbitrary, and have been
adopted only for demonstrative purposes. For example,
realistic errors in Teff are not likely to exceed ∼ 50 K,
whereas in column 2 of Table 2 we have used twice this
value. Similarly, the gravity is unlikely to be wrong by
much more than 0.1 dex and a conservative estimate of er-
rors in metallicity is about 0.1 dex (Carretta et al. 2003b),
hence the values listed in column 3 and 4 should be scaled
accordingly. Moreover, since our derivation of atmospheric
parameters uses the position of stars in the CMD, errors
in Teff and log g cannot be considered independent: e.g.,
an error in Teff of about 100 K would also correspond
to an uncertainty of about 0.25-0.30 dex in log g. So the
contributions from these two terms must first be summed
up algebrically, and the result be summed in quadrature
along with all other independent contributions to give the
final error due to uncertainties in all the adopted parame-
ters. These typically range from 0.06 dex for the brightest
E. Carretta et al.: Proton capture elements in NGC 2808
9
Table 2. Sensitivities of [Na/Fe] abundances to demon-
strative errors in the atmospheric parameters
Element
∆Teff
∆ log g ∆[A/H]
+100 K +0.3 dex
∆vt
0.2 dex +0.2 km s−1
star 51871: Teff =3970 K, log g = 0.62
[Na/Fe]
+0.205
−0.105
+0.011
−0.008
star 39060: Teff =4357 K, log g = 1.28
[Na/Fe]
+0.172
−0.132
+0.028
−0.016
star 56136: Teff =4905 K, log g = 2.38
[Na/Fe]
+0.292
−0.156
+0.041
−0.017
stars in the sample up to 0.10 dex for the faintest stars
observed at the RHB level.
To these, one has to add the uncertainties in the mea-
surements. These random errors (including e.g. the con-
tinuum placement, the visual estimate of Na abundances
from comparison with synthetic spectra) can be evaluated
from the rms deviations of the average of Na abundances
given by the two Na D lines. They range typically from
0.05 dex at the RGB-tip, to 0.06 dex around 4350 K, and
to 0.07 dex for the faintest stars in the sample.
The resulting total random errors then vary from 0.08-
0.09 dex for brighter stars and increase to as much as 0.12
dex at the RHB level. The random error bars are displayed
in the upper panel of Fig. 6. Hence, the observed spread
in Na abundances is more than 5 times the error for the
bright part of the sample and as much as 8-10 times the
error when considering the lower luminosity giants.
A first, firm result of these observations is then that
large variations do exist in the abundances of elements
produced in proton-capture reactions at all luminosities
along the Red Giant Branch of NGC 2808, as shown by
the large spread in [Na/Fe] abundances.
Apart from specific computations, Fig. 7 allows to im-
mediately appreciate how much Na can differ between
pairs of stars having the same temperature, hence the
same evolutionary physical status.
6. Comparison with the Na distribution in other
Globular Clusters
The proton capture elements as Na in RGB stars have
been studied in several GC's, mainly by the Lick and
Texas groups. We may compare here our findings with
what has been derived for the following GC's ([Fe/H]
values are all on the Carretta & Gratton 1997 metal-
licity scale): M5 (Ivans et al. 2001; [Fe/H]= -- 1.11), M13
(Pilachowski et al. 1996b, Kraft et al. 1997; [Fe/H]= --
1.39), M15 and M92 (Sneden et al. 2000; [Fe/H]= -- 2.12
and -- 2.16, respectively). M4, studied by Ivans et al.
(1999), will not be considered here because severe differ-
ential reddening makes the separation of true RGB from
AGB stars very difficult.
We have taken the [Na/Fe] values from the above
quoted papers, together with the photometric information
Fig. 7. Observed spectra of pairs of stars with different
strenghts in Na D lines, at the top (3970 K), in the mid-
dle (4450 K) and at the bottom (4906 K) of the tempera-
ture range spanned by our sample. Star identifications are
given at the top of each panel, together with the [Na/Fe]
value. A few photospheric lines are indicated, while 'in-
terst.' shows the position of the interstellar Na D lines.
that comes originally from: Sandquist et al. (1996, M5),
Cudworth & Monet (1979, M13), Cudworth (1976, M15),
and Rees (1992, M92). The distance moduli (m−M )V and
reddenings were taken from Harris (1996). These [Na/Fe]
values come from moderately high (R=11000) to very high
(R=60000) resolution spectra. The analyses differ from
ours in some parts: in some cases Na abundances have
been derived from the 5682-88 A doublet (M5 and M13),
in others from the Na D lines (M15 and M92); the Bell et
al. (1976) grid of model atmosphere was usually used by
the Texas-Lick group; finally, no correction for departure
from LTE has been applied (this is slightly less important
when the 5682-88 A lines are used).
6.1. NGC 2808 and M13
The largest samples of RGB stars are for M13 (112) and
NGC 2808 (81), while less than about 30 RGB stars have
been observed in the other GC's. The first comparison
then will be done between these two GC's, since sta-
tistical significance and large sample of stars in differ-
ent evolutionary phases are required to disentangle possi-
ble evolutionary effects, if any, overimposed to primordial
variations. Moreover, Pilachowski et al. (1996b, hereafter
PSKL) and Kraft et al. (1997) made a case for evolution-
ary effects being strongly at work in M13, which is also
the template cluster as far as extreme oxygen depletion in
RGB stars is concerned.
10
E. Carretta et al.: Proton capture elements in NGC 2808
The two clusters have been observed at similar resolu-
tions (R ∼ 11000 and 15000 respectively for M13 and NGC
2808), in both cases Na abundances come from spectral
synthesis (5682-88 A lines, and Na D lines, respectively),
the metallicities are quite similar (0.25 dex less for M13),
and only true RGB stars are considered in the comparison.
We note here that 13 more RGB tip stars were ob-
served in NGC 2808, with UVES, and preliminary Na
abundances were derived from abundance analysis based
on the EW's of four lines, the subordinate lines 5862-
88 A and 6154-60 A. We do not include these stars in the
present analysis to maintain a strict homogeneity in our
database, however it's worth mentioning that they agree
very well with the present results.
To evaluate the (possible) presence of evolutionary ef-
fects we have to compare samples of stars in different
phases along the RGB. PSKL define as "RGB tip" stars
< 1, since they find an abrupt change in
those having logg ∼
the Na abundance distribution at this point, with [Na/Fe]
becoming larger and less dispersed for higher luminosity
stars. We find a similar separation, but in our case the
Na abundances decrease in the RGB tip stars. While it
is not clear if this point has some physical relevance, e.g.
marking the onset of a particular regime when the star
approaches the last stages of lifetime as a red giant, in the
following we will adopt logg = 1.02 as separator (as done
e.g., in M5 by Ivans et al. 2001) between RGB-tip and
lower-RGB stars (see also PSKL for a discussion on this
point).
However, before attempting any comparison, we have
to spend a few cautionary words on the various factors
that may introduce spurious differences. When we com-
pare results for bright and fainter objects, the larger ran-
dom errors that are intrinsic to the analysis of fainter stars
can produce an artificial increase in the abundance dis-
persion. Moreover, when we compare abundance analyses
done by different groups we also have to take into ac-
count the different tools adopted: the use of the Kurucz
or BEGN models for atmospheres, or the choice about cor-
recting or not for NLTE, both introduce a tilt (see Fig. 6)
that influences the median values of the Na distributions.
Fig. 8 (lower panels) shows the distribution along the
RGB's of the observed stars for NGC 2808 and M13, in
the dereddened (B-V)0, MV plane, while in the upper pan-
els are plotted the distributions of the [Na/Fe] ratios for
RGB tip and lower RGB stars; the vertical lines indicate
the median value of each distribution. NGC 2808 and M13
appear to behave quite differently in both bins/ranges. As
already noted by PSKL, in M13 the brightest stars are
skewed towards high Na abundances (the median of the
[Na/Fe] values is +0.41 dex) and show a small dispersion,
while among the lower RGB stars there is a number of ob-
jects that are more Na-poor (median +0.16 dex) and with
a larger dispersion. Notice that the - small - differential
corrections (needed to translate abundances derived with
BEGN models to our own scale) between the two bins
would only increase the difference, shifting the medians
apart.
Fig. 8. Bottom panels: CMD's for the RGB stars studied
in NGC 2808 and M13; filled and open circles represent
stars with log g ≤ and > 1.02, respectively; crosses indi-
cate objects not considered in the histograms because they
are possible AGB stars. Upper panels: histograms for the
[Na/Fe] values, for log g ≤ and > 1.02 respectively, nor-
malized to the total number of stars used in the adopted
range (labelled in the box). For NGC 2808 we plot both
the LTE and NLTE cases. The vertical lines (dotted: NGC
2808 LTE; solid: NGC 2808 NLTE; dashed-dotted: M 13)
indicate the median values for each considered range and
case; the value for the NLTE distributions in NGC 2808
are shown also in the M13 panels for comparison.
In NGC 2808 the opposite seems to be true: the RGB
tip stars are Na-poorer, on average: a median value of
0.13 dex is found, to be compared with 0.40 dex for lower
luminosity RGB stars. When running simple Kolmogorov-
Smirnov tests, these differences (between the two bins for
the same GC, and between GC's) appear significant: the
two ranges in log g (or luminosity and/or temperature)
have very low probability of being extracted from the same
parent population. Moreover, both ranges seem to be sta-
tistically different between M13 and NGC 2808. While a
larger spread at hotter temperatures is expected because
of the worsening of spectra quality at lower luminosities,
it is the median value that matters, as far as the interpre-
tation of this data is concerned.
The interpretation by PSKL of this pattern of abun-
dances in M13 was that, in addition to the primordial
variations among stars, an evolutionary effect is present,
that increases the Na abundance somewhat differently for
each star, since the mixing efficiency is different from stars
to star. This evolutionary Na enhancement is dominant in
the upper RGB, as witnessed also by the lowering of car-
bon isotopic ratios and carbon abundances (accompanied
E. Carretta et al.: Proton capture elements in NGC 2808
11
In our opinion, the lack of a clear peak at high Na
values in the distribution for NGC 2808 is an indication
that the possible link between HB morphology (extension
and mass distribution) and other parameters such as the
extent of chemical anomalies may be not very tight.
6.2. Comparison with other clusters: M5, M15, M92
As a result of the above comparison, can we deduce that
NGC 2808 is a peculiar GC? Or is instead M13 that be-
haves unusually? Already Ivans et al. (2001) in their study
on M5 noted that this cluster does not behave like M13
(see their fig. 11). To M5 we may add M15 and M92:
even with smaller numbers, the samples in these three
GCs reach down to the same luminosity level (MV ∼ 0) of
the bulk of our program stars, so we feel entitled to repeat
the comparison.
The results are displayed in Fig. 9. In all three cases
the [Na/Fe] distribution does not vary significantly (as
deduced again by the Kolmogorov-Smirnov test) between
the two bins (upper and lower RGB), and the median
values are quite similar, in particular for the lower RGB.
Again, based on this test, M13 appears to stand out among
globular clusters, since none of the examined clusters
shows a clearcut shift toward Na-rich stars when climbing
up the RGB. This is not surprising, as M13 is peculiar in
other ways, like its anomalously extendend Na-O anticor-
relation. It seems that in every other cluster, the scatter in-
troduced by possible evolutionary effects is very small and
does not affect sensibly the resulting abundances. This is
more evident in the 3 other clusters shown in Fig. 9, where
stars are observed preferentially on the upper part of the
RGB. An increase of the scatter is hardly detectable.
7. SUMMARY
We have analysed moderately high-resolution spectra of
more than 80 red giant stars in the globular cluster
NGC 2808, taken during the Science Verification program
of the FLAMES multi-object spectrograph at the VLT.
Atmospheric parameters were derived from the photomet-
ric information, radial velocities (which permitted to ex-
clude a few field interlopers, leaving a sample of 81 RGB
stars) were derived from the spectra. We examined in this
paper only the Sodium abundances, that were derived
from the Na i D lines and spectral synthesis, and cor-
rected for NLTE effects. The main results of our analysis
are:
1- There are large variations in Na abundances at all
luminosity levels along the RGB in NGC 2808. This is an
intrinsic spread, since it is much larger than possible un-
certainties in the derived abundances (being from 0.5 to 1
dex, to compare with errors of about 0.08 dex increasing to
0.12 dex at the very faint end of our sample). While this is
not surprising, being commonly observed in several glob-
ular clusters, this is the first time that detailed abundance
analysis uncovers this pattern in NGC 2808. This spread
is not explained by standard stellar evolutionary models,
Fig. 9. The same as Fig. 8, but for M5, M15, and M92.
by increasing nitrogen abundances climbing up the RGB
in several clusters, see the review by Kraft 1994 for ref-
erences). Instead, in NGC 2808 this enhancement of the
number of Na-rich stars is not seen, and the variation goes
the opposite way, as in other clusters (see below).
The conclusions of this comparison are that:
i) in NGC 2808 the evolutionary effect claimed in M 13-
if at all present - is only at the noise level. This would
imply that the Na abundance distribution along the RGB
is most probably of primordial origin, either because it was
imprinted in the gas from which these stars originated, or
because it was produced by an early pollution event due
to a first generation of intermediate-mass AGB stars;
ii) this finding seems to be quite at odds with what one
may expect from the distribution of stars on the HB, which
is rather peculiar in NGC 2808 (an important example of
"second-parameter" effect). Gratton (1982) already noted
that about one fourth of the HB stars lie on the blue
HB, which is quite unexpected given the metallicity of
the cluster.
On the working hypothesis (recently explored by
D'Antona et al. 2002) that Na-rich stars are also enriched
in Helium (as well as depleted in Oxygen, as found in
NGC 2808 by Carretta et al. 2003b), they would populate
exactly the blue part of the HB. This, however, would re-
quire that about 1/4 of the stars in the upper RGB stand
out as Na-rich objects, which is maybe observed in M13
but not very clearly in NGC 2808. The scenario proposed
by D'Antona et al. (2002) also implies that more than
one episode of star formation may have occurred within a
GC, producing a bi(multi)-modal mass distribution on the
HB. A similar type of distribution in the Na abundances
should then be present along the RGB, which however is
not observed with a sufficient degree of confidence.
12
E. Carretta et al.: Proton capture elements in NGC 2808
Cudworth, K.M. 1976, AJ, 81, 519
Cudworth, K.M. & Monet, D.G. 1979, AJ, 84, 774
D'Antona, F., Caloi, V., Montalban, J, Ventura, P.& Gratton,
R.G. 2002, A&A, 395, 69
D'Antona, F., Gratton, R.G., & Chieffi, A. 1983, MSAIt, 54,
173
Denisenkov, P.A., Denisenkova, S.N. 1990, Soviet. Astron.
Lett., 16, 275
Gratton, R.G. 1982, A&A, 115, 171
Gratton, R.G., Bonifacio, P., Bragaglia, A., et al. 2001, A&A,
369, 87
Gratton, R.G., Carretta, E., Eriksson, K., & Gustafsson, B.
1999, A&A 350, 955
Gratton, R.G., Sneden, C., Carretta, E., & Bragaglia, A. 2000,
A&A, 354, 169
Harris, W.E. 1996, AJ, 112, 1487
Ivans, I., Kraft, R.P., Sneden, C., et al., 2001, AJ, 122, 1438
Ivans, I., Sneden, C., Kraft, R.P., et al., 1999, AJ, 118, 1273
Kraft, R.P. 1994, PASP, 106, 553
Kraft, R.P, Sneden, C., Smith, G.H., Shetrone, M.D., Langer,
G.E., & Pilachowski, C.A. 1997, AJ, 113, 279
Kurucz, R.L. 1995, CD-ROM 13, Smithsonian Astrophysical
Observatory, Cambridge
Langer, G.E., Hoffman, R., & Sneden, C. 1993, PASP, 105, 301
Lyons, M.A., Kemp, S.N., Bates, B. & Shaw, C.R., 1996,
MNRAS, 280, 835
McWilliam, A., Preston, G.W., Sneden, C. & Searle, L. 1995,
AJ, 109, 2757
Osborn, W. 1971, Observatory, 91, 223
Pasquini, L., Avila, G., Blecha, A. et al., 2002, The ESO
Messenger No. 110, p.1
Pilachowski, C.A., Sneden, C. & Kraft R.P. 1996a, AJ, 111,
1689
Pilachowki, C.A., Sneden, C., Kraft, R.P. & Langer, G.E.,
1996b, AJ, 112, 545
Rees, R. 1992, AJ, 103, 1573
Sandquist, E.L., Bolte, M., Stetson, P.B. & Hesser, J.E. 1996,
ApJ, 470, 910
Smith, G.H. 1987, PASP, 99, 67
Sneden, C. Pilachowski, C.A. & Kraft, R.P. 2000, 120, 1351
Ventura, P. D'Antona, F., Mazzitelli, I., & Gratton, R. 2001,
ApJ, 550, 65
Walker, A.R., 1999, AJ, 118, 432
and its presence requires variable amount of yields from
proton capture fusion on Neon, that can only happen in
the interior region near the H-burning shell, in the same
star examined or in other objects that were then able to
somehow pollute its surface.
2- We have compared RGB-tip to lower-RGB stars to
see whether there is any evolutionary effect acting like in
M13, and enhancing Na for the brighter, more evolved
giants. This effect does not appear to be important in
NGC 2808, since brighter RGB stars have lower average
Na abundances; if a similar mechanism is at work in this
cluster, it only contributes adding some noise to a primor-
dial dispersion. Our data are not sufficient to distinguish if
the different composition was pre-existing the star forming
gas, or is due to pollution on the star surface.
3- This Na abundance dispersion, and the (absence
of) variation towards the RGB tip is similar to what was
found for other GCs, e.g. M5, M15, and M92, but at odds
with what is seen in M13, the only other cluster for which
a very large sample of RGB stars has been observed.
Homogeneous analysis of large samples of stars ob-
served at high resolution and high S/N is now possible for
all Galactic globular clusters, with the new multiobject
spectrographs mounted on 8-10m class telescopes. This is
the road we have to follow if we wish to make significative
improvements on our knowledge of stellar formation and
evolution in globular clusters.
Acknowledgements. This research has made use of data taken
during FLAMES Science Verification. We wish to warmly
thank Luca Pasquini for having co-ordinated the building of
such a powerful instrument, and the ESO staff at Paranal for
their excellent work during SV. We are specially indebted to
Raffaele Gratton for many interesting and helpful suggestions
and discussions that greatly improved the present work. We
also thank the referee (Bob Rood) for his suggestions.
References
Alonso, A., Arribas, S., & Martinez-Roger, C. 1999, A&AS,
140, 261
Alonso, A., Arribas, S., & Martinez-Roger, C. 2001, A&A, 376,
1039
Bell, R.A., Eriksson, K., Gustafsson, B. & Nordlund A. 1976,
A&AS, 23, 37
Bragaglia, A., Carretta, E., Gratton, R.G., et al. 2001, AJ, 121,
327
Bruls, J.H.M., Rutten, R.J., & Shchukina, N.G. 1992, A&A,
265, 237
Cacciari, C., Bragaglia, A., Rossetti, E., et al., 2003, in prepa-
ration
Cannon, R.D., Croke, B.F.W., Bell, R.A., Hesser, J.E., &
Stathakis, R.A. 1998, MNRAS, 298, 601
Cardelli, J.A., Clayton, G.C. & Mathis, J.S., 1989, ApJ, 345,
245
Carretta, E., & Gratton, R.G. 1997, A&AS, 121, 95
Carretta, E., Gratton, R.G.,& Lucatello, S. 2003a, in prepara-
tion
Carretta, E., Bragaglia, A., Cacciari, C., Mulas, G. & Rossetti,
E. 2003b, in preparation
Cottrell, P.L., & Da Costa, G.S. 1981, ApJ, 245, L79
|
0809.5028 | 2 | 0809 | 2008-10-29T23:01:21 | WMAP 5-year constraints on time variation of $\alpha$ and $m_e$ in a detailed recombination scenario | [
"astro-ph"
] | We study the role of fundamental constants in an updated recombination scenario. We focus on the time variation of the fine structure constant, and the electron mass in the early Universe, and put bounds on these quantities by using data from CMB including WMAP 5-yr release and the 2dFGRS power spectrum. We analyze how the constraints are modified when changing the recombination scenario. | astro-ph | astro-ph |
WMAP 5-year constraints on time
variation of α and me in a detailed
recombination scenario
Claudia G. Sc´occola1,2∗, Susana. J. Landau3†, Hector Vucetich1 ,
1 Facultad de Ciencias Astron´omicas y Geof´ısicas. Universidad Nacional de La Plata,
Paseo del Bosque S/N 1900 La Plata, Argentina.
2 Instituto de Astrof´ısica La Plata - CONICET
3 Departamento de F´ısica, FCEyN, Universidad de Buenos Aires,
Ciudad Universitaria - Pab. 1, 1428 Buenos Aires, Argentina.
June 22, 2021
Abstract
We study the role of fundamental constants in an updated recom-
bination scenario. We focus on the time variation of the fine structure
constant α, and the electron mass me in the early Universe, and put
bounds on these quantities by using data from CMB including WMAP
5-yr release and the 2dFGRS power spectrum. We analyze how the
constraints are modified when changing the recombination scenario.
Time variation of fundamental constants is a prediction of theories that
attempt to unify the four interactions in nature. Many efforts have been
made to put observational and experimental constraints on such variations.
Primordial light elements abundances produced at Big Bang nucleosynthesis
(BBN) and cosmic microwave background radiation (CMB) are the most
powerful tools to study the early universe and in particular, to put bounds
∗fellow of CONICET
†member of the Carrera del Investigador Cient´ıfico y Tecnol´ogico CONICET
1
on possible variations of the fundamental constants between those early times
and the present.
Previous analysis of CMB data (earlier than the WMAP five-year release)
including a possible variation of α have been performed by refs. [1, 2, 3, 4]
and including a possible variation of me have been performed by refs. [3, 4, 5].
In March 2008, WMAP team released data collected during the last five years
[6]. Moreover, new processes relevant during the recombination epoch have
been taken into account.
Indeed, in the last years, helium recombination
has been studied in great detail [7, 8, 9, 10], revealing the importance of
these considerations on the calculation of the recombination history. Switzer
& Hirata [8] presented a multi-level calculation for neutral helium recombi-
nation including, among other processes, the continuum opacity from H I
photoionization. They found that at z < 2200 the increasing H I abundance
begins to absorb photons out of the He I 21p → 11s line, rapidly accelerating
He I recombination, which finishes at z ∼ 1800. Kholupenko et al [11] have
considered the effect of the neutral hydrogen on helium recombination and
proposed an approximated formula to take this effect into account. This has
enabled its implementation on numerical codes such as Recfast, since the
complete calculations done by Switzer & Hirata take a large amount of com-
putational time. Another improvement in the recombination scenario is the
inclusion of the semi-forbidden transition 23p → 11s, the feedback from spec-
tral distorsions between 21p → 11s and 23p → 11s lines, and the radiative
line transfer.
The release of new data from WMAP brings the possibility of updating
the constraints on the time variation of fundamental constants. In this paper
we study the variation of α and me in the improved recombination scenario.
It could be argued that me is not a fundamental constant in the same sense as
α is and that constraints on the Higgs vaccum expectation value (< v >) are
more relevant than bounds on me. However, in the recombination scenario
the only consequence of the time variation of < v > is a variation in me.
The effect of a possible variation of α and/or me in the recombination
scenario and in the CMB temperature and polarization spectra has been
analized previously [12, 13, 14, 5]. Here we focus in the effect of the variation
of α and me on the improved recombination scenario.
The recombination equations implemented in Recfast in the detailed
recombination scenario [15] including the fitting formulae of [11] are:
2
H(z)(1 + z)
dxp
dz
= (cid:16)xexpnHαH − βH(1 − xp)e−hνH2s/kTM(cid:17)CH,
(1)
=(cid:16)xHeIIxenHαHeI − βHeI(fHe − xHeII)e−hνHeI,21 s/kTM(cid:17) CHeI
+ xHeIIxenHαt
HeI(fHe − xHeII)e−hνHeI,23 s/kTM! C t
gHeI,23s
gHeI,11s
HeI −
βt
HeI ,
H(z)(1 + z)
dxHeII
dz
where
CH =
(2)
,
(3)
(4)
(5)
1 + KHΛHnH(1 − xp)
1 + KH(ΛH + βH)nH(1 − xp)
,
CHeI =
C t
HeI =
1 + KHeIΛHenH(fHe − xHeII)ehνps/kTM
1 + KHeI(ΛHe + βHeI)nH(fHe − xHeII)ehνps/kTM
1 + K t
HeInH(fHe − xHeII)ehνt
HeIβt
ps/kTM
1
.
The last term in eq. (2) accounts for the recombination through the
triplets by including the semi-forbidden transition 23p → 11s. As remarked
in [15], αt
HeI is fitted with the same functional form used for the αHeI singlets,
with different values for the parameters, so the dependences on the funda-
mental constants are the same, being proportional to α3m−3/2
. The two
photon transition rates ΛH and ΛHe depend on the fundamental constants
as α8me. The photoionization coefficients β are calculated as usual from the
recombination coefficients αc (with c standing for H, HeI and HeII), so their
dependencies are known (see for example [4]).
e
The KH, KHeI and K t
HeI quantities are the cosmological redshifting of the
H Ly α, He i 21p -- 11s and He i 23p -- 11s transition line photons, respectively.
In general, K and the Sobolev escape probability pS are related through the
following equations (taking He i as an example):
KHeI =
K t
HeI =
gHeI,11s
gHeI,21p
gHeI,11s
gHeI,23p
1
21p−11spS
1
23p−11spS
nHeI,11sAHeI
nHeI,11sAHeI
and
,
(6)
(7)
3
where AHeI,21p−11s and AHeI,23p−11s are the Einstein A coefficients of the He I
21p -- 11s and He I 23p -- 11s transitions, respectively. To include the effect of the
continuum opacity due to H, based on the approximate formula suggested
by ref.
[11], pS is replaced by the new escape probability pesc = ps + pcon,H
with
1
pcon,H =
1 + aHeγbHe
,
(8)
(9)
and
γ =
gHeI,11 s
gHeI,21 p
AHeI
21p−11s(fHe − xHeII)c2
8π3/2σH,1s(νHeI,21p)ν2
HeI,21p∆νD,21p(1 − xp)
where σH,1s(νHeI,21p) is the H ionization cross-section at frequency νHeI,21p
and ∆νD,21p = νHeI,21pq2kBTM/mHec2 is the thermal width of the He i 21p --
11s line. The cross-section for photo-ionization from level n is [16]:
σn(Z, hν) =
26απa2
0
3√3
n
Z 2 (1 + n2ǫ)−3gII(n, ǫ)
(10)
where gII(n, ǫ) ≃ 1 is the Gaunt-Kramers factor, and a0 = ¯h/(mecα) is the
Bohr radius, so σH,1s(νHeI,21p) is proportional to α−1m−2
e .
The transition probability rates AHeI,21p−11s and AHeI,23p−11s can be ex-
pressed as follows [17]:
AHeI
i−j =
4α
3c2 ω3
ij hψir1 + r2ψji2
(11)
where ωij is the frequency of the transition, and i(j) refers to the initial
(final) state of the atom. First we will analize the dependence of the bra-ket.
To first order in perturbation theory, all wavefunctions can be approximated
to the respective wavefunction of hydrogen. Those can be usually expressed
as exp(−qr/a0) where a0 is the Bohr radius and q is a number. It can be
shown that any integral of the type of Eq. (11) can be solved with a change
of variable x = r/a0.
If the wave functions are properly normalized, the
dependence on the fundamental constants comes from the operator, namely
r1 + r2. Thus, the dependence of the bra-ket goes as a0. On the other hand,
ωij is proportional to the difference of energy levels and thus its dependence
on the fundamental constants is ωij ≃ meα2. Consequently, the dependence
of the transition probabilities of HeI on α and me is
4
AHeI
i−j ≃ meα5
(12)
new recombination scenario
standard recombination scenario
1.2
1
0.8
e
n
0.6
0.4
0.2
0
3000
2500
2000
1500
z
1000
500
0
Ionization history allowing α to vary with time. From left to
Figure 1:
right, the values of α
are 1.05, 1.00, and 0.95, respectively. The dotted
α0
lines correspond to the standard recombination scenario, and the solid lines
correspond to the updated one.
In Fig. 1 we show how a variation in the value of α at recombination
affects the ionization history, moving the redshift at which recombination
occurs to earlier times for larger values of α. The difference between the
functions when the two different recombination scenarios are considered, for
a given value of α, is smaller than the difference that arise when varying the
value of α. Something similar happens when varying me.
With regards to the fitting parameters aHe and bHe, since detailed cal-
culation of their values are not available yet, it is not possible to determine
the effect that a variation of α or me would have on these new parameters.
Wong et al [15] have shown that they must be known at the 1% level for
future Planck data.
In this letter, however, we are dealing with the 5 yr
data from WMAP satellite and this accuracy is not required. To come to
this conclusion, we have calculated the temperature, polarization and cross
correlation CMB spectra, allowing the parameters aHe and bHe to vary at the
50% level. We found that for the temperature and polarization spectra, the
variation is always lower than the observational error (1% for temperature
and almost 40 % in polarization). The largest variations occur in the cross
5
).
correlation CMB spectra (C T E
In this case, we have calculated the ob-
servational errors divided by the value of the Cℓ's of all measured C T E
and
compared them with the relative variation in the Cl's induced when changing
aHe and bHe by a 50%. In all of the cases the first quantity is several orders
of magnitude greater than the variation of the Cℓ's. Therefore, in order to
analyze WMAP5 data, there is no need to modify these parameters.
ℓ
l
To put constraints on the variation of α and me during recombination
time in the detailed recombination scenario studied here, we introduced the
dependencies on the fundamental constants explicitly in the latest version
of Recfast code [18], which solves the recombination equations. We per-
formed our statistical analysis by exploring the parameter space with Monte
Carlo Markov chains generated with the publicly available CosmoMC code
of ref. [19] which uses the Boltzmann code CAMB [20] and Recfast to
compute the CMB power spectra. We modified them in order to include
the possible variation of α and me at recombination. We ran eight Markov
chains and followed the convergence criterion of ref. [21] to stop them when
R − 1 < 0.0180. Results are shown in Table 1 and Figure 2.
The observational set used for the analysis was data from the WMAP
5-year temperature and temperature-polarization power spectrum [6], and
other CMB experiments such as CBI [22], ACBAR [23], and BOOMERANG
[24, 25], and the power spectrum of the 2dFGRS [26]. We have considered
a spatially-flat cosmological model with adiabatic density fluctuations, and
the following parameters:
P = ΩBh2, ΩCDM h2, Θ, τ,
∆α
α0
,
∆me
(me)0
, ns, As! ,
(13)
where ΩCDM h2 is the dark matter density in units of the critical density, Θ
gives the ratio of the comoving sound horizon at decoupling to the angular
diameter distance to the surface of last scattering, τ is the reionization optical
depth, ns the scalar spectral index and As is the amplitude of the density
fluctuations.
In Fig. 2 we show the marginalized posterior distributions for the cosmo-
logical parameters, ∆α/α0, and ∆me/(me)0, which are the variation in the
values of those fundamental constants between recombination epoch and the
present. The three succesively larger two dimensional contours in each panel
correspond to the 68%-, 95%-, and 99%- probability levels, respectively. In
the diagonal, the one dimensional likelihoods show the posterior distribution
of the parameters.
6
1
0.8
0.6
0.4
0.2
0
0.13
0.09
1.15
1.1
1.05
1
0.95
0.11
0.05
x
a
m
L
/
L
2
h
M
D
C
Ω
Θ
τ
0.05
0
α
/
α
∆
0
e
m
/
e
m
∆
s
n
s
A
0
-0.05
0.1
0
-0.1
0.98
0.93
3.1
2.9
0.02
0.024
2
Ω
Bh
0.09
Ω
CDMh
0.13
2
0.95 1 1.05 1.1 1.15
Θ
0.05
0.11
τ
-0.05
0.05
0
∆ α / α
0
0
-0.1
∆me/me0
0.1
0.93
0.98
ns
2.9
3.1
As
Figure 2: Marginalized posterior distributions obtained with CMB data, in-
cluding the WMAP 5-year data release plus 2dFGRS power spectrum. The
diagonal shows the posterior distributions for individual parameters, the
other panels shows the 2D contours for pairs of parameters, marginalizing
over the others.
In Table 1 we show the results of our statistical analysis, and compare
them with the ones we have presented in ref. [4], which were obtained in the
standard recombination scenario (i.e. the one described in [27], which we
denote PS), and using WMAP3 [28, 29] data. The constraints are tighter
in the current analysis, which is an expectable fact since we are working
with more accurate data from WMAP. The bounds obtained are consistent
with null variation, for both α and me, but in the present analysis, the 68%
confidence limits on the variation of both constants have changed.
In the
case of α, the present limit is more consistent with null variation than the
previous one, while in the case of me the single parameters limits have moved
toward lower values. To study the origin of this difference, we perform an-
other statistical analysis, namely the analysis of the standard recombination
7
parameter
wmap5 + NS
wmap5 + PS wmap3 + PS
Ωbh2
ΩCDM h2
Θ
τ
∆α/α0
∆me/(me)0
ns
As
H0
0.02241+0.00084
−0.00084
0.1070+0.0078
−0.0078
1.033+0.023
−0.023
0.0870+0.0073
−0.0081
0.004+0.015
−0.015
-0.0193+0.049
−0.049
0.962+0.014
−0.014
3.053+0.042
−0.041
0.02242+0.00086
−0.00085
0.1071+0.0080
−0.0080
1.03261+0.024
−0.023
0.0863+0.0077
−0.0084
0.003+0.015
−0.015
-0.017+0.051
−0.051
0.963+0.015
−0.015
3.052+0.043
−0.043
0.0218+0.0010
−0.0010
0.106+0.011
−0.011
1.033+0.028
−0.029
0.090+0.014
−0.014
-0.023+0.025
−0.025
0.036+0.078
−0.078
0.970+0.019
−0.019
3.054+0.073
−0.073
70.3+5.9
−5.8
70.3+6.1
−6.0
70.4+6.6
−6.8
Table 1: Mean values and 1σ errors for the parameters including α and me
variations. NS stands for the new recombination scenario, and PS stands for
the previous one.
scenario (PS) together with WMAP5 data and the other CMB data sets and
the 2dFGRS power spectrum. The results are also shown in Table 1. We
see that the changes in the results are due to the new WMAP data set, and
not to the new recombination scenario. In Fig. 3 we compare the probability
distribution for ∆α/α0 and also for ∆me/(me)0, in different scenarios and
with different data sets.
In Fig. 4 we compare the 95%- probability contour level for the parame-
ters, and their one dimensional distributions, for two different analysis in the
standard recombination scenario, namely the one with WMAP5 data (dashed
lines) and the one with WMAP3 data (solid lines). The contours are smaller
in the former case, which is expectable since that data set is more accurate.
For the fundamental constants, the contours notably shrink. Moreover, the
constraints are shifted to a region of the parameter space closer to that of null
variation in the case of α. On the other hand, limits on the variation of me
are shifted to negative values, but still consistent with null variation. From
the one dimensional likelihoods we see that the the peak of the likelihood has
moved for Ωbh2. The obtained results for the cosmological parameters are
8
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
-0.06
-0.04
-0.02
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
x
a
m
L
L
/
x
a
m
L
L
/
WMAP 3
WMAP 5
new scenario
standard scenario
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
x
a
m
L
L
/
0
0.02
∆ α / α
0
0.04
0.06
0.08
0
-0.15
-0.1
-0.05
0
0.05
0.1
∆ α / α
0
WMAP 3
WMAP 5
new scenario
standard scenario
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
x
a
m
L
L
/
0
-0.2
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0
-0.3
-0.2
-0.1
me / (me)0
0
0.1
me / (me)0
0.2
0.3
0.4
Figure 3: One dimensional likelihood for ∆α
(lower
α0
for WMAP5 data and two different recombination scenarios.
row). Left:
Right: comparison for the standard recombination scenario, between the
WMAP3 and WMAP5 data sets.
(upper row) and ∆me
(me)0
in agreement within 1σ with the ones obtained by the WMAP collaboration
[30], without considering variation of fundamental constants.
Acknowledgments
Support for this work was provided by PIP 5284 CONICET. Numerical
calculations were performed in KANBALAM cluster at Universidad Na-
cional Aut´onoma de M´exico. The authors would like to thank E.C Flores,
I.R.Leobardo, J.L.Gordillo Ruiz and S. E. Frausto del Rio for technical sup-
port for the KANBALAM facilities.
References
[1] Martins, C. J., Melchiorri, A., Trotta, R., Bean, R., Rocha, G., Avelino,
P. P.,& Viana, P. T. 2002, Phys.Rev. D66, 023505
9
x
a
m
L
/
L
2
h
M
D
C
Ω
Θ
τ
0
α
/
α
∆
0
e
m
/
e
m
∆
s
n
s
A
1
0.8
0.6
0.4
0.2
0
0.13
0.09
1.1
1
0.16
0.085
0.01
0.05
0
-0.05
-0.1
0.3
0.15
0
-0.15
1.03
0.98
0.93
3.3
3.1
2.9
0.02
Ω
Bh
0.024
2
0.09
Ω
CDMh
0.13
2
1
1.1
Θ
0.01 0.085 0.16
τ
-0.1 -0.05 0 0.05
∆ α / α
0
-0.15 0 0.15 0.3
∆me/me0
0.93 0.98 1.03
2.9
3.1
3.3
ns
As
Figure 4: Comparison between the 95%- confidence levels of WMAP3 (solid
line) with those of WMAP5 (dashed line). In the diagonal, we compare the
one dimensional likelihoods in these two cases.
[2] Rocha, G., Trotta, R., Martins, C. J. A. P., Melchiorri, A., Avelino,
P. P., & Viana, P. T. P. 2003, New Astronomy Review, 47, 863
[3] Ichikawa, K., Kanzaki, T., & Kawasaki, M. 2006, Phys.Rev. D74, 023515
[4] Landau, S.J., Mosquera, M.E., Sc´occola, C. and Vucetich, H. 2008,
Phys.Rev. D78, 083527
[5] Yoo, J. J., & Scherrer, R. J. 2003, Phys.Rev. D67, 043517
[6] Nolta, M. R., et al. 2008, ArXiv e-prints, 803, arXiv:0803.0593
[7] Dubrovich, V. K., & Grachev, S. I. 2005, Astronomy Letters, 31, 359
[8] Switzer, E. R., & Hirata, C. M. 2008, Phys.Rev. D77, 083006
10
[9] Hirata, C. M., & Switzer, E. R. 2008, Phys.Rev. D77, 083007
[10] Switzer, E. R., & Hirata, C. M. 2008, Phys.Rev. D77, 083008
[11] Kholupenko, E. E., Ivanchik, A. V., & Varshalovich, D. A. 2007,
Mon.Not.Roy.Astron.Soc. 378, L39
[12] Kaplinghat, M., Scherrer, R. J., & Turner, M. S. 1999, Phys.Rev. D60,
023516
[13] Hannestad, S. 1999, Phys.Rev. D60, 023515
[14] Kujat, J., & Scherrer, R. J. 2000, Phys.Rev. D62, 023510
[15] Wong, W. Y., Moss, A., & Scott, D. 2008, Mon.Not.Roy.Astron.Soc.
386, 1023
[16] Seaton, M. J. 1959, Mon.Not.Roy.Astron.Soc. 119, 81
[17] Drake, G. W. F., & Morton, D. C. 2007, Astrophys.J.Suppl.Ser. 170,
251
[18] Seager, S., Sasselov, D. D., & Scott, D. 1999, Astrophys.J.Letters 523,
L1
[19] Lewis, A., & Bridle, S. 2002, Phys.Rev. D66, 103511
[20] Lewis, A., Challinor, A., & Lasenby, A. 2000, Astrophys.J.538, 473
[21] Raftery, A.E. & Lewis, S.M. (1992) In Bayesian Statistics, (Bernaro,
J.M.,ed), p.765. OUP.
[22] Readhead, A. C. S., et al. 2004, Astrophys.J.609, 498
[23] Kuo, C. L., et al. 2004, Astrophys.J.600, 32
[24] Piacentini, F., et al. 2006, Astrophys.J.647, 833
[25] Jones, W. C., et al. 2006, Astrophys.J.647, 823
[26] Cole, S., et al. 2005, Mon.Not.Roy.Astron.Soc. 362, 505
[27] Seager, S., Sasselov, D. D., & Scott, D. 2000, Astrophys.J.Supp.Ser. 128,
407
11
[28] Hinshaw, G., et al. 2007, Astrophys.J.Supp.Ser. 170, 288
[29] Page, L., et al. 2007, Astrophys.J.Supp.Ser. 170, 335
[30] Dunkley, J., et al. 2008, ArXiv e-prints, 803, arXiv:0803.0586
12
|
astro-ph/0403027 | 1 | 0403 | 2004-03-01T14:25:33 | Density waves in the shearing sheet IV. Interaction with a live dark halo | [
"astro-ph"
] | It is shown that if the self-gravitating shearing sheet, a model of a patch of a galactic disk, is embedded in a live dark halo, this has a strong effect on the dynamics of density waves in the sheet. I describe how the density waves and the halo interact via halo particles either on orbits in resonance with the wave or on non-resonant orbits. Contrary to expectation the presence of the halo leads to a very considerable enhancement of the amplitudes of the density waves in the shearing sheet. This effect appears to be the equivalent of the recently reported enhanced growth of bars in numerically simulated stellar disks embedded in live dark halos. Finally I discuss the transfer of linear momentum from a density wave in the sheet to the halo and show that it is mediated only by halo particles on resonant orbits. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
4
0
0
2
r
a
M
1
1
v
7
2
0
3
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Density waves in the shearing sheet
IV. Interaction with a live dark halo
B. Fuchs
Astronomisches Rechen -- Institut, Monchhofstrasse 12 -- 14, 69120 Heidelberg, Germany
Received 2003; accepted
Abstract. It is shown that if the self -- gravitating shearing sheet, a model of a patch of a galactic disk, is embedded
in a live dark halo, this has a strong effect on the dynamics of density waves in the sheet. I describe how the
density waves and the halo interact via halo particles either on orbits in resonance with the wave or on non-
resonant orbits. Contrary to expectation the presence of the halo leads to a very considerable enhancement of
the amplitudes of the density waves in the shearing sheet. This effect appears to be the equivalent of the recently
reported enhanced growth of bars in numerically simulated stellar disks embedded in live dark halos. Finally I
discuss the transfer of linear momentum from a density wave in the sheet to the halo and show that it is mediated
only by halo particles on resonant orbits.
Key words. galaxies: kinematics and dynamics -- galaxies: spiral
1. Introduction
The shearing sheet (Goldreich & Lynden -- Bell 1965, Julian
& Toomre 1966) model has been developed as a tool to
study the dynamics of galactic disks and is particularly
well suited to describe theoretically the dynamical mech-
anisms responsible for the formation of spiral arms. For
the sake of simplicity the model describes only the dynam-
ics of a patch of a galactic disk. It is assumed to be in-
finitesimally thin and its radial size is assumed to be much
smaller than the disk. Polar coordinates can be therefore
rectified to pseudo-Cartesian coordinates and the velocity
field of the differential rotation of the disk can be approx-
imated by a linear shear flow. These simplifications allow
an analytical treatment of the problem, which helps to
clarify the underlying physical processes operating in the
disk.
In the present paper of this series I discuss the dynam-
ical effects if the shearing sheet is immersed in a live dark
halo. Dark halos are usually thought to stabilize galac-
tic disks against non-axisymmetric instabilities. This was
first proposed by Ostriker & Peebles (1973) on the basis
of -- low -- resolution -- numerical simulations. Their phys-
ical argument was that the presence of a dark halo re-
duces the destabilizing self -- gravity of the disks. Doubts
about an entirely passive role of dark halos were raised
by Toomre (1977), but he (Toomre 1981) also pointed out
that a dense core of a dark halo may cut the feed -- back loop
of the corotation amplifier of bars or spiral density waves
and suppress thus their growth. Recent high-resolution nu-
merical simulations by Debattista & Sellwood (2000) and
Athanassoula (2002, 2003) have shown that quite the re-
verse, a destabilization of disks immersed in dark halos,
might be actually true. Athanassoula (2002) demonstrated
clearly that much stronger bars grow in the simulations if
the disk is embedded in a live dark halo instead of a static
halo potential. This is attributed to angular momentum
transfer from the bar to the halo via halo particles on res-
onant orbits. Angular momentum exchange between disk
and halo has been addressed since the pioneering work of
Weinberg (1985) in many studies theoretically or by nu-
merical simulations and I refer to Athanassoula (2003) for
an overview of the literature. Toomre (1981) has shown
how the bar instability can be understood as an interfer-
ence of spiral density waves in a resonance cavity between
the corotation amplifier and an inner reflector (cf. also
Fuchs 2004). Thus it is to be expected that a live dark
halo will be also responsive to spiral density waves and
develop wakes, which I investigate here using the shearing
sheet model.
2. Boltzmann equations
The evolution of the distribution function of the disk stars
in phase space is described by the linearized 4 -- dimensional
Boltzmann equation
Correspondence to: [email protected]
∂fd1
∂t
+ u
∂fd1
∂x
+ v
∂fd1
∂y
2
B. Fuchs: Density waves in the shearing sheet
∂Φd0 + Φh0
∂x
∂Φd1 + Φh1
∂x
−
−
∂fd1
∂u −
∂fd0
∂u −
∂Φd0 + Φh0
∂y
∂Φd1 + Φh1
∂y
∂fd1
∂v
∂fd0
∂v
= 0 ,
(1)
where (x, y) denote spatial coordinates with y pointing in
the direction of galactic rotation and (u, v) are the cor-
responding velocity components, respectively. Equation
(1) has been derived from the general 6 -- dimensional
Boltzmann equation assuming delta -- function like depen-
dencies of the distribution function on the vertical z coor-
dinate and the vertical w velocity component, respectively,
and integrating the Boltzmann equation with respect to
them. A perturbation Ansatz of the form
fd = fd0 + fd1 , Φd = Φd0 + Φd1 , Φh = Φh0 + Φh1
(2)
is chosen for the distribution function of the disk stars
and the gravitational potentials of the disk and the halo,
respectively, and the Boltzmann equation (1) has been
linearized accordingly.
Similarly the linearized Boltzmann equation for the
halo particles can be written as
+ u
∂fh1
∂fh1
∂x
∂t
∂Φd0 + Φh0
−
−
−
∂x
∂Φd0 + Φh0
∂z
∂Φh1 + Φd1
∂y
+ v
∂fh1
∂y
+ w
∂fh1
∂z
∂fh1
∂u −
∂fh1
∂w −
∂fh0
∂v −
∂Φd0 + Φh0
∂y
∂Φh1 + Φd1
∂x
∂Φh1 + Φd1
∂z
∂fh1
∂v
∂fh0
∂u
∂fh0
∂w
= 0 .
(3)
The choice of the dark halo model is lead by the following
considerations. One of the deeper reasons for the success of
the infinite shearing sheet model to describe spiral density
waves realistically is the rapid convergence of the Poisson
integral in self -- gravitating disks (Julian & Toomre 1966).
Consider, for example, the potential of a sinusoidal density
perturbation
where κ denotes the epicyclic frequency of the stellar or-
bits and Σd is the surface density of the sheet. In the
solar neighbourhood in the Milky Way, for instance, the
critical wave length is λcrit = 5 kpc. Thus it is reason-
able to neglect over such length scales, like in the shearing
sheet model, the curvature of the mean circular orbits of
the stars around the galactic center or the gradient of the
surface density. The curvature of the stellar orbits due
to the epicyclic motions of the stars, on the other hand,
cannot be neglected and is indeed not neglected in the
shearing sheet model. The radial size of an epicycle is ap-
proximately given by σu/κ, where σu denotes the radial
velocity dispersion of the stars, and the ratio of epicycle
size and critical wave length is given by
σu
κλcrit
= 0.085Q
(7)
in terms of the Toomre stability parameter Q (Toomre
1964) which is typically of the order of 1 to 2. Concurrent
to these approximations I assume a dark halo which is
homogeneous in its unperturbed state. Accordingly the
curvature of the unperturbed orbits of the halo parti-
cles is neglected on the scales considered here and the
particles are assumed to be on straight -- line orbits. The
equations of motion of the halo particles are the charac-
teristics of the Boltzmann equation. In a homogeneous
halo x = ∇(Φd0 + Φh0) = 0, and in accordance with
this assumption I neglect the force terms ∇Φd0 and ∇Φh0
in the Boltzmann equation (3). This simplifies its solu-
tion considerably. The disadvantage of such a model is
that there are no higher -- order resonances of the orbits of
the halo particles with the density waves as described by
Weinberg (1985) or observed in the high -- resolution simu-
lations by Athanassoula (2002, 2003). However their effect
was shown to be much less important than the main reso-
nances of the particles with the density waves, which are
properly described in the present model.
Φ(x, y) = −GZ ∞
−∞
dx′Z ∞
−∞
dy′
Σ10 sin (kx′)
p(x − x′)2 + (y − y′)2
where G denotes the constant of gravitation. At x = 0
Φ = −4GΣ10 sin (kx) lim
xL→∞
2πGΣ10 sin (kx)
−
k
Si(kxL)
k
=
.
(5)
The sine integral in equation (5) converges so rapidly that
it reaches at kxL = π
2 already 87% of its asymptotic value.
Thus the "effective range" of gravity is about only a quar-
ter of a wave length. The shearing sheet models effectively
patches of galactic disks of such size. The wave lengths of
density waves are of the order of the critical wave length
(Julian & Toomre 1966)
λcrit =
4π2GΣd
κ2
,
(6)
3. Halo dynamics
, (4)
The Boltzmann equation (3) can be viewed as a linear
partial differential equation for the perturbation of the
distribution function of the halo particles, fh1, with inho-
mogeneities depending on the perturbations of the grav-
itational potentials of the disk and the halo, ∇Φd1 and
∇Φh1, respectively. Thus the equation can be solved for
the disk and halo inhomogeneities separately and the so-
lutions combined afterwards by superposition.
3.1. Jeans instability of the halo
I consider first that part of the Boltzmann equation (3)
+ u
∂fh1
∂t
∂Φh1
∂x
−
∂fh1
∂x
∂fh0
∂u −
+ v
∂fh1
∂y
+ w
∂Φh1
∂y
∂fh0
∂v −
∂fh1
∂z
∂Φh1
∂z
(8)
∂fh0
∂w
= 0 .
B. Fuchs: Density waves in the shearing sheet
3
Equation (8) is Fourier transformed with respect to time
and spatial coordinates and I consider in the following
Fourier terms
fhk,ω, Φhk,ω × ei(ωt+(k,x))
with frequency ω and wave vector k. If orthogonal spatial
coordinates ξ, η, and ζ with ξ parallel to the wave vector
k are introduced, equation (8) takes the form
(9)
∂fh0
∂υ
(10)
= 0 ,
iωfhk + υikfhk − ikΦhk
where k = k and υ denotes the velocity component par-
allel to the ξ-axis. Equation (10) can be immediately in-
tegrated with respect to the velocity components perpen-
dicular to k. Adopting for the background distribution
function a Gaussian distribution, fh0 = ρb√2πσh
,
the solution of equation (10) is given by
exp− υ2
2σ2
h
fhk = −Φhk
υ
σ2
h
k
ω + kυ
ρb√2πσh
exp−
υ2
2σ2
h
.
The corresponding Fourier coefficient of the density per-
turbation of the halo particles is
ρhk = −Φhk
ρb√2πσ3
h Z ∞
−∞
dυ
υ
ω
k + υ
exp−
υ2
2σ2
h
.
(12)
The perturbations of the halo are supposed to be self --
gravitating and the gravitational potential has to solve
the Poisson equation
∂2
∂y2 +
∂2
∂2
∂z2}Φhkei(k,x) =
{
∂x2 +
−k2Φhkei(k,x) = 4πGρhkei(k,x) .
Combining equations (12) and (13) leads to the dispersion
relation
k2 =
4πGρb√2πσ3
h Z ∞
−∞
dυ
υ
ω
k + υ
exp−
υ2
2σ2
h
,
(14)
which is well known from plasma physics (cf. Kegel 1998).
Indeed, the dispersion relation (14) describes simply the
Jeans collapse of the halo. From the imaginary part of
equation (14) one concludes ℜ(ω) = 0 and using formula
3.466 of Gradshteyn & Ryzhik (2000) one obtains from its
real part the dispersion relation
k2 =
4πGρb
σ3
h
{
σh
2 −r π
2 ℑω
k e
(ℑω)2
2k2 σ2
h erfc(cid:18) ℑω
√2kσh(cid:19)} , (15)
larger than the Jeans length pπσ2
where erfc denotes the complementary error function. As
is well known perturbations will grow on length scales
h/Gρb. However, dark
halos are thought to be dynamically hot systems and
their Jeans lengths will be of the order of the size of the
halos themselves. Thus this part of the solution of the
Boltzmann equation (3) is uninteresting in the present
context and will be not considered in the following.
3.2. Response of the halo to a density wave in the disk
I concentrate now on the remaining part of the Boltzmann
equation (3),
+ u
∂fh1
∂t
∂Φd1
∂x
−
∂fh1
∂x
∂fh0
∂u −
+ v
∂fh1
∂y
+ w
∂Φd1
∂y
∂fh0
∂v −
∂fh1
∂z
∂Φd1
∂z
(16)
∂fh0
∂w
= 0 ,
which describes the halo response to a perturbation in the
disk. If the gravitational potential perturbation of the disk
is Fourier expanded the Fourier terms have the form (cf.
equation 33 of Fuchs 2001)
Φdk
ei(kxx+kyy)−kz
(17)
with k = k =qk2
y. This can be converted to the
form as in equation (9) by introducing Fourier coefficients
x + k2
Φdk =
(11)
1
2π Z ∞
−∞
dzΦdk
e−ikzz−kz =
1
π
k2
k
+ k2
z
Φdk
.(18)
Note that the coordinate y, which is defined in the refer-
ence system of the disk, is related to the y coordinate in
the reference system of the halo due to the motion of the
center of the shearing sheet as
y → y − r0Ω0t .
Fourier transforming the distribution function of the halo
particles as fh1 =R d3kfhk exp i(k, x) and changing again
to ξ, η, ζ coordinates and the corresponding velocity com-
ponents the Boltzmann equation (16) takes the form
(19)
(13)
∂fhk
∂t
+ υikfhk + ikΦdk
υ
σ2
h
fh0ei(ω−kyr0Ω0)t = 0 ,
(20)
+ k2
where k = qk2
z . Equation (20) has been integrated
with respect to the velocity components perpendicular to
the ξ-axis. Assuming again a Gaussian velocity distribu-
tion for the halo particles it can be solved straightforward
as
ei(ω−kyr0Ω0)t
υ
σ2
h
ρb√2πσh
υ2
2σ2
h
. (21)
exp−
ω − kyr0Ω0 + kυ
fhk = −kΦdk
The corresponding Fourier term of the density perturba-
tion of the halo particles is given by
υe− υ2
k − kyr0Ω0
ρhk = −Φdk
h Z ∞
ρb√2πσ3
k + υ
(22)
−∞
2σ2
h
dυ
ω
,
where the time -- dependent term has been split off and is
written no longer explicitely. Next the gravitational poten-
tial associated with this density distribution is calculated
from the Poisson equation,
− k2Φhk = 4πGρhk ,
resulting in
(23)
Φhk = Φdk
4πGρb√2πσ3
h
1
k2 Z ∞
−∞
dυ
2σ2
h
υe− υ2
k − kyr0Ω0
ω
k + υ
.
(24)
4
B. Fuchs: Density waves in the shearing sheet
The integral on the rhs of equation (24) is not trivial and
has to be evaluated as a Cauchy principal value and the
pole contribution at υ = − ω
, where the halo
particles travel at velocities in resonance with the Doppler-
shifted phase velocity of the wave. Following the method of
Kegel (1998) I obtain for the principal value of the integral
k + kyr0Ω0
k
Φnr
hk = Φdk
×erf(cid:18)i
1
σ2
h
4πGρb
k2{1 + i√π
√2kσh (cid:19) exp−
kyr0Ω0 − ω
kyr0Ω0 − ω
√2kσh
(kyr0Ω0 − ω)2
2k2σ2
h
(25)
} ,
where the error function with imaginary argument is re-
lated to Dawson's integral. Since the gravitational forces
in equation (1) have to be taken at the midplane z = 0,
it is necessary to convert Φhk from the representation in
k space to a mixed representation in (k, z) space leading
to
dkz
Φnr
hk
(z = 0) =Z ∞
−∞
kyr0Ω0 − ω
√2kσh
(kyr0Ω0 − ω)2
×{1 + i√π
× exp−
2k2σ2
h
4Gρb
(k2
k
z )2 Φdk
σ2
+ k2
h
√2kσh (cid:19)
erf(cid:18)i
kyr0Ω0 − ω
} .
(26)
The pole contribution to the integral in equation (24) has
to be calculated according to Landau's rule resulting in
Φres
hk = Φdk
4πGρb
σ2
h
1
k2 i√π
kyr0Ω0 − ω
√2kσh
× exp−
(kyr0Ω0 − ω)2
2k2σ2
h
and in the mixed representation
Φres
hk
dkz
(z = 0) = −Z ∞
k
+ k2
−∞
(kyr0Ω0 − ω)2
ω − kyr0Ω0
√2kσh
exp−
2k2σ2
h
(k2
×
z )2 Φdk
.
The final result can be formally written as
Φhk
(z = 0) = Υ(ω − kyr0Ω0, k)Φdk
,
(27)
i
√π
4πGρb
σ2
h
(28)
(29)
where the real and imaginary parts of Υ are defined by
equations (26) and (28), respectively. Thus for any given
frequency there is a contribution both from the non --
resonant and the resonant halo particles. I find that in
dimensionless form
Fig. 1. The function Υ(ω − kyr0Ω0, k). The real part is
shown in the upper panel, while the imaginary part is
shown in the lower panel. The Doppler shifted frequency
is given in terms of the epicyclic frequency and the wave
number k in units of the critical wave number. Contours
are given at levels of 10−2Gρb/κ2. Negative values are
indicated as dotted lines. A Q parameter value of Q = 1.4
and a ratio of disk to halo velocity dispersions of σd : σh =
1 : 5 are assumed.
=
+ k2
(ω − kyr0Ω0)2
z )σ2
2(k2
h
σh(cid:19)2 1
1.748(cid:18)σd
Q2
k2
crit
+ k2
z
k2
(ω − kyr0Ω0)2
κ2
,
where σd denotes the velocity dispersion of the disk stars
and kcrit = 2π/λcrit, the critical wave number of the disk.
The function Υ is illustrated in Fig. 1 for parameters typi-
cal for the solar neighbourhood in the Milky Way. The ve-
locity dispersion of the halo particles has been estimated
as σh = r0Ω0/√2 like in an isothermal sphere.
(30)
4. Disk dynamics
The halo response (29) to the perturbation in the disk has
to be inserted into equation (1). Solving the Boltzmann
equation (1) is greatly facilitated by the fact that its form
is identical to the case of an isolated shearing sheet with
the replacement
Φdk → (1 + Υ)Φdk
(31)
B. Fuchs: Density waves in the shearing sheet
5
and I can use directly the results of Fuchs (2001) even
though the Boltzmann equation is treated there using ac-
tion and angle variables instead of the Cartesian coordi-
nates as in equation (1). In particular the factor 1 + Υ
is carried straightforward through to the fundamental
Volterra integral equation (equation 68 of Fuchs 2001)
x
y,ω (32)
−∞
+ rk′,ω ,
dkxK (kx, k′x) (1 + Υ(kx, k′y, ω))Φkx,k′
Φk′,ω = Z k′
where the kernel K is given by equation (67) of Fuchs
(2001). rk′ describes an inhomogeneity of equation (32)
related to an initial non -- equilibrium state of the shear-
ing sheet. Equation (32) is separating in the circumferen-
tial wave number k′y. In equation (32) the wave numbers
are expressed in units of the critical wave number kcrit
(cf. equation 30). This implies that the Volterra equation
describing a shearing sheet embedded in a rigid halo po-
tential is formally the same as that of an isolated shear-
ing sheet, because in this case Υ = 0 and the halo mass
affects only the numerical values of the critical wave num-
ber kcrit and the stability parameter Q. It is advanta-
geous to consider equation (32) transformed back from
frequency to time domain. Splitting off the ω -- dependent
term exp iω kx−k′
from the kernel and making use of the
2Ak′
y
convolution theorem of the Fourier transform of products
of two functions leads to
x
x
−∞
Φk′,t = Z k′
× {Z ∞
+ Z ∞
0
0
+ rk′,t ,
kx − k′x
2Ak′y (cid:19)
y,t′ δ(cid:18)t − t′ +
y,t′F Υ(kx, k′y, ω)e
iω
kx−k′
x
2Ak′
(33)
y !t−t′
}
dkx K (kx, k′x)
dt′Φkx,k′
dt′Φkx,k′
where the operator F denotes the Fourier transform from
ω to time domain. In equation (33) I have assumed an
initial perturbation of the disk at time t = 0 so that
Φkx,k′
y,t′<0 = 0. The Fourier transform F can be calcu-
lated analytically using formulae 6.3171 and 3.952, and
the integrals with respect to kz using formula 3.466 of
Gradshteyn & Ryzhik (2000) leading to
F Υ(kx, k′y, ω)e
4π2Gρb
1
σ2
h
x + k′y
k2
=
iω
kx −k′
x
2Ak′
y !t−t′
2 δ(cid:18)t − t′ +
kx − k′x
2Ak′y (cid:19)
kx − k′x
2Ak′y (cid:19)]
kx − k′x
2Ak′y (cid:12)(cid:12)(cid:12)}
kx − k′x
+4π2Gρb exp [ik′yr0Ω0(cid:18)t − t′ +
×{(cid:18)t − t′ +
2Ak′y (cid:19) +(cid:12)(cid:12)(cid:12)
t − t′ +
dω cos (bω)aωerf(iaω)e−a2ω2
1 I use the identity R ∞
∂bR ∞
dω sin (bω)erf(iaω)e−a2ω2
with formula 6.317.
0
0
a ∂
(34)
=
kx − k′x
x + k′y
t − t′ +
2Ak′y (cid:12)(cid:12)(cid:12)(cid:19) ,
×erfc(cid:18) σh√2qk2
absolute value (cid:12)(cid:12)(cid:12)
2(cid:12)(cid:12)(cid:12)
y (cid:12)(cid:12)(cid:12)
t − t′ + kx−k′
where the terms depending on the delta function and the
in the curly bracket repre-
sent the contribution from the non-resonant halo particles,
while the remaining term gives the contribution by the res-
onant halo particles. If this is inserted into equation (33)
it takes the form
x
2Ak′
x
Φk′,t =Z k′
+Z t+
−∞
kx−k′
x
2Ak′
y
0
+rk′,t .
dkx K (kx, k′x) {Φ
kx,k′
y,t+
kx−k′
x
2Ak′
y
(35)
dt′Φkx,k′
y,t′F Υ(kx, k′y, ω)e
iω
kx −k′
x
2Ak′
y !t−t′
}
Equation (35) can be integrated numerically with very
modest numerical effort. In Fig. 2 I illustrate the response
of the shearing sheet now embedded in a live halo to
an initial sinusoidal perturbation of unit amplitude. For
this purpose I use the inhomogeneity term of the Volterra
equation
x
rk′,ω =Z k′
−∞
dkxL (kx, k′x) fkx,k′in
y
(0)
(36)
derived in Fuchs (2001) with fkx,k′in
x ). The
response of the shearing sheet to this initial impulse is
a swing amplification event. The radial wave number kx
evolves as
(0) ∝ δ(kx − kin
y
kx = kin
x + 2Ak′y
int ,
(37)
in is constant,
while the circumferential wave number k′y
which means that the wave crests swing around from lead-
ing to trailing orientation during the amplification phase.
Around t = 6 the amplitudes become negative which in-
dicates that the swing amplified density wave is also os-
cillating. As can be seen from Fig. 2 comparing the evo-
lution of shearing sheets embedded either in a rigid halo
potential or a live halo this characteristic behaviour of the
density wave is not changed by the responsive halo, but
the maximum growth factor of the amplitude of the wave
is enhanced by a surprisingly large amount. In Fig. 2 the
evolution of a shearing sheet embedded in a static halo
potential has been calculated from equation (35) setting
F = 0. The maximum growth factor is identical to within
1% accuracy with the maximum growth factor calculated
by Fuchs (2001) in an alternative way. In order to test
the accuracy of the numerical solution of the full equation
(35) I have increased the velocity dispersion of the halo
particles to σd : σh = 1 : 13. The halo is then very stiff
and the enhancement of the maximum growth factor of
the amplitude of the wave by the responsive halo is only
5.3% so that the solution of equation (35) is very similar
to that for a density wave in a shearing sheet in a static
halo. If I use this, which is correctly calculated, in the
second term of equation (35) I confirm by this iterated so-
lution the maximum growth factor of the self -- consistent
6
B. Fuchs: Density waves in the shearing sheet
solution of equation (35) to an accuracy better than 1%.
The enhanced maximum growth factor of swing amplified
stance, by keeping in equation (34) only that part of the Υ
function related to the resonant halo particles and solving
equation (35) approximately by iterating the solution for
the isolated sheet,
x
Φk′,t ≈ Z k′
+ Z ∞
−∞
0
dkx K (kx, k′x){Φ
kx,k′
y,t+
kx−k′
x
2Ak′
y
dt′Φiso
kx,k′
y,t′F Υres(kx, k′y, ω)e
+ rk′,t .
(38)
iω
kx−k′
x
2Ak′
y !t−t′
}
The amplification of density waves depends critically
on the Toomre stability parameter Q. This is illustrated
in Fig. 3 where the response of the shearing sheet to the
same initial impulse as in the previous example is shown,
but assuming a stability parameter of Q = 2. As can be
seen in Fig. 3 there is neither effective amplification of
density waves in a shearing sheet in a rigid halo potential
or in a shearing sheet embedded in a live dark halo.
In Fig. 4 I illustrate the maximum growth factors as
function of the circumferential wave number ky for shear-
ing sheets embedded in a rigid halo potential or a live
dark halo, respectively. As can be seen from Fig. 4 in
both cases maximum amplification occurs for wave num-
bers around ky = 0.5 or in the notation of Toomre (1981)
X = k−1
y = 2. Thus the preferred circumferential wave
length of the density waves is 2λcrit.
Following Athanassoula et al. (1987) arguments of
swing amplification theory have been used to constrain
the mass -- to-light ratios of galactic disks. The expected
number of spiral arms is determined by how often the az-
imuthal wave length 2π/ky fits onto the circular annulus,
m =
2πr0
2π
ky
.
(39)
As shown in Fig. 4 density waves grow preferentially with
an azimuthal wave number around ky ≈ 0.5kcrit. This as-
sumes a disk with a flat rotation curve, A/Ω0 = 0.5, but
the preferred wave length varies with the slope of the ro-
tation curve measured by Oort's constant A (cf. Fuchs
2001). The observed number of spiral arms of a galaxy
taken together with the rotation curve allows then to de-
termine the surface density of the disk. Since the wave
number of maximum growth of the density wave ampli-
tudes is the same for shearing sheets embedded either in a
rigid halo potential or in a live dark halo, the presence of
a responsive dark halo does not influence the constraints
on the disk mass obtained this way.
Another constraint on the disk mass depends on the
implied value of the Toomre stability parameter Q (Fuchs
1999, 2003). Q must be larger than Q = 1 in order to avoid
Jeans like instabilities in the disks (Toomre 1964). A disk
violating this condition will evolve rapidly by fierce dy-
namical instabilities above the threshold of Q = 1 (Fuchs
& von Linden 1998). On the other hand, I have demon-
strated above that the Q value must be less than Q = 2 in
Fig. 2. Swing amplified density wave in the shearing sheet.
The upper diagram shows the evolution in a shearing
sheet embedded in a static halo potential triggered by
an impulse with unit amplitude and wave vector kin =
(−2, 0.5)kcrit. Time is given in units of 0.045 epicycle pe-
riods. The middle diagram shows the evolution of a shear-
ing sheet embedded in a live dark halo triggered by the
same impulse. The lower diagram shows the difference.
The model parameters are A/Ω0 = 0.5, Q = 1.4, σd : σh =
1 : 5, Gρb/κ2 = 0.01, and r0Ω0 : σd = 220 : 44.
density waves due to a responsive halo seems to be the
equivalent of the enhanced growth of bars of stellar disks
embedded in live dark halos seen in the numerical simula-
tions. However, the interaction of the shearing sheet and
the surrounding halo is not only mediated by the resonant
halo particles, but the non -- resonant halo particles play an
important role as well. This can be demonstrated, for in-
B. Fuchs: Density waves in the shearing sheet
7
Fig. 4. Maximum growth factors of swing amplified den-
sity waves in a shearing sheet embedded in a rigid halo
potential (squares) and in a shearing sheet embedded in a
live dark halo (triangles). The same model parameters as
above have been adopted.
where I assume a gravitational potential of the form
Φ(x, t) = ℜ{Φ(x)eiωt} ,
with complex frequency ω. Thus the overall acceleration
of the halo is given by
(41)
< v >= −Z d3xZ d3vfh∇Φ .
(42)
If both the gravitational potential and the distribution
function of the halo particles are Fourier expanded2,
< v > = −Z d3xZ d3vZ d3k ikΦk(t)ei(k,x)
× Z d3k′fhk′(t)ei(k′,x)
= (2π)3Z d3vZ d3k ikΦ−k(t)fhk(t) .
(43)
In order that the gravitational potential Φ be a real
quantity its Fourier coefficients must obey the relation
Φ−k(t) = Φ∗k(t), and for a potential of the form (41) one
can write
Φk(t) =
1
2{Φkeiωt + Φ∗
−ke−iω∗t} .
(44)
2 It can be shown (Binney & Tremaine 1987 and refer-
ences therein) that if the Fourier transformed potential of a
point mass is inserted into equation (43) this leads exactly to
Chandrasekhar's dynamical friction formula, although with a
Coulomb logarithm defined in a slightly different way.
Fig. 3. Same as in Fig. 2, but adopting Q = 2.
order that the disk can develop appreciable spiral struc-
ture. If the radial velocity dispersions of the stars are
known, the allowed range of Q sets constraints on the
disk mass, which are independent from the previous ones
(cf. equations 6 and 7). As shown in Fig. 3 this condition
is not changed by the presence of a live dark halo.
5. Momentum transfer to the halo
The analogue of the angular momentum transfer from bars
to dark halos is in the shearing sheet the transfer of linear
momentum to the surrounding halo particles. In order to
describe this I adapt the discussion of Dekker (1976) to the
shearing sheet model. The acceleration of a halo particle
is given by
v = −∇Φ ,
(40)
8
B. Fuchs: Density waves in the shearing sheet
Again spatial coordinates ξ, η, and ζ are introduced with
the ξ -- axis parallel to the wave vector k and using equation
(11) one obtains from equation (43)
(45)
(46)
(47)
(48)
6. Summary and conclusions
If a self -- gravitating shearing sheet is embedded in a live
dark halo, the halo particles respond unexpectedly strong
to density waves in the sheet. The interaction between the
density waves and the halo particles is mediated both by
halo particles on orbits in resonance with the waves and
on non -- resonant orbits. If the embedded shearing sheet
is initially perturbed by a sinusoidal wave, a swing am-
plified density wave develops in the disk, which is of the
same type as in an isolated sheet or a sheet in a static halo
potential, but with an amplitude enhanced by a surpris-
ingly large amount. This appears to be the equivalent of
the enhanced bar growth in stellar disks embedded in live
dark halos instead of static halo potentials seen in numer-
ical simulations. There is transfer of linear momentum of
density waves in the shearing sheet to the halo particles.
This is mediated, however, entirely by halo particles on
orbits in resonance with the waves similar to the torque
exerted by bars on the surrounding halo.
Acknowledgements. I thank E. Athanassoula for stimulating
discussions and the anonymous referee for helpful comments.
References
Athanssoula , E., 2002, ApJ, 569, L83
Athanssoula , E., 2003, MNRAS, 341, 1179
Athanssoula , E., Bosma, A., Papaioannou, S., 1987, A&A,
179, 23
Binney, J., Tremaine, S., 1987, Galactic Dynamics, Princeton
Univ. Press, Princeton, p.487
Debattista, V.P., Sellwood, J.A., 2000, ApJ, 543, 704
Dekker, E., 1976, Phys. Reports, 24, 315
Fuchs, B., 1999, Astron. Soc. Pac. Conf. Ser., 182, 365
Fuchs, B., 2001, A&A, 368, 107
Fuchs, B., 2003, in: N.J.C Spooner, V. Kudryavtsev (eds.)
Proc. Fourth Int. Workshop on the Identification of Dark
Matter, World Scientific, Singapore, p. 72
Fuchs, B., 2004, A&A, submitted
Fuchs, B., von Linden, S., 1998, MNRAS, 294, 513
Goldreich, P., Lynden -- Bell, D., 1965, MNRAS, 130, 125
Gradshteyn, I.S., Ryzhik, I.M., 2000, Table of Integrals, Series,
and Products (6th Ed.), Academic Press, New York
Julian, W.H., Toomre, A., 1966, ApJ, 146, 810
Kegel, W.H., 1998, Plasmaphysik, Springer, Berlin
Lynden -- Bell, D., Kalnajs, A.J., 1972, MNRAS, 157, 1
Ostriker, J.P., Peebles, P.J.E., 1973, ApJ, 186, 467
Sellwood, J.A., 2003, ApJ, 587, 638
Stix, T.H., 1962, The theory of plasma waves, McGraw-Hill,
New York
Toomre, A., 1964, ApJ, 139, 1217
Toomre, A., 1977, ARAA, 15, 437
Toomre, A., 1981, in: S.M. Fall, D. Lynden -- Bell (eds.) The
Structure and Evolution of Normal Galaxies, Cambridge
Univ. Press, Cambridge, p. 111
Weinberg, M.D., 1985, MNRAS, 213, 451
Weinberg, M.D., Katz, N., 2002, ApJ, 580, 627
< υ > = i(2π)3Z d3kZ ∞
−∞
dυ
ρb√2πσ3
h
e− υ2
2σ2
h
k
×
4{Φ−keiωt + Φ∗ke−iω∗t}
× {−kΦk
υeiωt
ω + kυ − kΦ∗
−k
υe−iω∗t
−ω∗ + kυ} .
In equation (45) I have integrated already over the velocity
components perpendicular to the ξ-axis. Two of the four
terms of the integrand cancel out leading to
< υ > = −2π3iZ d3kZ ∞
× k2e−2ℑωtΦk2{
ρb√2πσ3
−∞
ω + kυ −
dυ
υ
h
2σ2
h
e− υ2
υ
ω∗ + kυ} .
In the limit of −ℑω → 0
ω + kυ −
υ
υ
ω∗ + kυ → −i2πυδ(ℜω + kυ) ,
and I find
< υ > = −4π4Z d3k
ρb√2πσ3
(ω − kyr0Ω0)2
2k2σ2
h
,
× exp−
h Φk2(ω − kyr0Ω0)
where ω denotes now the frequency of the wave in the
reference frame of the shearing sheet as in the previous
sections. Note that according to the notation used here
(ω − kyr0Ω0) < 0 for waves traveling along the y-axis
in the forward direction, so that < υ > > 0 for such
waves. From the form of the Fourier coefficients (18) it
becomes immediately clear that only planar momentum
is transferred to the halo, < vz >= 0, as expected from
symmetry reasons. < vx > and < vy > depend on the
exact form of the power spectrum Φk2. Equation (47)
shows that linear momentum transfer to the halo is medi-
ated entirely by halo particles on orbits resonant with the
wave. This is exactly the same as for angular momentum
transfer from bars to halos (Weinberg 1985, Weinberg &
Katz 2002, Athanassoula 2003, Sellwood 2003). It has also
been known for a long time for the angular momentum
exchange between spiral density waves and galactic disks
(Lynden-Bell & Kalnajs 1972) or for momentum transfer
from longitudinal plasma waves (Stix 1962). It is interest-
ing that the exchanged linear momentum scales as ρb/σ3
h
in equation (48). This implies that at given mass a dy-
namical hot halo absorbs less momentum than a cooler
one. On the other hand, the shearing sheet would loose
less momentum. The same effect has been described by
Athanassoula (2003) for the angular momentum transfer
from bars to halos.
|
astro-ph/0702268 | 1 | 0702 | 2007-02-09T18:30:27 | Pre-main sequence spectroscopic binaries suitable for VLTI observations | [
"astro-ph"
] | A severe problem of the research in star-formation is that the masses of young stars are almost always estimated only from evolutionary tracks. Since the tracks published by different groups differ, it is often only possible to give a rough estimate of the masses of young stars. It is thus crucial to test and calibrate the tracks. Up to now, only a few tests of the tracks could be carried out. However, with the VLTI it is now possible to set constrains on the tracks by determining the masses of many young binary stars precisely. In order to use the VLTI efficiently, a first step is to find suitable targets, which is the purpose of this work. Given the distance of nearby star-forming regions, suitable VLTI targets are binaries with orbital periods between at least 50 days, and few years. Although a number of surveys for detecting spectroscopic binaries have been carried out, most of the binaries found so far have periods which are too short. We thus surveyed the Chamaeleon, Corona Australis, Lupus, Sco-Cen, rho Ophiuci star-forming regions in order to search for spectroscopic binaries with periods longer than 50 days, which are suitable for the VLTI observations. As a result of the 8 years campaign we discovered 8 binaries with orbital periods longer than 50 days. Amongst the newly discovered long period binaries is CS Cha, which is one of the few classical T Tauri stars with a circumbinary disk. The survey is limited to objects with masses higher than 0.1 to 0.2 Modot for periods between 1 and 8 years. We find that the frequency of binaries with orbital periods < 3000 days is of 20+/-5 percent. The frequency of long and short period pre-main sequence spectroscopic binaries is about the same as for stars in the solar neighbourhood. In total 14 young binaries are now known which are suitable for mass determination with the VLTI. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. ttauribinaries
October 11, 2018
c(cid:13) ESO 2018
Pre-main sequence spectroscopic binaries suitable for VLTI
observations ⋆
E.W. Guenther1, M Esposito1
,
2, R. Mundt3, E. Covino4, J.M. Alcal´a4, F. Cusano1, and B. Stecklum1
1 Thuringer Landessternwarte Tautenburg, Sternwarte 5, D-07778 Tautenburg, Germany
2 Dipartimento di Fisica "E.R. Caianiello", Universit`a di Salerno, via S. Allende, 84081 Baronissi (Salerno), Italy
3 Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany
4 INAF-Osservatorio Astronomico di Capodimonte via Moiariello 16, I-80131 Napoli, Italy
Received 20.05.2006; accepted 6.2.2007
ABSTRACT
Context. A severe problem of the research in star-formation is that the masses of young stars are almost always estimated only
from evolutionary tracks. Since the tracks published by different groups differ, it is often only possible to give a rough estimate
of the masses of young stars. It is thus crucial to test and calibrate the tracks. Up to now, only a few tests of the tracks could
be carried out. However, with the VLTI it is now possible to set constrains on the tracks by determining the masses of many
young binary stars precisely.
Aims. In order to use the VLTI efficiently, a first step is to find suitable targets, which is the purpose of this work. Given the
distance of nearby star-forming regions, suitable VLTI targets are binaries with orbital periods between at least 50 days, and
few years. Although a number of surveys for detecting spectroscopic binaries have been carried out, most of the binaries found
so far have periods which are too short.
Methods. We thus surveyed the Chamaeleon, Corona Australis, Lupus, Sco-Cen, ρ Ophiuci star-forming regions in order to
search for spectroscopic binaries with periods longer than 50 days, which are suitable for the VLTI observations.
Results. As a result of the 8 years campaign we discovered 8 binaries with orbital periods longer than 50 days. Amongst the
newly discovered long period binaries is CS Cha, which is one of the few classical T Tauri stars with a circumbinary disk. The
survey is limited to objects with masses higher than 0.1 to 0.2 M⊙ for periods between 1 and 8 years.
Conclusions. We find that the frequency of binaries with orbital periods ≤ 3000 days is of 20 ± 5%. The frequency of long and
short period pre-main sequence spectroscopic binaries is about the same as for stars in the solar neighbourhood. In total 14
young binaries are now known which are suitable for mass determination with the VLTI.
7
0
0
2
b
e
F
9
1
v
8
6
2
2
0
7
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Key words. binaries, radial velocity monitoring
1. Introduction
The most fundamental parameter of a star is its mass,
which determines almost everything about its birth, life,
and death. The masses of low mass pre-main sequence
(pms) stars are often derived by comparing the location
of the star in the Hertzsprung-Russell diagram with the-
oretically calculated evolutionary models. Unfortunately,
the evolutionary models published by various authors dif-
fer considerably due to the differences in the input physics,
such as the treatment of convection, magnetic fields, chem-
offprint
Send
[email protected]
requests
to: Eike Guenther,
e-mail:
⋆ based on observations obtained at the European Southern
Observatory at La Silla, Chile in program 62.I-0418(A);
63.I-0096(A); 64.I-0294(A); 65.I-0012(A); 67.C-0155(A); 68.C-
0292(A); 68.C-0561(A); 69.C-0207(A); 70.C-0163(A); 073.C-
0355(A); 074.A-9018(A); 075.C-0399(A-F)
ical abundance of the star etc. (Palla & Stahler, 1992;
D'Antona & Mazzitelli 1994; Swenson et al. 1994; Burrows
et al. 1997; Forestini 1994; Siess et al. 1997; Baraffe et al.
1998; Palla & Stahler 1999; Tout et al. 1999; Chabrier et
al. 2000, D'Antona et al. 2000; Wuchterl & Tscharnuter,
2003; Montalb´an et al. 2004). Testing and calibrating evo-
lutionary models of pms-stars are thus of key importance
for the understanding of young stars, the determination
of the initial mass function, and studies of the galactic
star-formation history in general. The lack of knowledge,
which of the models to choose, is thus a severe prob-
lem for the whole field of star-formation. Classical T Tauri
stars (henceforth called CTTSs) are young, low mass, op-
tically visible pms-emission line stars with an accretion
disk. Weak-line T Tauri stars (WTTSs) are similar to the
classical ones but have much smaller accretion rates, and
less massive disks, if any.
2
Guenther et al.: PMS spectroscopic binaries
An ideal test of the tracks would be to compare the
true masses of CTTSs and WTTSs with the masses de-
rived from the models. This is, for example, possible for
eclipsing binary stars. Unfortunately, by now there are
only three eclipsing young binaries known after we elimi-
nated RXJ 1608.6-3922 by showing that it is not an eclips-
ing binary but a spotted single star (Joergens et al. 2001).
Casey et al. (1998) derived the masses of the eclipsing
binary TY CrA to 3.16 ± 0.02 M⊙ and 1.64 ± 0.01 M⊙
and compared these values with three sets of models.
Evolutionary models do not differ too much in the mass-
range between 1.5 and 3.0 M⊙, and thus the authors find
that all three sets of models are reasonably consistent
with the observations. Since many pms-stars have masses
smaller than 1.5 M⊙ it is important to have direct mass
determinations also in the low-mass domain. It is thus
better to focus on stars with masses lower than 1.5 M⊙.
Covino et al. (2004) analysed the eclipsing binary RXJ
0529.4+0041 which consists of a 1.25 M⊙ primary star
and a 0.91 M⊙ companion star and compared these val-
ues with three sets of evolutionary models. They find that
the models published by Baraffe et al. (1998) and Swenson
et al. (1994) are in reasonable agreement with the obser-
vations but the ones published by D'Antona & Mazzitelli
(1994) are not. Stassun et al. (2004) analysed the eclips-
ing binary V1174 Ori which consists of a 1.01 ± 0.02 M⊙
primary and a 0.731 ± 0.008 M⊙ secondary. They find that
models by Montalb´an et al. (2004) seem to agree with the
observations but those published by Siess et al. (1997) are
in conflict with them.
By measuring the orbital motion of the molecular gas
in the disk, Simon et al. (2000) concluded that models
of the pms-evolution with lower Teff values are in better
agreement with the observations (see, e.g., Baraffe et al.
1998 and Palla & Stahler 1999). Another possibility to
measure masses directly is to combine astrometric data
obtained with the HST Fine Guidance Sensors with ra-
dial velocity (RV) measurements. In this way Steffen et al.
(2001) derived masses of 1.5 ± 0.2 M⊙ and 0.81 ± 0.09 M⊙
for the binary system NTT 045251+3016. They find that
the Baraffe et al. (1998) models with a mixing-length pa-
rameter of α = 1.0 are closest to the measured primary
mass. The models published by D'Antona & Mazzitelli
(1994) are clearly inconsistent with the observations. The
first determination of the masses of the components of
a young binary star using an interferometer was recently
carried out by Boden et al.(2005). They used the Keck in-
terferometer and determined the masses of the two compo-
nents of HD 98800 B to 0.70±0.06 M⊙ and 0.58±0.05 M⊙.
With an age of 8 to 20 Myr it is however much older than a
typical T Tauri star. Nevertheless, the object is still young
enough to be used for testing the tracks. However, a rea-
sonable agreement with the models published by Baraffe
(1998) could only be achieved by assuming a rather low
metallicity of the star.
While all these data allow a very first test of evolu-
tionary models, more mass determinations are definitely
needed, especially of stars of lower masses. Shown in Fig. 1
is log(L/L⊙) versus log(M/M⊙) for all young stars of
which the masses have so far been derived. However, we
should keep in mind that the position of young stars in
the Hertzsprung-Russell diagram depends on the history
of the accretion rate. Thus, short-period, eclipsing bina-
ries might not be ideal for testing the tracks, because the
components might have had recent mass-exchange. An ex-
ample where this could have been the case is RXJ1603.8-
3938. In this system both components have almost identi-
cal masses but one is a factor two brighter than the other
(Guenther et al. 2001).
Fig. 1. The figure summarises the current knowledge of
empirical mass determination of young stars. The masses
of the eclipsing objects in Orion and CrA were taken from
Covino et al. (2004), Stassun et al. (2004), and Casey et
al. (1998). Alencar et al. (2003) derived the masses of
the binary in upper Sco. Using the HST Fine Guidance
Sensor Steffen et al. (2001) determined the masses of NTT
045251+3016 in CrA. Also shown are the masses derived
by Simon et al. (2000) by measuring the orbital motion of
gas in the disk.
In order to determine the masses of many more pms-
stars over a large mass range we have carried out a survey
for long-period binary stars. For single-line spectroscopic
binaries (SB1) only the mass-function f (m) = m3
2 sin3 i
(m1+m2)2
can be derived using the RV-method. If f (m) is combined
with the relative astrometric orbit of the two stars (posi-
tion angle and relative distance of the stars) and if also its
parallax is known, the masses of the two stars can be de-
termined. Using this method, we have derived the masses
of the two components of the SB1-system EK Dra (Konig
et al. 2005). This system has, however, an age of about 125
Myr and thus, it is not really young. For double-line spec-
troscopic binaries (SB2), it is possible to measure m1 sin3 i
and m2 sin3 i. The individual masses of the stars can then
be derived if the RV-data is combined with the relative
astrometric orbit, or if the system is eclipsing.
Unfortunately, the distance to the nearest star-forming
regions is so large, that binaries that could be resolved
Guenther et al.: PMS spectroscopic binaries
3
with AO-systems on 8 m-class telescopes have orbital pe-
riods of many decades, or even centuries. However, the
big leap forward in this work is the Very Large Telescope
Interferometer, which is now available. The VLTI instru-
ment AMBER (Astronomical Multi BEam combineR) of-
fers the unique opportunity to spatially resolve binaries
down to separations of only 4 mas in the K-band. Thus,
with AMBER it is possible to resolve binaries with separa-
tions down to 0.6 AU (corresponding to an orbital period
of ∼ 120 d for systems with m1 = m2 = M⊙) for distances
≤ 150 pc. Because AMBER works in the near infrared, it
is possible to detect a young 0.1 M⊙ star next to a young
1.0 M⊙ star due to the achievable brightness contrast ra-
tio.
In order to carry out such measurements with AMBER
a sufficient number of pms-binaries with suitable separa-
tions (i.e. sufficiently long periods) have to be identified
first. The aim of this work thus is the identification of pms-
binaries suitable for the VLTI observations. For the survey
we selected known pms-stars in southern star-forming re-
gions which have a distance ≤ 150 pc. Stars fainter than 10
mag in the K-band, visual pairs with a separation of ≤ 10
arcsec and stars with a v sin i ≥ 50 km s−1, are excluded.
Before we started our survey in 1998, only three
pms-spectroscopic binaries located in nearby star-forming
regions and with orbital periods longer than 50 days
were known (Mathieu 1994). Four additional spectroscopic
binary candidates in the Ophiuchus-Scorpius (Oph),
Chamaeleon, Lupus, and Corona Australis star form-
ing regions were identified by Melo et al. (2003). Thus,
while quite a number of authors have already surveyed
the nearby star-forming regions for spectroscopic binaries,
most of the objects found are binaries of short period and
unsuitable for VLTI observations. The main intention of
our survey is a search for long-period binaries, suitable for
VLTI-observations, not a comprehensive survey of spec-
troscopic binaries. In this paper we to present the out-
come of this extensive survey. The short-period systems,
which were also discovered in the course of the survey will
be subject of a forthcoming paper. In section 2, we de-
scribe the observations, in section 3, we give a full list
of all long period pms-binaries suitable for the VLTI ob-
servations together with the RV-values and the orbits of
the newly discovered binaries. In section 4, we discuss the
spectroscopic binary candidates. In the last section we dis-
cuss the results obtained.
2. The sample
For our survey we selected 122 late-type pms-stars in the
Chamaeleon (Cha), Corona Australis (CrA), ρ Ophiuchi
(Oph), Lupus (Lup), and Scorpius Centaurus (SC) star-
forming regions. We did not include the 12 SBs known at
the star of the survey (Mathieu et al. 1994; Melo 2003;
Guenther et al. 2001), and also did in general not include
visual binaries, unless the separation was ≥ 10 arcsec.
After the first observing run, we removed another 14 stars
from the list, because their v sin i turned out to be too
high, or because these stars are not young. Apart from
the known binaries, and stars with very large v sin im, we
observed essentially all stars with spectral types between
F6 and M3 in these regions. The sample that was finally
observed comprises 108 stars, of which 26 are CTTS, and
82 WTTS.
Determinations of the distance of the Chamaeleon as-
sociation give values of 160 ± 15 pc for ChaI, 178 ± 18 pc
for ChaII (Whittet et al. 1997), 171 ± 20 pc for ChaI
(Wichmann et al. 1998), and 168+14
−12 pc (Bertout et al.
1999). For the Corona Australis region Knude & Høg
(1998) derive a distance of 170 pc which is different from
the ∼ 130 pc found by other authors (Knacke et al. 1973;
Marraco & Rydgren 1981). For the ρ Ophiuchi region,
the distance of 160 pc derived by Knude & Høg (1998)
agrees well with the results obtained by other authors
(Whittet 1974; Chini 1981). Quite a number of authors
have determined the distance to the Lupus star-forming
region: Hughes et al. (1993) find 140 ± 20 pc, Knude &
Høg (1998) 100 pc, Nakajima et al. (2000) 150 pc, Sartori
et al. (2003) 147 pc, Franco (2002) 150 pc, de Zeeuw et
al. (1999) 142 ± 2 pc, and Teixeira et al. (2000) 85 pc but
note that 14 stars of this group have measured parallax-
distances, with an average of 138 pc.
The Upper Scorpius OB association has an age of 5-
6 Myr (Preibisch & Zinnecker 1999) and thus is slightly
older than the other regions that all have ages of 1-3 Myr.
The average distance of the Upper Scorpius OB associa-
tion is 145 ± 2 pc (de Zeeuw et al. 1999). In Tab. 1 we
also give the parameters of the binaries found in the TW
Hydra (TWA) and Tucana (Tuc) associations. Both re-
gions are older but also closer (45-60 pc) than the other
regions. Since low-mass stars in these regions are still pms,
these objects are interesting for VLTI observations. The
TW Hydra region has a age of 8-12 Myr, and the Tucana
association an age of about 30 Myr (Webb et al. 1999;
Torres et al. 2000; Torres et al. 2003; Weinberger et al.
2004).
3. Observations
All stars were observed with the ESO Echelle spec-
trograph FEROS (Fiber-fed Extended Range Optical
Spectrograph). FEROS was operated up to October 2002
(HJD 245 2550) at the ESO 1.5-m-telescope, and was then
moved to the MPG/ESO-2.20-m-telescope. The spectra
cover the wavelength region between about 3600 A and
9200 A with a resolving power of λ/∆λ = 48000. On the
1.5-m-telescope the entrance aperture of the fibre had a
projected diameter of 2.′′7 on the sky, while at MPG/ESO
2.2-m-telescope 2.′′0. As long as FEROS was at the ESO
1.5-m-telescope, exposure-times were set so that a S/N-
ratio of about 30 was achieved. The exposure-times were
thus between 5 and 45 minutes. Since the performance of
FEROS dramatically improved when FEROS was moved
to the MPG/ESO 2.2-m-telescope, the S/N-ratio went up
to typically 50, although we shortened the exposure times.
The standard FEROS data-reduction pipeline was used
4
Guenther et al.: PMS spectroscopic binaries
for bias subtraction, flat-fielding, scattered light removal,
Echelle order extraction, and wavelength calibration of the
spectra.
When measuring the radial velocity of stars, the ques-
tion arises whether it is better to use several templates
of different spectral types, or only one template. We tried
out both methods using HR3862 (G0), HR6748 (G5), HR
5777 (K1), HR5568 (K4), and HR6056 (M0.5) as tem-
plates. The advantage in using several templates is that
the match between the template and the star is better,
but the disadvantage is that the absolute RVs of the tem-
plates have to be known to a high accuracy, if the absolute
RVs of a stars are of interest. If the absolute RVs of the
templates are not well known to a high accuracy, the ab-
solute values of the RVs of the stars are lost. In our case
this would be a disadvantage, because the first step for
detecting pms-spectroscopic binaries is to determine their
absolute RV. A star with a RV that differs by more than
2 km s−1 from that of the star-forming region is likely to
be either not a member of that region, or an SB1.
The other alternative is to use just one template.
This approach is used in the HARPS search for extra-
solar planets of F,G, and K-stars, where an accuracy
of typically 1 m s−1 is achieved (Lovis et al. 2005). A
good choice in our case is the K1-star HR 5777, which is
bright (mv = 4.6), and its absolute RV is well determined
(+49.12±0.06 kms−1, Murdoch et al. 1993). If we use just
one template, the question is whether the mismatch of the
spectral type between the template and the star reduces
the accuracy of the RV-measurements. If this were the
case, the accuracy of the RV-measurements of the K-stars
would be higher than those of stars with other spectral
types. Fig. 2 shows the variance of the RV-measurements
versus the spectral type. There is no obvious trend of the
variance with the spectral type. This is possibly because
we have only very few M-stars in our survey, were the mis-
match with the template really matters. We thus use only
HR 5777 as template. In this way, it is possible to compare
the RVs of different stars, and it is possible to merge our
data with data taken with other instruments.
We split up the spectrum in six wavelength regions
which are practically free of stellar emission lines, and
virtually free of telluric lines. The wavelength regions are:
4000 to 4850 A, 4900 to 5850 A, 5900 to 6500 A, 6600 to
6860 A, 7400 to 7500 A, and 7700 to 8100 A. For each
of the wavelength-band we obtained the RV separately,
and then averaged these 6 values. The errors of the RV-
measurements of the T Tauri stars are determined from
the variance of the RV-values of the individual spectral
regions. In many stars, we could use only the first three
regions, because the S/N ratio of the other regions was too
low. In such cases we always used the same regions for the
same star. In order to account for any instrumental shift,
we measured the position of the telluric spectral lines in
the 6860 to 6930 A region by cross-correlating this part
of the stellar spectrum with one taken with the Fourier
Transform Spectrometer of the McMath-Pierce telescope
at Kitt Peak (Wallace et al. 1998; Brault 1978). We find
an instrumental shift of typically 0.5 km s−1 which differs
typically by less than 0.1 km s−1 from frame to frame.
Fig. 2. The figure shows the scatter of the measurements
of all single stars listed in Table 18, 19, 20 versus the spec-
tral type. Assuming that the scatter is the error of the
measurements, we conclude that the error dominated by
the v sin i of the stars, and there is no obvious trend with
the spectral type.
In order to investigate the accuracy of the RV-
measurements obtained, we took 31 spectra of HR 5777
spread out over the whole observing campaign. These
spectra were reduced in the same way as the spectra of
our targets. That is, we used one spectrum of HR 5777
as template (taken on the March 23, 1999), and cross-
correlated this spectrum with all other spectra taken of
HR 5777. The absolute RV of the template was deter-
mined by cross-correlating it with the FTS-spectrum of
the sun which was rebinned to the same resolution as the
FEROS spectra. (Wallace et al. 1998; Brault 1978). Fig. 3
shows a histogram of the RV-measurements of HR 5777.
The RV-values determined for HR 5777 are perfectly con-
sistent with the published RV of +49.12 ± 0.06 kms−1
(Murdoch et al. 1993). Fitting a Gaussian to this distri-
bution, we derive a FWHM of 0.30 kms−1. This implies
that the error is ±0.15 kms−1, which agrees well with the
error derived from the variance of the RV-values which is
±0.12 kms−1. We thus take ±0.15 kms−1 as the intrinsic
uncertainty of the measurements. The uncertainty of the
RV-measurements of the pre-main sequence stars as deter-
mined from the variance of the RV-measurements of the
individual spectral regions, are usually larger than those
of HR 5777, because of the smaller signal-to-noise and the
larger v sin i of the stars.
4. Results: Long-period spectroscopic binaries
The aim of this survey is to detect young binaries suit-
able for VLTI-observations, i.e. binaries with orbital pe-
riods longer than 50 days located in the nearby star-
forming regions. While in our optical observations these
Guenther et al.: PMS spectroscopic binaries
5
only about 0.1 M⊙. However, it is still possible that the
orbital period is even longer than 2482 days.
Fig. 3. The figure shows a histogram of the RV measure-
ments obtained for HR 5777. The published RV of this star
is +49.12±0.06 kms−1 (Murdoch 1993) which is shown as
a dashed line. The FEROS measurements are consistent
with this value. The FWHM of a Gaussian fitted to this
the distribution is 0.30 kms−1. The error of the measure-
ments thus is ± 0.15 kms−1.
systems usually appear as SB1s, many of these can be
converted into SB2s,
if high-resolution infrared spec-
tra were taken (e.g. Prato et al. 2002). Table 1 gives
a short overview of all known long-period binaries in
nearby southern star-forming regions found during this
study (CS Cha, RXJ1220.6-7539, MO Lup, RXJ1534.1-
3916, RXJ1559.2-3814, GSC 06209-00735, GSC 06213-
00306, BS Indi) and by other authors. In the follow-
ing, we give details on the individual systems listed in
Table 1. Additionally to the binaries found in this survey,
we also list in Tab. 1 the long-period spectroscopic bina-
ries published elsewhere (HIP50796, Torres et al. 2003;
HD97131, Torres et al. 2003; NTTS160814-1857, Mathieu
1994; Haro 1-14c, Reipurth et al. 2002, Simon & Prato
2004; NTTS162819-2423s, Mathieu 1994). While we can
not completely rule out that the VW Cha and GSC 06793-
00569 are binaries, these objects are in any case unsuitable
for the VLTI observations, and thus not listed in Tab. 1.
4.1. CS Cha
CS Cha is a CTTS of solar abundance (Padgett 1996).
This object was observed by Ghez et al. (1997) by means
of speckle-imaging but no companion was found within
0.′′1. However, Takami et al. (2003) already conclude from
the fact that there is a large gap in the inner disk that this
object might actually be a binary. The basic properties
of the disk of this star were derived by Henning et al.
(1993). We took 32 spectra of this object over a period
of 2570 days Table 2. CS Cha is an SB1-binary system
with a very long period. A possible orbit of 2482 days is
shown together with the RV-measurements in Fig. 4. The
minimum mass of the companion in this case would be
Fig. 4. The figure shows the RV-measurements of the
CTTS CS Cha, together with a possible orbital solution
with a period of 2482 days. However, we can not rule out
that the period is even longer than that.
4.2. RXJ1220.6-7539
RXJ1220.6-7539 is a WTTS. In most of our spectra the
Hα line is filled in. In some occasions we observe a small
double line emission profile with an equivalent width of
only −0.4 A. At other occasions, Hα appears in absorp-
tion but the equivalent width then is only 0.08 A. We
took 33 spectra of this star over a time-span of 2587
days. The object is an SB1-binary with an orbital period
of 613.9 ± 0.4 days and a eccentricity of 0.225 ± 0.005.
Fig. 5 shows the phase-folded RV-measurements. The RV-
measurements are given in Tab. 3, and the orbital elements
in Tab. 4.
4.3. RXJ1534.1-3916
RXJ1534.1-3916 is a WTTS, with Hα being in absorption.
We took 24 spectra of this star over a time-span of 2574
days. The RV-values are shown in Tab. 5, and Fig. 6. The
orbital period is certainly much longer than the 2574 days
over which we have observed the star. Thus we can not
derive an orbit yet. The eccentricity has to be very high
6
Guenther et al.: PMS spectroscopic binaries
Table 1. Spectroscopic binaries with periods longer than 50 days
this srvey/ EW Hα
region
HIP507962
CS Cha
HD971314
RXJ1220.6-7539
MO Lup5
RXJ1534.1-3916
RXJ1559.2-3814
GSC 06209-00735
NTTS160814-18576
GSC 06213-00306
Haro 1-14c7
NTTS162819-2423s6
BS Indi8
no / TWA
yes / Cha
no / TWA
yes / Cha
yes / Lup
yes / Lup
yes / Lup
yes / SC
no / SC
yes / SC
no / Oph
no / Oph
yes / Tuc
0.203
-40
fi
-2.3
abs
-1.4
0.3
0.7
fi
[A]1
EW LiI
spec type
[A]
K5/WTTS
0.53 ± 0.01 K4/CTTS
F2
0.21 ± 0.06 K2/WTTS
0.37 ± 0.02 K7/WTTS
0.21 ± 0.02 K1/WTTS
0.23/0.14 WTTS
0.37 ± 0.01 K2/WTTS
K2/WTTS
0.24/0.18 WTTS
K3/WTTS
G8/WTTS
0.18 ± 0.02 K0/WTTS
mK
type
RA
[mag]
7.66 ± 0.03
8.20 ± 0.03
7.70 ± 0.02
7.93 ± 0.02
8.64 ± 0.02
8.55 ± 0.02
9.34 ± 0.03
8.43 ± 0.02
7.69 ± 0.02
7.43 ± 0.02
7.78 ± 0.03
7.44 ± 0.02
6.57 ± 0.02
10 22 18.0
11 02 26.3
11 10 34.2
12 20 34.4
15 24 03.5
15 34 07.4
15 59 16.1
16 08 14.8
16 11 09.0
16 13 18.5
16 31 04.4
16 31 20.0
21 20 59.8
Dec
(2000.0)
-10 32 15
-77 33 36
-30 27 19
-75 39 29
-32 09 51
-39 16 18
-38 14 42
-19 08 33
-19 04 45
-22 12 48
-24 04 33
-24 30 04
-52 28 40
[days]
type
(2000.0)
SB1
SB1
ST3
SB1
ST3
SB1
SB2
SB1
SB1
SB2
SB2
SB1
SB1
period
570
≥ 2482
134
613
> 3000
> 3000
474
2045
145
167
591
89
1222
em
abs
1 em=emission, abs=absorption fi=filled in
2 triple system Torres et al. (2003)
3 Song et al. (2002)
4 Torres et al. (2003), possibly also a triple system
5 triple system consisting of a binary with 12 days, and one of with a period > 3000 days. (Esposito et al. 2006).
6 Mathieu (1994)
7 Reipurth et al. (2002), Simon & Prato(2004), eccentricity 0.617 ± 0.008, f (m) = 0.018 ± 0.001 M⊙
8 triple system, see Guenther et al. (2005)
figure
Fig. 5. The
phase-folded RV-
measurements of RXJ1220.6-7539. The orbital period is
612.9 ± 0.4 days.
shows
the
(≈ 0.9). RXJ1534.1-3916 nicely illustrates the difficulty in
finding long-period spectroscopic binaries with eccentric
orbits. If we would not have taken the first two spectra, we
probably would have classified this object as a single star.
According to Hogeveen (1992) only 0.45% of the binaries
have e ≥ 0.9.
Fig. 6. The figure shows
the RV-measurements of
RXJ1534.1-3916 The orbital period is apparently longer
than 7 years.
4.4. RXJ1559.2-3814
RXJ1559.2-3814 is a WTTS and an SB2. The average
equivalent width of Hα is −1.4 ± 0.2 A. The equivalent
Table 2. CS Cha
Table 3. RXJ1220.6-7539
Guenther et al.: PMS spectroscopic binaries
7
HJD
RV [km s−1]
245 1264.54715
245 1332.55406
245 1621.62514
245 2026.54612
245 2031.51533
245 2089.60062
245 2093.54486
245 2097.50960
245 2098.55118
245 2319.67764
245 2322.65188
245 2373.56573
245 2710.51867
245 2717.55061
245 2723.53930
245 3097.51064
245 3104.56677
245 3105.53213
245 3122.55011
245 3137.56055
245 3155.57941
245 3455.56786
245 3168.54063
245 3501.61418
245 3524.53172
245 3552.52001
245 3601.52246
245 3811.53842
245 3811.58558
245 3811.60357
245 3813.78161
245 3834.09873
15.7 ± 0.2
15.5 ± 0.1
15.9 ± 0.3
14.7 ± 0.1
14.8 ± 0.1
13.5 ± 0.6
14.1 ± 0.1
14.5 ± 0.1
14.1 ± 0.2
13.0 ± 0.3
13.3 ± 0.3
12.9 ± 0.3
13.3 ± 0.4
13.2 ± 0.3
13.3 ± 0.2
13.5 ± 0.2
14.6 ± 0.1
14.2 ± 0.2
13.6 ± 0.2
13.7 ± 0.2
13.7 ± 0.3
14.6 ± 0.1
14.6 ± 0.1
14.9 ± 0.1
14.9 ± 0.1
13.5 ± 0.1
14.6 ± 0.2
15.8 ± 0.2
15.8 ± 0.2
15.8 ± 0.2
15.6 ± 0.2
14.9 ± 0.2
HJD
245 1262.64298
245 1333.65495
245 1621.67014
245 1737.48553
245 2089.58864
245 2093.53472
245 2097.55429
245 2098.52034
245 2319.78210
245 2372.63728
245 2373.61486
245 2384.59905
245 2395.67658
245 2710.65545
245 2717.69773
245 2723.66034
245 3096.66843
245 3097.71156
245 3104.70590
245 3105.63828
245 3122.64088
245 3137.72151
245 3168.57480
245 3455.66760
245 3522.60916
245 3550.54048
245 3811.55386
245 3813.79633
245 3835.16435
245 3849.21861
RV [km s−1]
0.7 ± 0.1
−5.4 ± 0.1
7.9 ± 0.1
7.5 ± 0.1
1.2 ± 0.1
1.7 ± 0.1
2.0 ± 0.1
1.7 ± 0.1
8.4 ± 0.1
7.6 ± 0.1
7.2 ± 0.1
7.1 ± 0.1
6.2 ± 0.1
1.4 ± 0.1
2.2 ± 0.1
2.6 ± 0.1
0.9 ± 0.1
0.9 ± 0.1
0.4 ± 0.1
0.2 ± 0.1
−1.4 ± 0.1
−2.8 ± 0.1
−5.1 ± 0.1
7.8 ± 0.1
8.3 ± 0.1
7.8 ± 0.1
−6.1 ± 0.1
−6.2 ± 0.1
−5.6 ± 0.1
−4.8 ± 0.1
Table 4. Orbital elements of RXJ1220.6-7539
width of the LiI 6708 A line is 0.23 A for the primary,
and 0.14 A for the secondary component. We find that
the ratio of the height of the peak of the cross-correlation
function of the two components is 1.9 ± 0.2. Tab. 6 gives
the RV measurements obtained for both components and
the orbital elements are listed in the Tab. 7. Good fit is
obtained for a period of 474 days (Fig. 7).
element
P
T0 [HJD]
γ
K1
e
ω
a1sin i
f(m)
value
613.9 ± 0.4 d
2450123 ± 3
2.74 ± 0.03 km s−1
7.15 ± 0.07 km s−1
0.225 ± 0.005
171.7 ± 1.9o
0.399 ± 0.005 AU
0.0228 ± 0.0008 M⊙
4.5. GSC 06209-00735
4.6. GSC 06213-00306
GSC 06209-00735 was first identified as a member Upper
Sco association based on the spectroscopic and X-ray
properties by Preibisch et al. 1998. We confirm the large
equivalent width of the LiI 6708 A line of 0.37 ± 0.01 AA,
and found that the object is a SB1. Hα is in absorption,
the equivalent width is 0.3 ± 0.1 A. In total, we took 26
spectra within 2287 days. We find an orbital period of
2045 ± 16 days, which is only slightly lower than the time
over which we observed the star. Fig. 8 shows the phase-
folded RV-measurements together with the orbit. The RV-
measurements are given in Tab. 8, and the orbital elements
in Tab. 9.
GSC 06213-00306 is another SB2 WTTS. The equivalent
widths of the Li I 6708 A line are 0.241 ± 0.004 A, and
0.18 ± 0.04 A, for the primary and secondary, respectively.
The Hα-line is often filled in, occasionally it appears as an
emission line with an equivalent width of -0.1 A, at other
occasions as an absorption line with an equivalent width of
0.1 A. In total, we took 22 spectra of this binary. We find
an orbital solution with a period of 166.9±0.1 days. Thus,
GSC 06213-00306 is a long-period young SB2-binary, the
orbital elements for both components have been solved.
Given that it has a K-magnitude of 7.43±0.02 it is an ideal
VLTI target. The RV-measurements are listed in Tab. 10,
8
Guenther et al.: PMS spectroscopic binaries
Table 5. RXJ 1534.1-3916
Table 6. RXJ 1559.2-3814
HJD
245 1261.76390
245 1335.77111
245 2093.73652
245 2372.74166
245 2384.71011
245 2395.72062
245 2710.78396
245 2717.78285
245 2723.77086
245 3096.73954
245 3104.77290
245 3105.73478
245 3122.76000
245 3137.77322
245 3168.62275
245 3501.73605
245 3523.74779
245 3556.77084
245 3556.78210
245 3601.55318
245 3603.51374
245 3811.68239
245 3813.84489
245 3835.23052
RV [km s−1]
−25.0 ± 0.1
−27.5 ± 1.0
−12.7 ± 1.5
−10.6 ± 0.1
−10.9 ± 0.5
−11.4 ± 0.4
−12.3 ± 0.5
−12.0 ± 0.2
−11.3 ± 0.2
−12.4 ± 0.7
−11.3 ± 0.4
−11.5 ± 0.5
−13.0 ± 0.1
−12.6 ± 0.4
−11.6 ± 0.1
−13.7 ± 0.1
−13.8 ± 0.1
−11.3 ± 0.2
−11.1 ± 0.5
−13.6 ± 0.3
−10.8 ± 0.2
−12.0 ± 0.2
−12.1 ± 0.7
−11.9 ± 0.3
HJD
245 1264.71496
245 1331.71643
245 1622.78442
245 1623.78926
245 1625.71327
245 3097.64088
245 3105.69324
245 3122.77331
245 3137.78841
245 3168.81587
245 3168.82963
245 3501.75306
245 3503.78331
245 3523.76173
245 3556.79982
245 3556.81105
245 3601.56957
245 3603.53176
245 3811.70502
245 3811.72015
245 3813.81043
245 3813.82554
245 3838.28599
−6.7 ± 0.3
−7.9 ± 0.2
19.6 ± 0.4
20.6 ± 0.8
19.1 ± 0.9
4.2 ± 0.11
2.6 ± 0.41
0.9 ± 0.31
−2.8 ± 0.9
−8.1 ± 0.4
−7.8 ± 0.3
18.6 ± 0.1
19.2 ± 0.2
18.5 ± 0.2
7.7 ± 1.61
8.1 ± 1.31
0.6 ± 0.41
−0.1 ± 0.61
2.0 ± 1.11
1.4 ± 1.01
1.4 ± 0.71
1.2 ± 0.51
3.1 ± 1.01
RV [km s−1]
RV [km s−1]
A component B component
12.2 ± 0.6
12.3 ± 1.1
−14.2 ± 0.4
−14.6 ± 0.6
−15.2 ± 0.7
8.4 ± 0.8
9.1 ± 1.0
−16.4 ± 1.1
−15.1 ± 0.4
−15.2 ± 1.1
1 Just A-component.
Table 7. Orbital elements RXJ 1559.2-3814
0.5
1
phase
1.5
2
0.5
1
phase
1.5
2
element
P
T0 [HJD]
γ
K1
K2
e
ω1
ω2
a1sin i
a2sin i
q = m2
m1
m1sin3 i
m2sin3 i
value
474.0 ± 1.0 d
2447834 ± 3
2.0 ± 0.4 km s−1
13.4 ± 0.2 km s−1
14.2 ± 0.2 km s−1
0.336 ± 0.005
336.8 ± 1.9o
156.8 ± 1.9o
0.549 ± 0.010 AU
0.581 ± 0.010 AU
0.945 ± 0.027
0.444 ± 0.026 M⊙
0.419 ± 0.025 M⊙
20
10
0
]
s
/
m
k
[
V
R
−10
−20
0
l
]
s
/
m
k
[
s
a
u
d
s
e
R
i
6
4
2
0
−2
−4
0
Fig. 7. The figure shows the RV-measurements of RXJ
1559.2-3814. Shown is the orbit of the A and the B com-
ponent for a period of 474 days.
the orbit in Tab. 11. The RV-measurements and the orbit
are shown in Fig. 9. It is interesting to note that the masses
of both components are almost identical.
5. Results: Spectroscopic binary candidates
In this section we discuss two objects that were suspected
to be spectroscopic binaries of long period. We argue that
in both cases it is unlikely that these are spectroscopic
binaries.
5.1. VW Cha
VW Cha was already observed by Melo (2003) with
FEROS, who found an occasional doubling of the cross-
correlation peak and thus suggested that this object is a
spectroscopic binary. Brandner et al. (1996) found that
VW Cha is a visual binary with a separation of 0.′′72 ±
0.′′03 . The companion is only 0.25 mag fainter in J than
the primary. Using K-band speckle observations Ghez et
al. (1997) confirmed the presence of this companion, and
determine a separation of 0.′′66 ± 0.′′03. Surprisingly, Ghez
et al. 1997 give a flux-ratio in the K-band of 4.5 ± 0.7.
Additionally to this inner component, Ghez et al. (1997)
detected another component with a separation of 17 ± 2′′.
Table 8. GSC 06209-00735
Table 9. Orbital elements of GSC 06209-00735
Guenther et al.: PMS spectroscopic binaries
9
HJD
245 1260.79532
245 1333.79536
245 1622.87285
245 1737.59186
245 2093.77154
245 2395.75208
245 2396.70830
245 2372.81128
245 2384.73607
245 2710.79456
245 2717.79687
245 2723.78469
245 3096.76265
245 3097.75000
245 3104.79092
245 3105.87514
245 3122.78632
245 3137.81404
245 3168.63464
245 3601.59793
245 3616.52275
245 3811.75370
245 3838.36166
245 3847.35257
RV [km s−1]
−5.7 ± 0.1
−6.7 ± 0.1
−9.3 ± 0.1
−10.1 ± 0.2
−10.4 ± 0.4
−8.6 ± 0.1
−8.5 ± 0.1
−8.7 ± 0.1
−8.4 ± 0.1
−6.8 ± 0.1
−6.4 ± 0.1
−6.5 ± 0.1
−4.1 ± 0.1
−4.2 ± 0.1
−4.3 ± 0.1
−4.3 ± 0.1
−4.4 ± 0.1
−4.5 ± 0.1
−5.0 ± 0.1
−8.8 ± 0.1
−8.8 ± 0.1
−9.7 ± 0.1
−9.5 ± 0.1
−9.8 ± 0.1
figure
the
Fig. 8. The
phase-folded RV-
measurements of GSC 06209-00735. The orbital period is
2045 ± 16 days.
shows
As pointed out above, FEROS was operated up to
October of 2002 (HJD 245 2550) at the ESO 1.5-m tele-
scope, and was then moved to the MPG/ESO-2.20-m tele-
scope. The entrance apertures were 2.′′7 and 2.′′0 on the
ESO 1.5-m telescope, and MPG/ESO 2.2-m telescope, re-
spectively. Thus, if the fibre is centred on the primary,
element
P
T0 [HJD]
γ
K1
e
ω
a1sin i
f(m)
value
2045 ± 16 d
2451022 ± 12
−7.62 ± 0.02 km s−1
2.87 ± 0.05 km s−1
0.20 ± 0.03
8.1 ± 2.3o
0.53 ± 0.018 AU
0.0049 ± 0.0005 M⊙
Table 10. GSC 06213-00306
HJD
RV [km s−1]
RV [km s−1]
A component B component
245 1333.82932
245 1622.89550
245 1623.75952
245 1624.73672
245 1625.73770
245 1737.66113
245 2089.46697
245 2093.50599
245 2097.59641
245 2396.72380
245 2372.76696
245 2384.74816
245 2385.71382
245 2710.82314
245 2717.82525
245 2723.81319
245 1260.89577
245 1260.89577
245 3603.57132
245 3811.76438
245 3813.87632
245 3838.38271
−5.5 ± 0.51
−24.3 ± 0.1
−24.1 ± 0.1
−24.8 ± 0.2
−25.3 ± 0.3
−18.0 ± 0.2
−9.0 ± 0.21
−7.0 ± 0.21
−6.9 ± 0.11
4.5 ± 0.8
4.1 ± 0.9
5.3 ± 0.4
4.8 ± 0.6
4.4 ± 0.5
6.4 ± 0.5
5.5 ± 0.5
−6.0 ± 0.41
−6.1 ± 0.51
5.7 ± 0.6
−2.6 ± 1.02
−2.8 ± 1.02
−7.0 ± 0.42
1 both components, unresolved.
2 just A-component.
10.9 ± 0.3
11.4 ± 0.2
11.9 ± 0.2
12.4 ± 0.3
4.1 ± 0.5
−18.6 ± 0.2
−17.5 ± 0.2
−18.6 ± 0.1
−18.8 ± 0.3
−18.8 ± 0.3
−19.2 ± 0.2
−19.1 ± 0.2
−19.7 ± 0.9
the secondary would almost be at the edge. The amount
of light from the primary and secondary thus would vary
depending on the exact placement of the fibre and the see-
ing. The question thus arises whether the occasional dou-
bling of the spectral-lines is due to this effect, or whether
there is really another component. If VW Cha were a spec-
troscopic binary, its separation would have to be much
smaller than 0.′′72 ± 0.′′03, which implies that we should
observe large RV-variations.
In order to clarify the situation, we took 16 spectra of
this star. The cross-correlation peak is in fact asymmetric.
In contrast to our expectation, we find only rather mod-
est variations of the RV (Tab. 12). Additionally, the RV-
measurements are erratic, and not periodic. While we can
not completely rule out that this is an SB2, it seems more
likely that the asymmetric form of the cross-correlation
is due to the visual companion moving in and out of the
fibre, and not due to a spectroscopic companion. This hy-
pothesis would also explain the observed scatter of the
10
]
s
/
m
k
[
V
R
20
10
0
−10
−20
−30
0
l
]
s
/
m
k
[
s
a
u
d
s
e
R
i
2
0
−2
0
0.5
1
phase
1.5
2
0.5
1
phase
1.5
2
figure
the
Fig. 9. The
measurements of GSC 06213-00306,
the residuals. The orbital period is 166.87 ± 0.13 days.
phase-folded RV-
together with
shows
Table 11. Orbital elements GSC 06213-00306
element
P
T0 [HJD]
γ
K1
K2
e
ω1
ω2
a1sin i
a2sin i
q = m2
m1
m1 sin3 i
m2 sin3 i
value
166.9 ± 0.1 d
2452124.2 ± 1.1
−6.76 ± 0.06 km s−1
15.1 ± 0.1 km s−1
15.65 ± 0.09 km s−1
0.226 ± 0.006
161.8 ± 2.2o
341.8 ± 2.2o
0.226 ± 0.002 AU
0.234 ± 0.002 AU
0.97 ± 0.01
0.246 ± 0.006 M⊙
0.239 ± 0.006 M⊙
RV-measurements, because the velocity difference of two
stars of one solar-mass are in a circular orbit with a sepa-
rtion of 0.′′7 is about 4 km s−1. This object is in any case
unsuitable for VLTI-observations.
Guenther et al.: PMS spectroscopic binaries
Table 12. VW Cha
HJD
245 1262.57801
245 1333.57002
245 2031.53243
245 2319.70304
245 2322.67626
245 2372.53625
245 2384.53498
245 2710.59595
245 2717.59503
245 2723.58209
245 3096.65990
245 3104.58111
245 3105.55467
245 3122.59151
245 3137.59880
245 3155.64238
245 3122.59151
245 3522.57695
245 3524.54647
RV [km s−1]
14.9 ± 0.5
15.4 ± 0.2
17.2 ± 0.2
19.8 ± 1.0
18.6 ± 0.6
18.4 ± 1.2
15.3 ± 0.4
16.3 ± 0.2
16.3 ± 0.4
22.0 ± 2.4
18.7 ± 0.9
15.2 ± 0.2
17.5 ± 0.1
18.4 ± 0.5
16.1 ± 0.4
14.6 ± 2.6
18.5 ± 0.4
16.4 ± 0.1
18.3 ± 0.4
Table 13. GSC 06793-00569
HJD
245 1260.90498
245 1333.84131
245 1737.67610
245 2089.47704
245 2093.64727
245 2097.60976
245 2373.71233
245 2385.67462
245 2395.83337
245 2396.81929
245 2710.86108
245 2717.86133
245 2723.82421
245 3097.68383
245 3104.81253
245 3105.78900
245 3122.83664
245 3137.83685
245 3168.70120
RV [km s−1]
−6.5 ± 0.2
−4.2 ± 0.3
−6.5 ± 0.2
−6.8 ± 0.4
−4.5 ± 0.4
−8.6 ± 0.9
−5.7 ± 0.2
−5.1 ± 0.1
−5.8 ± 0.1
−5.7 ± 0.1
−7.8 ± 0.1
−6.3 ± 0.1
−7.0 ± 0.1
−5.8 ± 0.1
−5.2 ± 0.1
−6.2 ± 0.1
−5.5 ± 0.1
−6.4 ± 0.1
−6.9 ± 0.1
6. Results: Single stars
5.2. GSC 06793-00569
GSC 06793-00569 is a WTTS with a weak Hα-emission
line of -0.4 A. We have taken 19 spectra of this star.
Although the cross-correlation occasionally shows a blue
asymmetry, and on other occasions a red asymmetry, the
RV-variations are moderate, and we do not find any sig-
nificant periodicity (Tab. 13). We thus interpret the data
in the same way as for VW Cha, the star could be a visual
binary with such a separation that roughly corresponds to
the radius of the fibre.
Unfortunately, not all previous surveys for pms-binaries
give a list of the objects that were found to be single.
Such a list is however very useful, because it avoids dupli-
cation of work. In Table 18, 19, and 20 we list the average
RV-values obtained for all single stars together with the
variance of the RV-values determined for each star. Also
given in these tables are the average equivalent widths of
Hα and the Li I 6708 A line, as derived from the spectra.
The K-band brightness is taken from the 2MASS All-Sky
Catalogue of Point Sources (Cutri et al 2003). The number
of spectra taken for each star is given in the last column.
Guenther et al.: PMS spectroscopic binaries
11
The RV-curve of a star with a spot is periodic, where
the period corresponds to the rotation period of the star.
However, once this spot vanishes and other spots appear
at differnt longitudes, the amplitude and the phase of the
RV-signal changes. As a result, the periodic signal van-
ishes, and the power in a periodogram decreases. In con-
trast to this, the signal of an orbiting companion remains
unchaged and thus, the power in a periodogram always
increases, if more data is added. The differnce of the RV-
signal caused by activity, and cuased by an orbiting com-
panion is nicely illustrated for EK Dra, ǫ Eri, and β Gem.
EK Dra is an active binary star with an orbital period
of 45 years, ǫ Eri is an active star with a planet, and β
Gem an osciallting star with a planet (Konig et al. 2005;
Hatzes et al. 2000; Hatzes et al. 2006). For more informa-
tion on the effects of stellar activity see also Paulson et
al. 2002), and Saar et al. 1998. Variations of the RV due
to stellar activity have already been observed in T Tauri
stars. Because of the large activity level of these stars, the
effects are even larger than for older stars. RU Lup for
example, shows RV-variations with an amplitude of 2.17
km s−1 (Stempels et al 2007). It is thus not surpising that
we observe the same effects in the T Tauri stars that we
observed. Nice examples are RXJ1415.0-7822 (Tab. 14),
HK Lup, Tab. 15), GSC 06793-01406 Tab. 16), and GSC
06793-00994 (Fig. 11), of which we took 13 to 19 spectra
each. All of these stars show RV-variations with a semi-
amplitude of 1 to 3 km s−1 but in none of them a signifi-
cant period could be detected. Fig. 10 shows a histogram
of the RV-variations observed in single stars.
The survey is limited to companions that cause a RV-
amplitude of ≥ 3 km s−1. For example, a system consisting
of a 1.0 M⊙ primary, and a 0.1 M⊙ secondary would only
be detectable, if the period were equal, or shorter than
a year. Similarly, if the companion would have an orbital
period of 3000 days, the mass of the companion would
have to be ≥ 0.23 M⊙ in order to be detectable. The
survey thus is incomplete for long-period companions with
masses ≤ 0.2 M⊙.
We took 13 to 21 spectra of all those stars which show
unusually large RV-variations, in order to find out whether
these are caused by stellar activity, or by an orbiting com-
panion. Apart from the objects listed in Table 1, none of
them turned out to be a binary. Only in the case of the
two rapidly rotating stars GSC 06781-01046 and MN Lup,
we can not fully exclude that they are binaries. One of the
spectra of GSC 06781-01046 shows a double-line appear-
ance. This object thus might be an SB2 but more spectra
are needed.
7. Discussion and conclusions
The goal of the survey was to compile a list of young
long-period binary stars suitable for calibrating the evo-
lutionary tracks by measuring the masses by combining
spectroscopic and VLTI data. Although it was not our in-
tention to carry out a statistical compleate survey, the sur-
vey does contains some useful informtion on the frequency
Fig. 10. Histogram of the RV-variations observed for the
stars listed in Table 18, 19 20. The amplitude of RV-
variations caused are typically less than 3 km s−1.
Fig. 11. A typical example for the non-periodic variations
of a young star. Shown are the RV-values of GSC 06793-
00994.
of young binaries. Mathieu (1994) listed in his classical re-
view article only 3 pms-spectroscopic binaries located in
the nearby southern star-forming regions (Chamaeleon,
Corona Australis, Lupus, Sco-Cen, ρ Ophiuchi) with or-
bital periods longer than 50 days, and 4 short period sys-
tems in these regions. Another short-period binary sys-
tem (RXJ1108.8-7519) was found by Covino et al. (1997).
The long period binaries are NTTS162819-2423S (period
89.1 days, ρ Oph), Wa Oph 1 (=NTTS160814-1857, pe-
riod 144.7 days, SC), and Haro 1-14c (period 591 days, ρ
Oph) (see also Reipurth et al. 2002). NTTS162819-2423S
and Haro 1-14c are both members of hierarchical systems.
The distance between the spectroscopic binary Haro 1-
14c and Haro 1-14 is 12.′′9, and NTTS162819-2423S is
separated from spectroscopic binary NTTS162819-2423N
by 6.′′. Additionally to these stars, Melo (2003) identified
VW Cha (ChaI), BF Cha (ChaII), CHX 18N (ChaI), and
AS 205 (SC) as spectroscopic binaries but did not derive
the orbital periods of these objects. However, we regard
VW Cha not as a spectroscopic binary. Another spectro-
scopic binary found is RXJ1603.8-3938 (Guenther et al.
2001). Thus, in all previous surveys, 12 SBs were found, 6
have periods longer than 50 days, of two are triple stars.
12
Guenther et al.: PMS spectroscopic binaries
Table 18. The single stars: Chamaeleon
region
EW Hα1
EG Cha
EO Cha
EQ Cha3
RXJ0850.1-7554
RXJ0915.5-7609
RXJ0917.2-7744
HD84075
HD86356
HD86588
RXJ1001.1-7913
RXJ1005.3-7749
SX Cha E4
SY Cha
TW Cha
CR Cha (=Sz6)
CT Cha
DI Cha5
VW Cha6
WW Cha2
HBC 584
Sz 417
VZ Cha2
CV Cha2, 5
GSC 07739-02180
RXJ1129.2-7546
RXJ1140.3-8321
RXJ1150.4-7704
T Cha
RXJ1159.7-7601
RXJ1202.1-7853
HD106772
RXJ1219.7-7403
RXJ1220.4-7407
HD107722
RXJ1233.5-7523
RXJ1239.4-7502
BC Cha
RXJ1325.7-7955
CPD-75 902
RXJ1415.0-7822
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
SC
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
Cha
[A]
−1.1 ± 0.1
−1.0 ± 0.2
−6.5 ± 1.1
abs
−1.1 ± 0.2
abs
abs
fi
abs
−2.5 ± 0.7
−4.2 ± 0.1
−20 ± 5
−13 ± 2
−28 ± 11
−34 ± 2
−40 ± 21
−17 ± 1
−57 ± 11
−57 ± 13
−77 ± 1
−2.0 ± 0.9
−32 ± 4
−67 ± 16
−3.9 ± 1.2
fi
−0.3 ± 0.2
−1.3 ± 0.5
−8.8 ± 9.9
−0.2 ± 0.1
−3.3 ± 0.5
abs
−4.5 ± 1.6
−2.9 ± 0.6
abs
abs
abs
−72 ± 18
abs
abs
fi
EW LiI
[A]
spec
type
0.53 ± 0.01
K7
0.53 ± 0.04 M0
0.57 ± 0.04 M3
0.27 ± 0.01
G5
K6
0.55 ± 0.03
G2
0.33 ± 0.06
G2
0.17 ± 0.01
G8
0.34 ± 0.05
0.11 ± 0.01
F6
0.10 ± 0.01
K8
0.58 ± 0.01 M1
0.65 ± 0.04 M0
0.56 ± 0.03 M0
K7
0.41 ± 0.01
K2
0.43 ± 0.04
K7
0.40 ± 0.05
0.26 ± 0.01
G2
0.44 ± 0.05 M0.5
K5
K7
K0
K6
G8
0.3 ± 0.2
0.3 ± 0.1
0.43 ± 0.02
0.31 ± 0.01
0.34 ± 0.02
0.52 ± 0.01
0.49 ± 0.02
0.21 ± 0.01
0.40 ± 0.06
0.41 ± 0.03
0.46 ± 0.01
0.56 ± 0.02
0.21 ± 0.01
0.56 ± 0.01
0.61 ± 0.01
0.08 ± 0.01
0.14 ± 0.01
0.41 ± 0.01
0.52 ± 0.03
0.10 ± 0.05
0.22 ± 0.01
0.26 ± 0.05
K3
K2
K2
G8
K2
K7
G2
K8
K7
F6
K0
K2
K1
K0
G5
mK
[mag]
7.338 ± 0.021
8.732 ± 0.021
8.410 ± 0.031
8.704 ± 0.019
8.488 ± 0.033
8.812 ± 0.023
7.160 ± 0.016
8.040 ± 0.029
7.994 ± 0.029
9.216 ± 0.021
8.892 ± 0.019
8.685 ± 0.024
8.631 ± 0.019
8.616 ± 0.021
7.310 ± 0.023
8.661 ± 0.021
6.217 ± 0.020
6.962 ± 0.026
6.083 ± 0.053
9.175 ± 0.024
7.999 ± 0.031
8.242 ± 0.038
6.845 ± 0.026
8.053 ± 0.029
8.878 ± 0.024
8.635 ± 0.019
8.970 ± 0.021
6.954 ± 0.018
8.304 ± 0.027
8.307 ± 0.023
6.231 ± 0.031
8.858 ± 0.023
8.367 ± 0.022
7.135 ± 0.034
7.756 ± 0.040
7.777 ± 0.021
9.354 ± 0.021
8.932 ± 0.022
8.015 ± 0.053
9.876 ± 0.023
RA
(2000.0)
08 36 56.2
08 44 31.9
08 47 59.2
08 50 05.4
09 15 29.1
09 17 11.4
09 36 17.8
09 51 50.7
09 53 13.7
10 01 08.8
10 05 17.6
10 55 59.8
10 56 30.5
10 59 01.1
10 59 07.0
11 04 09.1
11 07 20.7
11 08 01.8
11 10 00.7
11 10 49.6
11 12 24.5
11 09 23.8
11 12 27.7
11 21 05.5
11 29 12.7
11 40 16.6
11 50 28.9
11 57 13.5
11 59 42.3
12 02 03.8
12 17 26.9
12 19 43.8
12 20 21.9
12 23 29.0
12 33 32.0
12 39 21.3
13 01 59.0
13 25 42.8
13 49 12.9
14 15 01.8
Dec
RV
[kms−1]
(2000.0)
17.4 ± 0.2
-78 56 46
17.1 ± 0.3
-78 46 31
17.9 ± 1.0
-78 54 58
18.1 ± 1.0
-75 54 38
20.3 ± 0.1
-76 08 47
5.7 ± 1.8
-77 44 15
5.2 ± 0.1
-78 20 42
13.2 ± 0.4
-79 01 38
2.4 ± 0.5
-79 33 28
14.4 ± 0.4
-79 13 08
16.3 ± 0.2
-77 49 06
13.9 ± 0.5
-77 24 40
14.0 ± 0.3
-77 11 39
17.8 ± 0.1
-77 22 41
16.7 ± 1.1
-77 01 40
15.2 ± 0.3
-76 27 19
14.6 ± 0.2
-77 38 07
17.2 ± 0.5
-77 42 29
15.0 ± 2.0
-76 34 59
13.6 ± 0.1
-77 17 52
14.6 ± 0.3
-76 37 06
16.3 ± 0.4
-76 23 21
16.1 ± 1.2
-76 44 22
12.1 ± 0.7
-38 45 17
14.6 ± 0.5
-75 46 26
12.7 ± 0.1
-83 21 00
−3.3 ± 1.0
-77 04 38
14.0 ± 1.3
-79 21 32
13.8 ± 0.1
-76 01 26
-78 53 01
17.1 ± 0.2
-80 35 06 −12.9 ± 0.2
13.8 ± 0.1
-74 03 57
12.3 ± 0.4
-74 07 39
13.4 ± 0.4
-77 40 51
15.5 ± 0.1
-75 23 25
-75 02 39
13.9 ± 0.2
14.3 ± 0.2
-77 51 22
−8.4 ± 3.6
-79 55 25
−0.8 ± 0.1
-75 49 48
-78 22 12
−3.1 ± 0.3
σ RV No. of
spectra
3
2
8
2
3
3
3
2
3
2
2
5
3
2
3
13
2
17
2
2
18
2
2
3
3
3
4
17
3
3
2
2
2
2
2
2
2
4
2
18
[kms−1]
0.4
0.5
2.7
1.3
0.1
3.2
0.1
0.5
2.4
0.6
0.3
1.1
0.5
0.1
2.0
0.9
0.3
2.0
0.5
0.2
1.1
0.6
1.8
1.2
0.8
0.1
2.1
5.3
0.1
0.3
0.3
0.2
0.6
0.6
0.1
0.3
0.4
7.1
0.2
1.3
1 em=emission, abs=absorption fi=filled in
2 WW Cha, VZ Cha, CV Cha: also observed by Melo et al. (2003) and found to be single
3 EQ Cha : difficult object, cc-function asymmetric, unlikely to be binary
4 SX Cha : visual binary M0 primary and M3 secondary with a separation of 2.′′2. (Reipurth & Zinnecker 1993), only eastern
component observed
5 DI Cha (LkHα 332-17): binary with separation of 4.′′9 ± 0.′′2 (Ghez al. 1997)
6 VW Cha : triple system, separation of 0.′′66 ± 0.′′03 and 17.′′ ± 2.′′ (Ghez et al. 1997)
7 Sz 41: triple system, separation of 1.′′5 ± 0.′′8 and 12.′′4 ± 0.′′6 (Ghez et al. 1997)
To these we add 7 long-period SBs, and 3 short-period
ones. These short-period systems will be discussed in a
forthcoming paper (Covino et al. 2007). One of the newly
found long-period systems is a triple star. The newly found
tripple system is hirarchical. It consits of two components
which are separated by 0.′′26 ±0.′′03, were one of these com-
ponents is a binary with an orbital period of possibly 16
days. If we count the triple star as two binaries, because
it consist of a long and a short period system, the total
number of SBs in these regions is 25. By adding in the pre-
viously known binaries, the sample comprises 120 stars in
total. Thus, the frequency of SBs is 20 ± 5%. This num-
ber can now be compared with the results by Duquennoy
and Mayor (1991), who found a frequency of binaries with
Table 19. The single stars: Lupus, Scorpius Centaurus
Guenther et al.: PMS spectroscopic binaries
13
region
EW Hα1
LQ Lup
LS Lup
RXJ1516.6-4406
MM Lup
MN Lup3
MP Lup
MQ Lup
MR Lup
HIP758364
MT Lup
RXJ1538.6-3916
RXJ1539.2-4455
MU Lup
GSC 06785-00476
GSC 06781-010465
GW Lup2
HBC 6036
RXJ1555.4-3338
HD142987
GSC 06191-00552
GSC 06195-00768
HBC 609
RY Lup
MZ Lup
GSC 06204-00812
GSC 06784-01219
EX Lup2
HBC 613
NQ Lup
GSC 06784-00039
HK Lup2
V1094 Sco
GSC 06793-00868
GSC 06793-00797
GSC 06793-005697
GSC 06793-00994
GSC 06801-00186
GSC 06793-01406
GSC 06214-02384
GSC 06794-00156
GSC 06794-00537
GSC 06798-00035
GSC 06794-00337
Lup
Lup
Lup
Lup
Lup
Lup
Lup
Lup
Lup
Lup
Lup
Lup
SC
SC
Lup
Lup
Lup
SC
SC
SC
Lup
Lup
Lup
SC
SC
Lup
Lup
Lup
SC
Lup
Lup
SC
SC
SC
SC
SC
SC
SC
SC
SC
SC
SC
[A]
−1.5 ± 0.5
−0.7 ± 0.1
fi
fi
−6.1 ± 0.8
fi1
−0.5 ± 0.2
−1.7 ± 0.2
abs
−1.3 ± 0.5
abs
−1.8 ± 0.4
−0.8 ± 0.1
abs
abs
−52.8 ± 6.8
−10.4 ± 2.2
−0.7 ± 0.5
−2.1 ± 0.1
−0.6 ± 0.1
−0.5 ± 0.2
−30 ± 1.0
−3.2 ± 0.3
abs
−0.1 ± 0.1
−0.1 ± 0.1
−24 ± 8
−32 ± 2
−1.4 ± 0.1
fi
−22 ± 17
−8.2 ± 3.3
−2.8 ± 0.2
−0.7 ± 0.2
−0.4 ± 0.2
−0.3 ± 0.2
abs
−0.6 ± 0.2
fi
−0.3 ± 0.1
fi
fi
fi
spec
type
G8
K1
K2
≤ 0.2
∼ 0.3
EW LiI
[A]
0.26 ± 0.03
0.47 ± 0.01
0.54 ± 0.13
0.47 ± 0.01
0.44 ± 0.05 M2
K1
0.36 ± 0.01
K2
0.39 ± 0.03
K6
0.47 ± 0.03
0.04 ± 0.01
K0
K5
0.48 ± 0.01
0.39 ± 0.01
K4
0.17 ± 0.03
0.49 ± 0.01
0.30 ± 0.01
K6
G7
G5
0.48 ± 0.04 M0
0.58 ± 0.01 M0
0.44 ± 0.01
K5
G3
K3
0.48 ± 0.01
K7
0.46 ± 0.02
K8
0.39 ± 0.06
G0
0.38 ± 0.01
G8
0.33 ± 0.02
K4
0.47 ± 0.01
0.37 ± 0.01
G7
0.36 ± 0.08 M0
K8
0.46 ± 0.02
K7
0.51 ± 0.03
G7
0.37 ± 0.01
0.54 ± 0.03
K8
0.49 ± 0.02
K6
0.55 ± 0.01 M1
K4
0.52 ± 0.01
K1
0.47 ± 0.20
0.38 ± 0.01
G4
G5
0.35 ± 0.02
G7
0.36 ± 0.01
K0
0.41 ± 0.01
G6
0.34 ± 0.01
0.48 ± 0.02
K2
G1
0.35 ± 0.02
0.44 ± 0.01
K1
mK
[mag]
8.809 ± 0.021
9.454 ± 0.021
9.193 ± 0.019
9.260 ± 0.023
9.496 ± 0.019
8.930 ± 0.019
8.842 ± 0.021
8.963 ± 0.019
6.760 ± 0.020
9.376 ± 0.024
8.854 ± 0.023
9.989 ± 0.021
9.187 ± 0.026
8.920 ± 0.023
8.177 ± 0.020
8.630 ± 0.021
8.271 ± 0.026
9.353 ± 0.022
7.614 ± 0.021
8.325 ± 0.024
8.372 ± 0.025
8.608 ± 0.023
6.976 ± 0.019
8.528 ± 0.030
8.727 ± 0.025
8.461 ± 0.023
8.496 ± 0.021
8.724 ± 0.023
8.714 ± 0.021
7.908 ± 0.016
8.014 ± 0.021
8.658 ± 0.021
8.815 ± 0.034
8.455 ± 0.027
8.494 ± 0.019
8.608 ± 0.023
8.686 ± 0.022
8.102 ± 0.018
8.509 ± 0.019
7.084 ± 0.018
8.184 ± 0.024
7.695 ± 0.023
8.084 ± 0.026
RA
(2000.0)
15 08 37.8
15 15 52.8
15 16 36.8
15 23 25.7
15 23 30.4
15 24 32.4
15 25 33.2
15 25 36.7
15 29 26.9
15 38 02.7
15 38 38.3
15 39 12.0
15 40 41.2
15 41 06.8
15 42 49.9
15 46 44.7
15 51 47.0
15 55 26.3
15 58 20.6
15 58 47.8
15 57 02.4
15 59 16.5
15 59 28.4
16 01 09.0
16 03 02.7
16 05 50.5
16 03 05.5
16 07 10.0
16 08 18.3
16 08 43.4
16 08 22.5
16 08 36.2
16 11 56.4
16 13 02.8
16 13 29.3
16 14 02.1
16 14 59.3
16 16 18.0
16 19 34.0
16 24 51.4
16 23 07.8
16 23 32.3
16 27 39.6
Dec
RV
[kms−1]
(2000.0)
7.6 ± 1.5
-44 23 17
6.8 ± 0.4
-44 18 16
4.4 ± 1.3
-44 07 19
4.4 ± 0.1
-40 55 45
6.6 ± 3.4
-38 21 29
3.3 ± 0.3
-36 52 02
3.6 ± 0.1
-36 13 46
-35 37 32
−1.0 ± 1.9
-28 50 52 −19.5 ± 0.1
3.0 ± 0.2
-38 07 22
2.5 ± 0.2
-39 16 54
5.5 ± 0.1
-44 55 29
-37 56 18
2.2 ± 0.1
−2.0 ± 0.1
-26 56 26
-25 36 41
SB2?
−3.5 ± 0.5
-34 30 36
−2.6 ± 0.1
-35 56 43
-33 38 22
0.0 ± 0.1
6.0 ± 1.5
-18 37 25
−6.4 ± 0.5
-17 57 59
−5.1 ± 0.2
-19 50 41
-41 57 09
0.2 ± 0.7
−0.4 ± 0.5
-40 21 51
1.8 ± 0.2
-33 20 14
−4.9 ± 0.1
-18 06 04
−4.2 ± 0.1
-25 33 12
-40 18 25
−0.3 ± 0.1
−0.5 ± 0.3
-39 11 03
0.8 ± 0.1
-38 44 04
−4.6 ± 0.5
-26 02 17
-39 04 46
−2.2 ± 0.8
0.4 ± 0.3
-39 23 03
−6.0 ± 0.8
-23 04 04
−6.5 ± 0.4
-22 57 43
−6.1 ± 0.2
-23 11 06
-23 01 01
−3.1 ± 0.3
−1.7 ± 0.1
-27 50 22
−6.7 ± 0.2
-23 39 47
−3.5 ± 0.1
-22 28 29
−3.7 ± 0.7
-22 39 32
-23 00 59
−6.5 ± 0.6
−7.1 ± 0.3
-25 23 48
-22 45 22
−6.1 ± 0.1
σ RV No. of
spectra
17
2
18
3
4
4
2
5
3
2
3
2
2
3
2
2
3
3
2
3
4
2
2
2
2
2
3
2
2
2
21
9
3
4
19
18
4
13
3
4
3
2
4
[kms−1]
5.8
0.5
5.5
0.2
6.9
0.6
0.1
4.2
0.1
0.3
0.4
0.1
0.2
0.2
-
0.8
0.2
0.2
2.2
0.9
0.5
1.0
0.8
0.3
0.1
0.2
0.3
0.5
0.1
0.7
2.8
1.0
1.3
0.8
1.0
1.3
0.3
0.8
0.1
1.3
1.0
0.2
0.1
1 em=emission, abs=absorption fi=filled in
2 GW Lup, EX Lup, HK Lup : also observed by Melo et al. (2003) and found to be single
3 MN Lup: large v sin i possibly a spectroscopic binary but data insufficient
4 HIP75836 : listed in Simbad as eclipsing binary but no RV-variations observed, unlikely to be member of Lup
5 GSC 06781-01046 : this object might be a short-period SB2
6 HBC 603 (Sz77) : binary with separation of 1.′′8 ± 0.′′1 (Ghez et al. 1997) 7 GSC 06793-00569 : asymmetric cross-correlation
function, could be a visual binary
periods less than 3000 days amongst old G-dwarfs in the
solar neighbourhood of 21 ± 4%. The frequency of old and
young binaries thus is roughly the same.
Acknowledgements. We are grateful to the user support group
of ESO/La Silla. EC and JMA acknowledge financial sup-
port from INAF and Italian MIUR. This research has made
use of the SIMBAD database, operated at CDS, Strasbourg,
France. This publication makes use of data products from the
Two Micron All Sky Survey, which is a joint project of the
University of Massachusetts and the Infrared Processing and
Analysis Center/California Institute of Technology, funded by
14
Guenther et al.: PMS spectroscopic binaries
Table 20. The single stars: ρ Ophiuchi, Corona Australis
region
EW Hα1
RXJ1612.3-1909 Oph
Oph
V1002 Sco
V2503 Oph2,3
Oph
RXJ1625.3-2402 Oph
Oph
V2058 Oph
V2129 Oph2,4
Oph
HBC 268 2
Oph
HD1731485
CrA
S CrA 2,6
CrA
CrA
V709 CrA
CrA
DG CrA
CrA
V702 CrA
HBC 679
CrA
Kn Hα147
CrA
[A]
−3.9 ± 0.1
fi
−57 ± 1
f1
−91 ± 12
−22 ± 4
−42 ± 2
abs
−51 ± 6
fi
−53 ± 16
abs,fi
fi
−27 ± 13
EW LiI
[A]
≤ 1.0
spec
type
M2.5
0.48 ± 0.05
K0
0.50 ± 0.03 K6-7
0.52 ± 0.12
K5
0.45 ± 0.04
K5
0.53 ± 0.02 K3.5
0.44 ± 0.02 K2-3
G5
0.27 ± 0.03
0.20 ± 0.01
K6
0.40 ± 0.02
K0
0.36 ± 0.03
G5
0.32 ± 0.01
K5
0.48 ± 0.03
0.53 ± 0.08 M0
mK
[mag]
9.605 ± 0.019
7.494 ± 0.026
7.509 ± 0.031
8.764 ± 0.025
7.518 ± 0.024
7.207 ± 0.023
7.610 ± 0.024
7.204 ± 0.020
6.107 ± 0.021
7.713 ± 0.021
7.952 ± 0.026
8.354 ± 0.027
9.229 ± 0.030
8.446 ± 0.017
RA
(2000.0)
16 12 20.9
16 12 40.5
16 25 10.5
16 25 22.4
16 25 56.2
16 27 40.3
16 31 33.5
18 45 34.8
19 01 08.6
19 01 34.9
19 01 55.2
19 02 02.0
19 02 22.1
19 02 33.1
Dec
RV
[kms−1]
(2000.0)
-19 09 04 −1.6 ± 1.6
-18 59 28 −3.7 ± 1.0
-23 19 14 −7.1 ± 0.4
-24 02 06 −4.8 ± 0.3
-24 20 48 −4.9 ± 0.3
-24 22 03 −6.8 ± 0.6
-24 27 37 −6.3 ± 0.5
-37 50 20
0.1 ± 0.4
-36 57 20
0.9 ± 0.9
-37 00 57 −1.3 ± 0.3
-37 23 41 −1.8 ± 0.4
-37 07 44 −0.7 ± 0.1
-36 55 41 −0.1 ± 0.3
10.5 ± 4.4
-36 58 21
σ RV No. of
spectra
2
2
2
2
3
8
4
17
3
6
6
19
3
9
[kms−1]
2.2
1.4
0.6
0.4
0.7
1.7
0.9
1.6
1.6
0.7
0.9
0.5
0.5
12
1 em=emission, abs=absorption fi=filled in
2 V2503 Oph (Haro 1-4), V2129 Oph (SR 9), HBC 268 (Haro 1-16), S CrA also observed by Melo et al. (2003) and found to
be single
3 V2503 Oph (Haro 1-4) : binary with separation of 0.′′72 (Ghez et al. 1997)
4 V2129 Oph (SR 9) : binary with separation of 0.′′59 (Ghez et al. 1997)
5 HD173148 : binary with separation of 1.′′1
6 S CrA : binary with separation of 1.′′41 ± 0.′′06 (Ghez et al. 1997)
7 Kn Hα14 : this might be a spectroscopic binary
Table 14. RXJ1415.0-7822
Table 15. HK Lup
HJD
245 1261.69317
245 1622.67334
245 1737.50348
245 2019.67697
245 2026.69069
245 2031.67124
245 2098.53621
245 2372.67493
245 2373.64078
245 2384.64679
245 2710.69367
245 2717.71153
245 2723.69779
245 3096.86233
245 3104.74397
245 3105.70977
245 3122.73806
245 3137.75246
245 3168.65138
RV [km s−1]
19.7 ± 0.4
18.1 ± 1.3
17.1 ± 1.7
20.1 ± 0.3
19.0 ± 0.6
19.1 ± 0.1
16.2 ± 1.0
21.0 ± 0.2
19.3 ± 0.2
19.8 ± 0.5
19.2 ± 0.3
19.0 ± 0.5
18.9 ± 0.4
18.7 ± 0.2
20.3 ± 0.5
20.2 ± 0.4
18.6 ± 0.5
20.5 ± 0.3
18.7 ± 0.3
the National Aeronautics and Space Administration and the
National Science Foundation. The authors would also like to
thank Andrea Mehner for critically reading the text, and the
anonymous referee for improoving it.
References
Alencar, S.H.P., Melo, C.H.F., Dullemond, C.P., Andersen, J.,
Batalha, C., Vaz, L.P.R., & Mathieu, R.D. 2003, A&A, 409,
1037
HJD
245 1261.90059
245 1262.69194
245 1622.81043
245 1733.68220
245 2026.83013
245 2031.79988
245 2093.75963
245 2098.59938
245 2395.79978
245 2396.77110
245 2372.75478
245 2384.72338
245 2385.69702
245 2710.80837
245 2717.81062
245 2723.79847
245 3096.77689
245 3104.80132
245 3105.76881
245 3122.81946
245 3455.75212
RV [km s−1]
−1.3 ± 0.5
−5.0 ± 0.5
−5.3 ± 0.7
−2.5 ± 0.2
−5.2 ± 0.7
−5.7 ± 1.2
−1.4 ± 0.4
−1.9 ± 0.3
0.2 ± 0.5
4.5 ± 0.2
−0.8 ± 0.4
−2.3 ± 0.1
−0.6 ± 0.1
−0.6 ± 0.4
−2.3 ± 0.1
−0.4 ± 0.1
−1.8 ± 0.3
−2.3 ± 2.9
−8.6 ± 0.4
−8.7 ± 0.2
1.6 ± 0.8
Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P.H. 1998,
A&A, 337, 403
Bertout, C., Robichon, N., & Arenou, F. 1999, A&A, 352, 574
Beuzit, J.L., Hubin, N., Gendron, E., Demailly, L., Gigan, P.,
Lacombe, F., Chazallet, F., Rabaud, D., & Rousset, G.
1994, in: Adaptive Optics in Astronomy, eds. M.A. Ealey
& F. Merkle, Proc. SPIE, 2201 955
Guenther et al.: PMS spectroscopic binaries
15
Table 16. GSC 06793-01406
HJD
245 1625.80168
245 1737.70415
245 2089.73677
245 2093.78125
245 2384.81131
245 2717.88212
245 2723.87107
245 3096.84966
245 3104.85577
245 3105.83868
245 3122.87716
245 3137.88922
245 3168.74239
RV [km s−1]
−6.9 ± 0.6
−7.4 ± 0.6
−7.3 ± 0.4
−5.8 ± 0.7
−5.5 ± 0.1
−7.9 ± 0.2
−6.8 ± 0.1
−5.7 ± 0.3
−5.6 ± 0.1
−7.1 ± 0.1
−7.0 ± 0.2
−8.0 ± 0.2
−6.2 ± 0.1
Table 17. GSC 06793-00994
HJD
245 1262.70144
245 1737.69110
245 2089.51975
245 2093.68633
245 2396.80296
245 2372.82468
245 2373.69694
245 2384.79676
245 2385.73404
245 2710.87369
245 2717.83773
245 2723.83505
245 3096.79173
245 3104.83363
245 3105.80278
245 3122.84741
245 3137.84760
245 3168.72226
RV [km s−1]
−2.8 ± 0.1
−4.2 ± 0.5
−3.9 ± 0.1
−2.6 ± 0.3
−0.9 ± 0.1
−0.7 ± 0.2
−4.4 ± 0.3
−0.9 ± 0.1
−5.0 ± 0.1
−5.4 ± 0.1
−3.4 ± 0.1
−3.6 ± 0.1
−2.2 ± 0.1
−3.6 ± 0.1
−2.5 ± 0.1
−3.6 ± 0.1
−3.0 ± 0.1
−3.0 ± 0.1
Boden, A.F., Sargent, A.I., Akeson, R.L., Carpenter, J.M.,
Torres, G., Latham, D.W., Soderblom, D.R., Nelan, E.,
Franz, O.G., & Wasserman, L.H. 2005, ApJ, 635, 442
Brault, J.W. 1978, in: Proceedings of JOSO Workshop, G.
Godoli, G. Noci, & A. Righini, Arcetri: Osser. Mem.
Osserv. Astrofis., Number 106, 33
Brandner, W., Alcal´a, J.M., Kunkel, M., Moneti, A., &
Zinnecker, H. 1996, A&A, 307, 121
Burrows, A., et al. 1997, ApJ, 491, 856
Casey, B.W., Mathieu, R.D., Vaz, L.P.R., Andersen, J., &
Suntzeff, N.B. 1998, AJ, 115, 1617
D'Antona, F., & Mazzitelli, I. 1994, ApJS, 90, 467
de Zeeuw, P.T., Hoogerwerf, R., de Bruijne, J.H.J., Brown,
A.G.A., & Blaauw, A., 1999, AJ, 117, 354
Duquennoy, A., & Mayor, M. 1991, A&A248, 485
Esposito, M., Covino, E., Alcal´a, J.M., Guenther, E.W., &
Schisano, E., supmitted to MNRAS
Forestini, M. 1994, A&A, 285, 473
Franco, G.A.P. 2002, MNRAS, 331, 474
Ghez, A.M., Neugebauer, G., & Matthews, K., 1993, AJ, 106,
2005
Ghez, A.M., McCarthy, D.W., Patience, J.L., & Beck, T.L.
1997, ApJ, 481, 378
Guenther, E.W., Torres, G., Batalha, N., Joergens, V.,
Neuhauser, R., Vijapurkar, J., & Mundt, R. 2001, A&A,
366, 965
Guenther, E.W., Covino, E., Alcal´a, J.M., Esposito, M., &
Mundt, R. 2005, A&A, 433, 629
Hatzes, A. P., Cochran, W. D., Endl, M., Guenther, E. W.,
Saar, S. H., Walker, G. A. H., Yang, S., Hartmann, M.,
Esposito, M., Paulson, D. B., Dollinger, M. P. 2006, A&A,
457, 335
Hatzes, Artie P., Cochran, William D., McArthur, Barbara,
Baliunas, Sallie L., Walker, Gordon A. H., Campbell,
Bruce, Irwin, Alan W., Yang, Stephenson, Kurster, Martin,
Endl, Michael, Els, Sebastian, Butler, R. Paul, Marcy,
Geoffrey W. 2000, ApJ, 544, L145
Henning, T., Pfau, W., Zinnecker, H., & Prusti, T. 1993, A&A,
276, 129
Hogeveen, S.J. 1992, Astrophysics and Space Science, 193, no.
1, 1992, pp. 29-46
Hughes, J., Hartigan, P., & Clampitt, L. 1993, AJ, 105, 571
Joergens, V., Guenther, E., Neuhauser, R., Fern´andez, M., &
Vijapurkar, J. 2001, A&A, 373, 966
Knacke, R. F., Strom, K. M., Strom, S. E., Young, E. T., &
Kunkel, W., 1973, ApJ, 179, 847
Knude, J., & Høg, E. 1998, A&A, 338, 897
Kohler, R. 2001, AJ122, 3325
Konig, B., Guenther, E.W., Woitas, J., & Hatzes, A.P. 2005,
A&A, 435, 215
Lovis, C., Mayor, M., Bouchy, F., Pepe, F., Queloz, D., Santos,
N. C., Udry, S., Benz, W., Bertaux, J.-L., Mordasini, C.,
Sivan, J.-P. 2005, A&A, 437, 1121
Marraco, H.G., & Rydgren, A.E. 1981, AJ, 86, 62
Mathieu, R.D. 1994, ARA&A, 32, 465
Melo, C.H.F. 2003, A&A, 410, 269
Montalb´an J., D'Antona, F., Kupka, F., & Heiter, U. 2004,
A&A, 416, 1081
Murdoch, K.A., Hearnshaw, J. B., & Clark, M., 1993, ApJ,
413, 349
Nakajima, Y., Tamura, M., Oasa, Y., & Nakajima, T. 2000,
AJ, 119, 873
Chabrier, G., & Baraffe, I. 2000, ARA&A, 38, 337
Chini, G. 1981, A&A, 99, 346
Covino, E., Alcala, J. M., Allain, S., Bouvier, J., Terranegra,
L., & Krautter, J. 1997, A&A, 328, 187
Padget, D.L., 1996, ApJ, 471, 847
Palla, F., & Stahler, S.W. 1992, ApJ, 392, 667
Palla, F., & Stahler, S.W. 1999, ApJ, 525, 772
Prato, L., Simon, M., Mazeh, T., McLean, I.S., Norman, D.,
Covino, E., Frasca, A., Alcal´a, J.M., Paladino, R., & Sterzik,
& Zucker, S. 2002, ApJ, 569, 863
M.F. 2004, A&A, 427, 637
Covino, E. et al. in preparation
Cutri, R. M., et al. 2003, VizieR Online Data Catalog: II/246.
Originally published in: University of Massachusetts and
Infrared Processing and Analysis Center, (IPAC/California
Institute of Technology) (2003)
D'Antona, F., Ventura, P., & Mazzitelli, I. 2000, ApJ, 543, L77
Quist, C.F., & Lindegren, L., 2000, A&A, 361, 770
Paulson, D. B., Saar, S.H., Cochran, W.D., Hatzes, A. P. 2002,
AJ, 124, 572
Preibisch, T., Guenther, E., Zinnecker, H., Sterzik, M., Frink,
S., Roeser, S. 1998, A&A, 333, 619
Preibisch, T., & Zinnecker, H. 1999, AJ, 117, 2318
Reipurth, B., & Zinnecker, H. 1993, A&A, 278, 81
16
Guenther et al.: PMS spectroscopic binaries
Reipurth, B., Lindgren, H., Mayor, M., Mermilliod, J.-C., &
Cramer, N. 2002, AJ, 124, 2813
Saar, S. H., Butler, R.P., & Marcy, G. W. 1998, ApJ, 498, L153
Sartori, M.J., Lepine, J.R.D., & Dias, W.S. 2003, A&A, 404,
913
Siess, L., Forestini, M., & Dougados, C. 1997, A&A, 324, 556
Silverstone, M.D., Meyer, M.R., Mamajek, E.E., Hines, D.C.,
Hillenbrand, L.A., Najita, J., Pascucci, I., Bouwman, J.,
Kim, J.S., Carpenter, J.M., Stauffer, J.R., Backman, D.E.,
Moro-Martin, A., Henning, Th., Wolf, S., Brooke, T.Y. &
Padgett, D.L., 2006, ApJ639, (astro-ph/0511250)
Simon, M., Dutrey, A., & Guilloteau, S. 2000, ApJ, 545, 1034
Simon, M., & Prato, L. 2004, ApJ, 613, L69
Skrutskie, M. F., Cutri, R. M., Stiening, R. et al., 2006, AJ,
131, 1163
Song, I., Bessell, M.S., Zuckerman, B. 2002, A&A, 385, 862
Stassun, K.G., Mathieu, R.D., Vaz, L.P.R., Stroud, N., & Vrba,
F.J. 2004, ApJS, 151, 357
Steffen, A.T., Mathieu, R.D., Lattanzi, M.G., Latham, D.W.,
Mazeh, T., Prato, L., Simon, M., Zinnecker, H., & Loreggia,
D. 2001, AJ, 122, 997
Stempels, H. C., Gahm, G.F., & Petrov, P.P. 2007, A&A, 461,
253
Swenson, F.J., Faulkner, J., Iglesias, C.A., Rogers, F.J., &
Alexander, D.R. 1994, ApJ, 422, L79
Takami, M., Bailey, J., & Chrysostomou, A., 2003, A&A, 397,
675-691
Teixeira, R., Ducourant, C., Sartori, M.J., Camargo, J.I.B.,
P´eri´e, J.P., L´epine, J.R.D., & Benevides-Soares, P., 2000,
A&A, 361, 1143
Torres, C.A.O., da Silva, L., Quast, G. R., de la Reza, R., &
Jilinski, E. 2000, AJ, 120, 1410
Torres, G., Guenther, E.W., Marschall, L.A., Neuhauser, R.,
Latham, D.W., & Stefanik, R.P. 2003, AJ, 125, 825
Tout, C.A., Livio, M., & Bonnell, I.A. 1999, MNRAS, 310, 360
Wallace, L., Hinkle, K., & Livingston, W., 1998, An atlas of
the spectrum of the solar photosphere from 13,500 to 28,000
cm−1 (3570 to 7405 A), Publisher: Tucson, AZ: National
Optical Astronomy Observatories.
Webb, R.A., Zuckerman, B., Platais, I., Patience, J., White,
R.J., Schwartz, M.J., & McCarthy, C. 1999, ApJ, 512, L63
Weinberger, A.J., Becklin, E.E., Zuckerman, B., & Song, I.
2004, AJ, 127, 2246
Whittet, D.C.B. 1974, MNRAS, 168, 371
Whittet, D.C.B., Prusti, T., Franco, G.A.P., Gerakines, P.A.,
Kilkenny, D., Larson, K.A., & Wesselius, P.R. 1997, A&A,
327, 1194
Wichmann, R., Bastian, U., Krautter, J., Jankovics, I., &
Ruciski, S.M. 1998, MNRAS, 301, L39
Wuchterl, G., Tscharnuter, W.M. 2003, A&A, 398, 1081
List of Objects
'RXJ 1608.6-3922' on page 2
'TY CrA' on page 2
'RXJ 0529.4+0041' on page 2
'NTT 045251+3016' on page 2
'HD 98800 B' on page 2
'RXJ1603.8-3938' on page 2
'EK Dra' on page 2
'HR3862' on page 4
'HR6748' on page 4
'HR 5777' on page 4
'HR5568' on page 4
'HR6056' on page 4
'HR 5777' on page 4
'HR 5777' on page 4
'HR 5777' on page 4
'HR 5777' on page 4
'HR 5777' on page 4
'HR 5777' on page 4
'HR 5777' on page 4
'HR 5777' on page 4
'CS Cha' on page 4
'RXJ1220.6-7539' on page 4
'MO Lup' on page 4
'RXJ1534.1-3916' on page 4
'RXJ1559.2-3814' on page 4
'GSC 06209-00735' on page 4
'GSC 06213-00306' on page 4
'BS Indi' on page 4
'HR 5777' on page 5
'HIP50796' on page 5
'HD97131' on page 5
'NTTS160814-1857' on page 5
'Haro 1-14c' on page 5
'NTTS162819-2423s' on page 5
'VW Cha' on page 5
'GSC 06793-00569' on page 5
'CS Cha' on page 5
'CS Cha' on page 5
'CS Cha' on page 5
'RXJ1534.1-3916' on page 5
'RXJ1534.1-3916' on page 5
'RXJ1220.6-7539' on page 6
'RXJ1559.2-3814' on page 6
'RXJ1534.1-3916' on page 6
'GSC 06209-00735' on page 7
'GSC 06213-00306' on page 7
'GSC 06213-00306' on page 7
'RXJ 1559.2-3814' on page 8
'VW Cha' on page 8
'VW Cha' on page 8
'GSC 06209-00735' on page 9
'VW Cha' on page 9
'GSC 06213-00306' on page 10
'GSC 06793-00569' on page 10
'VW Cha' on page 10
'EK Dra' on page 11
'β Gem' on page 11
'EK Dra' on page 11
'β Gem' on page 11
'RU Lup' on page 11
'RXJ1415.0-7822' on page 11
'HK Lup' on page 11
'GSC 06793-01406' on page 11
'GSC 06793-00994' on page 11
'GSC 06781-01046' on page 11
'MN Lup' on page 11
'GSC 06781-01046' on page 11
'RXJ1108.8-7519' on page 11
'RXJ1603.8-3938' on page 11
Guenther et al.: PMS spectroscopic binaries
17
|
astro-ph/0412706 | 1 | 0412 | 2004-12-31T13:16:01 | NIR Spectroscopy of Luminous Infrared Galaxies and the Hydrogen Recombination Photon Deficit | [
"astro-ph"
] | We report on near-infrared medium-resolution spectroscopy of a sample of luminous and ultra luminous infrared galaxies (LIRGs-ULIRGs), carried out with SOFI at the ESO 3.5m New Technology Telescope. Because of wavelength dependence of the attenuation, the detection of the Pa_alfa or Br_gamma line in the Ks band should provide relevant constraints on SFR and the contribution of an AGN. We find, however, that the intensities of the Pa_alfa and Br_gamma lines, even corrected for slit losses, are on average only 10% and 40%, respectively, of that expected from a normal starburst of similar bolometric luminosity. The corresponding star formation rates, after correcting for the attenuation derived from the NIR-optical emission line ratios, are 14% and 60% of that expected if the far infrared luminosity were entirely powered by the starburst. This confirms the existence of a recombination photon deficit, particularly in the case of the Pa_alfa line, already found in the Br_gamma line in other infrared galaxies of similar luminosity. In discussing the possible causes of the discrepancy, we find unlikely that it is due to the presence of an AGN, though two objects show evidence of broadening of the Pa_alfa line and of the presence of coronal line emission. In fact, from our own observations and data collected from the literature we argue that the studied galaxies appear to be predominantly powered by a nuclear starburst. Two scenarios compatible with the present data are that either there exists a highly attenuated nuclear star forming region, and/or that a significant fraction of the ionizing photons are absorbed by dust within the HII regions. We suggest that observations in the Br_alpha spectral region could constitute a powerful tool to disentangle these two possibilities. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(DOI: will be inserted by hand later)
August 8, 2018
NIR Spectroscopy of Luminous Infrared Galaxies and the
Hydrogen Recombination Photon Deficit ⋆
J.R. Vald´es1
,
2, S. Berta3, A. Bressan1, A. Franceschini3, D. Rigopoulou4, and G. Rodighiero3
1 INAF, Osservatorio Astronomico di Padova, vicolo dell'Osservatorio 5, 35122 Padova, Italy
e-mail: [email protected],[email protected]
2 Instituto Nacional de Astrof´ısica, Optica y Electr´onica, Apdos. Postales 51 y 216, C.P. 72000 Puebla, Pue.,
M´exico
e-mail: [email protected]
3 Dipartimento di Astronomia, vicolo dell'Osservatorio 5, 35122 Padova, Italy
e-mail: [email protected],[email protected],[email protected]
4 Department of Physics (Astrophysics), University of Oxford, Denys Wilkinson Building, 1 Keble Road, Oxford
OX1 3RH, UK
e-mail: [email protected]
Received date; accepted date
Abstract. We report on near-infrared medium-resolution spectroscopy of a sample of luminous and ultra-luminous
infrared galaxies (LIRGs-ULIRGs), carried out with SOFI at the ESO 3.5m New Technology Telescope. Because
of wavelength dependence of the attenuation, the detection of the Paα or Brγ line in the Ks band should provide
relevant constraints on SFR and the contribution of an AGN.
We find, however, that the intensities of the Paα and Brγ lines, even corrected for slit losses, are on average
only 10% and 40%, respectively, of that expected from a normal starburst of similar bolometric luminosity. The
corresponding star formation rates, after correcting for the attenuation derived from the NIR-optical emission line
ratios, are 14% and 60% of that expected if the far infrared luminosity were entirely powered by the starburst.
This confirms the existence of a recombination photon deficit, particularly in the case of the Paα line, already
found in the Brγ line in other infrared galaxies of similar luminosity.
In discussing the possible causes of the discrepancy, we find unlikely that it is due to the presence of an AGN,
though two objects show evidence of broadening of the Paα line and of the presence of coronal line emission. In
fact, from our own observations and data collected from the literature we argue that the studied galaxies appear
to be predominantly powered by a nuclear starburst.
Two scenarios compatible with the present data are that either there exists a highly attenuated nuclear star
forming region, and/or that a significant fraction (≃80% ) of the ionizing photons are absorbed by dust within
the HII regions. We suggest that observations in the Brα spectral region could constitute a powerful tool to
disentangle these two possibilities.
Key words. ISM: dust extinction -- Galaxies:starburst -- Infrared:galaxies
4
0
0
2
c
e
D
1
3
1
v
6
0
7
2
1
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1. Introduction
The advent of the Infrared Space Observatory (ISO) com-
bined with the availability of new ground based facilities
such as SCUBA on the JCMT have resulted in the discov-
ery of numerous distant galaxies with enhanced infrared
(IR) emission (e.g. Elbaz et al. 1999, Smail et al. 2000,
Barger et al. 2000, Gear et al. 2000, Franceschini et al.
2001, Smail et al. 2003, Ivison et al. 2002).
Send offprint requests to: A.Bressan
⋆ based on observations collected at the European Southern
Observatory, Chile, ESO No. 67.A-0593, 71.A-0707.
It is not yet clear what mechanisms power the bolo-
metric luminosity of these luminous (LIRGs) and ultra
(ULIRGs) luminous infrared galaxies, with bolometric lu-
minosities exceeding 1011L⊙ and 1012L⊙, respectively.
Perhaps, the tightest constraint on the nature of the in-
frared luminosity is that they mostly obey the FIR/Radio
correlation of normal star forming galaxies (Sanders &
Mirabel 1996). However the presence of broad emission
line components, a certain degree of polarization, power-
law near infrared colours, warm far-infrared (FIR) spectra,
high radio brightness temperature in the milli-arcsecond
central structure and, finally, hard X-ray emission are of-
2
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
ten detected, suggesting that a significant contribution to
the luminosity may come from mass accretion onto a com-
pact central object.
These galaxies are often found to be interacting sys-
tems, indicating that the two phenomena may be triggered
at once by dynamical interaction (e.g. Rigopoulou et al.
1999).
Optical studies have been performed in order to estab-
lish the nature of these objects (e.g. Veilleux et al. 1999),
but it is difficult to disentangle between AGN and star for-
mation processes, because the bulk of their luminosity is
strongly attenuated by dust and the re-emitted FIR spec-
trum keeps little memory of the primary power source.
Even in objects without clear signatures of the pres-
ence of AGN, there are significant discrepancies between
different observables. For example Poggianti, Bressan &
Franceschini (2001),
in analyzing a sample of Infrared
Galaxies with log(LIR/L⊙)≥11.5, concluded that the Star
Formation Rate (SFR) detected from optical diagnostic
tools (even corrected for attenuation) amounts to about
1/3 of that inferred from the FIR. However, if such a large
fraction of stellar activity is hidden by dust, one may won-
der whether also the nuclear activity is invisible at optical
wavelengths.
A natural way to avoid strong dust obscuration is to
look in the near infrared where, being the attenuation only
a fraction (≃1/7) of that in the optical, one may hope to
obtain an unbiased estimate of the SFR and at the same
time to search for the possible presence of an obscured
AGN by means of broadening of permitted emission line
and/or presence of coronal lines.
With these purposes we begun a medium resolution
NIR spectroscopic study with SOFI at the ESO 3.5m New
Technology Telescope, of a sample of ULIRGs selected
from Genzel et al. (1998) and Rigopoulou et al. (1999)
samples. In this paper we present and discuss these new
observations.
The plan of the paper is as follows. In Sections 2 we
present the observations and data reduction, with an anal-
ysis of the emission line properties of target galaxies pre-
sented in subsections 2.1 and 2.2. All the process to calcu-
late the Star Formation rates are discussed in Section 3.
In particular, the technique we adopted to calculate the
slit losses and to evaluate the extinction are presented in
subsections 3.1 and 3.2. In Section 4 we show that, even
correcting for extinction and losses from the slit, there is
a large Paα/IR deficit, and we discuss its possible origin.
Our conclusions are summarized in Section 5. We assume
H0=65 [km/s/Mpc], ΩM =0.3, ΩΛ=0.7 throughout.
2. Observations and Data Reduction
The selected galaxies were drawn from the IRAS 1.2Jy
sample (Fisher et al. 1995), and from Genzel et al.
(1998) and Rigopoulou et al. (1999). The sample galaxies
cover the interval of solar luminosities 11.45≤log(LIR/L
⊙)≤12.73. The IR and radio properties of our target galax-
ies are listed in Table 1. The penultimate column indicates
the FIR/Radio correlation parameter at 1.4GHz, (see eq.
1 below), which for star forming galaxies is known to be
q≃2.35±0.15 over a wide range of infrared luminosities
(e.g. Sanders & Mirabel 1996).
Medium-resolution Ks-band NIR spectroscopy of the
galaxies was performed with the SOFI infrared spectrom-
eter (Moorwood et al. 1998) at the Nashmyth A focus of
the ESO 3.5m New Technology Telescope (NTT), during
August 2001 and August 2003. During the 2003 run, 5
minutes of Ks-band imaging were obtained as well.
The instrumental set consisted of the medium resolu-
tion grism (pixel scale = 0.292 arcsec), with a dispersion
of 4.62 A/pixel and a spectral coverage between 2.0 µm
and 2.3 µm. The Ks filter works as order sorting filter.
We adopted 1 arcsec slit, providing a spectral resolution
of R ≡ ∆ λ/λ ≈ 2000 (∆v ≈ 150 km s−1) at 2 µm.
Data were acquired by placing the galaxy at two differ-
ent positions along the slit, through the standard nodding
technique; the nod throw was set to 60 arcsec.
In some cases, given the peculiar structure of the
galaxy, the slit has been oriented with two different posi-
tion angles (see Table 2).
The standard reduction pipeline for long-slit NIR
spectra was performed by using routines of the Image
Reduction and Analysis Facility (IRAF1). The spectra
were flat-field corrected in the usual manner and sky was
subtracted through the nod on slit procedure.
Wavelength calibration and slit-curvature correction
was obtained with the Xenon/ Neon internal lamp and
checked on sky IR emission lines.
Corrections for telluric features and flux calibration
have been performed by observing nearly featureless hot
B and/or solar-like telluric and spectro- photometric stan-
dard stars, from the Hipparcos catalogue. Atmospheric
transmission variations were compensated by observing
the standard stars contiguously, at similar (as near as pos-
sible to target galaxies) air masses. The adopted procedure
consists in dividing the raw galaxies spectra by that of the
standard star and then multiplying by its intrinsic emit-
ted spectrum. The latter has been obtained by normalizing
Pickles (1998) stellar templates of the proper spectral type
to the V-to-K observed magnitudes of the standard stars.
In the case of not optimal seeing conditions, we corrected
the raw spectra of the standard stars for slit-losses, before
applying flux calibration.
The exposure time, the position angle (PA) of the slit,
the spectral type of the atmospheric calibrators and their
air masses with respect to those of the observed galaxies
are given in Table 2.
1 The package IRAF is distributed by the National Optical
Astronomy Observatory which is operated by the Association
of Universities for Research in Astronomy, Inc., under cooper-
ative agreement with the National Science Foundation.
6
6
5.5
5.5
5
5
4.5
4.5
4
4
1.2
1.2
1
1
0.8
0.8
0.6
0.6
0.4
0.4
2
2
0.3
0.3
0.25
0.25
0.2
0.2
0.15
0.15
0.1
0.1
0.05
0.05
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
3
IRAS00085-1223
IRAS00085-1223
2.05
2.05
2.1
2.1
2.15
2.15
2.2
2.2
IRAS01077-1707
IRAS01077-1707
0.5
0.5
0.4
0.4
0.3
0.3
0.2
0.2
0.3
0.3
0.2
0.2
IRAS00188-0856
IRAS00188-0856
IRAS00582-0258
IRAS00582-0258
0.16
0.16
0.12
0.12
0.08
0.08
1.85
1.85
1.9
1.9
1.95
1.95
2
2
0.04
0.04
1.86
1.86
1.88
1.88
1.9
1.9
1.92
1.92
1.94
1.94
1.96
1.96
1.98
1.98
2
2
IRAS06206-6315
IRAS06206-6315
0.4
0.4
0.3
0.3
0.2
0.2
0.1
0.1
IRAS19335-3632
IRAS19335-3632
2.05
2.05
2.1
2.1
2.15
2.15
2.2
2.2
0.1
0.1
1.85
1.85
1.9
1.9
1.95
1.95
2
2
2.05
2.05
2.1
2.1
0
0
1.85
1.85
1.9
1.9
1.95
1.95
2
2
2.05
2.05
2.1
2.1
IRAS22491-1808
IRAS22491-1808
1
1
0.8
0.8
0.6
0.6
0.4
0.4
IRAS23128-5919
IRAS23128-5919
0.3
0.3
0.2
0.2
0.1
0.1
IRAS23230-6926
IRAS23230-6926
1.9
1.9
1.95
1.95
2
2
2.05
2.05
2.1
2.1
2.15
2.15
2
2
2.05
2.05
2.1
2.1
2.15
2.15
1.9
1.9
1.95
1.95
2
2
2.05
2.05
IRAS23389-6139
IRAS23389-6139
0.2
0.2
0.1
0.1
1.85
1.85
1.9
1.9
1.95
1.95
2
2
Fig. 1. Spectra of the sample galaxies showing emission in only one nucleus. No slit-losses, nor extinction correction have been
applied. The identified emission features are highlighted. Fluxes are in units of [10−15ergs−1cm−2A−1].
2.1. Emission line properties
NIR spectra of galaxies in our sample, presented in Figures
1 and 2, are characterized by strong Paα or Brγ emission,
depending on their redshift. In the Paα region we also
detect weak H2 lines from higher order vibrational transi-
tions (H2 1-0S(5) at 1.835µm, H2 1-0S(3) at 1.957µm, and
H2 1-0S(2) at 2.033µm). Almost all these galaxies show a
contamination of the Paα blue wing by the HeI emission
at 1.868µm, and in some of them the weak Brδ emission at
1.945µm is detected. Galaxies observed in the Brγ region,
show H2 lines from lower order vibrational transitions (H2
1-0S(3) at 1.957µm, H 2 1-0S(2) at 2.033 µm, H2 1-0S(1)
at 2.121µm, and H 2 1-0S(0) at 2.223µm). It is to note the
relatively strong HeI emission at 2.058µm in the case of
IRAS 23128-5919.
4
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
0.5
0.5
0.4
0.4
0.3
0.3
0.2
0.2
0.1
0.1
IRAS02411+0354NE
IRAS02411+0354NE
0.15
0.15
0.1
0.1
0.05
0.05
IRAS02411+0354SW
IRAS02411+0354SW
0
0
1.8
1.8
1.85
1.85
1.9
1.9
1.95
1.95
2
2
0
0
1.8
1.8
1.85
1.85
1.9
1.9
1.95
1.95
2
2
IRAS20100-4156S
IRAS20100-4156S
1.85
1.85
1.9
1.9
1.95
1.95
2
2
IRAS22206-2715SW
IRAS22206-2715SW
0.3
0.3
0.2
0.2
0.1
0.1
0
0
0.1
0.1
0.08
0.08
0.06
0.06
0.04
0.04
0.08
0.08
0.06
0.06
0.04
0.04
0.02
0.02
0
0
IRAS20100-4156N
IRAS20100-4156N
0.2
0.2
0.1
0.1
IRAS22206-2715NE
IRAS22206-2715NE
1.85
1.85
1.9
1.9
1.95
1.95
2
2
1.85
1.85
1.9
1.9
1.95
1.95
2
2
1.85
1.85
1.9
1.9
1.95
1.95
2
2
Fig. 2. Spectra of the sample galaxies showing emission in two nuclei. No slit-losses, nor extinction correction have been applied.
The identified emission features are highlighted. Fluxes are in units of [10−15erg s−1cm −2A−1].
Table 1. Infrared and radio properties of the observed galaxies
Object
z
D
S12µm S25µm S60µm S100µm S25/S60
IRAS00085-1223
IRAS00188-0856
IRAS00582-0258
IRAS01077-1707
IRAS02411+0354
IRAS06206-6315
IRAS19335-3632
IRAS20100-4156
IRAS22206-2715
IRAS22491-1808
IRAS23128-5919
IRAS23230-6926
IRAS23389-6139
0.0198
0.1284
0.0874
0.0334
0.1436
0.0924
0.0821
0.1295
0.1314
0.0777
0.0446
0.1063
0.0927
[Mpc]
80.3
648.6
373.7
158.4
638.1
455.9
349.8
655.0
579.1
379.7
184.7
529.2
457.5
[Jy]
0.395
0.117
0.110
0.301
0.085
0.069
0.107
0.135
0.096
0.119
0.250
0.058
0.063
[Jy]
2.372
0.372
0.341
0.846
0.224
0.294
0.154
0.343
0.160
0.549
1.590
0.295
0.244
[Jy]
16.62
2.59
1.21
6.48
1.37
3.96
1.19
5.23
1.75
5.44
10.80
3.74
3.63
[Jy]
16.97
3.40
0.99
10.4
1.94
4.58
1.78
5.16
2.33
4.45
10.99
3.42
4.26
0.143
0.144
0.282
0.130
0.164
0.074
0.129
0.065
0.091
0.101
0.147
0.079
0.067
SIR
[W m2]
1.39×10−12
2.44×10−13
1.33×10−13
6.40×10−13
1.40×10−13
3.10×10−13
1.27×10−13
4.00×10−13
1.61×10−13
4.13×10−13
9.07×10−13
2.77×10−13
2.83×10−13
SRadio
[mJy]
66.9b
15.7b
10.1b
43.8b
6.6b
21.9a
7.1b
20.3c
6.3b
5.9b
37.3a
32.0a
163.9a
qd
2.48
2.36
2.14
2.32
2.44
2.53
2.36
2.49
2.56
3.02
2.72
2.31
1.62
L/C e
7.7µm
1.53
3.17
3.69
1.92
2.85
1.50
1.33
a) 843MHz flux from Mauch et al (2003); b) 1.4GHzflux from NRAO VLA Sky Survey (Condon et al 1998); c) Condon et al
(1996)
d) FIR/Radio correlation parameter (Sanders & Mirabel 1996). 843MHz fluxes were converted to 1.4GHz fluxes by assuming
F(ν) ∝ ν −0.8
e) L/C from Lutz et al. 1998
We find a good correlation between the Paα and IR
luminosity, as shown in Figure 4. The dashed line in the
Figure is the zero-intercept linear correlation we obtain,
log(LP aα/LIR) = -4.69. IRAS 02411+0354 has been ex-
cluded from the analysis because it falls significantly off
the correlation. In the case of Brγ luminosity, we have
only three galaxies, but our values fall on top of the re-
lation found by Goldader et al. (1997). The dashed line
is a zero-intercept linear correlation with slope quoted by
Goldader et al. (1997), log(LBrγ/LIR) = -4.96.
The spatial distribution of the Paα emission in IRAS
00582-0258, IRAS 23230-6926 and IRAS 23389-6139 ap-
pears compact. The peak of the Gaussian profile of the
line emission corresponds to the peak obtained from the
continuum emission alone, coming from the nucleus of the
galaxies. In contrast, IRAS 22491-1808 shows a broader
spatial distribution in Paα emission, with some minor
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
5
Table 2. Observational Parameters
IRAS name
00085-1223
00188-0856
00582-0258
01077-1707
02411+0354
06206-6315
19335-3632
20100-4156
22206-2715
22491-1808
23128-5919
23230-6926
23389-6139
Integ.
(seg)
3600
7200
3600
3600
480
6600
6600
2400
5400
7500
8400
3600
3600
7200
4800
4500
480
P.A. Atmos. Gal./Std.
(deg)
Air Mass
1.11/1.11
38
1.13/1.46
0
1.18/1.17
55
67
1.11/1.16
1.22/1.25
30
1.22/1.07
-54
1.19/1.07
90
1.57/1.69
145
43
1.19/1.47
1.32/1.05
25
1.11/1.12
-56
1.11/1.04
-32
61
1.37/1.04
1.02/1.08
74
1.32/1.10
35
1.45/1.51
158
8
1.20/1.03
Calib.
B9V
F8V
G0V
B9V
F6V
G0V
G0V
G0V
B9V
G2V
B9V
G0V
G0V
G3V
B4V
F7V
F5V
peaks superposed on a broader single Gaussian profile
along the slit.
Six galaxies, IRAS 00188-0856, IRAS 02411+0354,
IRAS 19335-3632, IRAS 20100-4156, IRAS 22206-2715,
and IRAS 22491-1808 show double profiles, corresponding
to the double nuclei observed in the corresponding NIR
images. Finally, IRAS 06206-6315 is best fitted with three
Gaussian profiles. This complexity is associated with the
bright regions seen in the H and I images (Bushouse et al.
2002).
In the case of galaxies observed in the Brγ spectral do-
main, IRAS00085-1223 and IRAS23128-5919 show a sin-
gle Gaussian profile, while IRAS 01077-1707 is best fitted
with three Gaussian profiles.
The observed emission line properties of target galaxies
are shown in Table 3. Only for those galaxies with suffi-
cient S/N have we considered the emission properties of
multiple components separately while, in all other cases,
we have folded the separate contribution together. The
FWHM velocities have been corrected by instrumental re-
sponse after subtracting in quadrature the instrumental
FWHM.
2.2. Detection of AGN signatures
Typical signatures of central non thermal activity are the
presence of broad emission line components and/or high
excitation lines.
None of the objects observed in the Brγ region show a
broad component of the Brγ line. In the case of the Paα
galaxies, we have attempted to isolate a broad compo-
nent of the Paα emission, by forcing the fit of the shape of
the line with two Gaussians. With this procedure, in IRAS
00188-0856 and IRAS 00582-0258 we could detect a broad
component of FWHM ≃ 2339 and 2210 km/s, respectively.
Figure 3 shows the results of the multi Gaussian fitting for
these galaxies. In the case of IRAS 00188-0856 we have
also taken into account the contribution of the HeI line at
1.8689µm to the general shape of the line. The broad com-
ponent corresponds to about 42% and 80% of the Paα flux
in the narrow component for IRAS 00188-0856 and IRAS
00582-0258, respectively. The Paα intensities reported in
Table 3 correspond only to the narrow component.
Another AGN diagnostic in the NIR, is the presence of
the high excitation line [SiVI]λ1.9628µm. This diagnostic
is particularly useful in Sy2 galaxies, that lack broad-line
components. The [SiVI]λ1.9628µm line falls slightly long
wards of the H2 1-0S(3) 1.957 µm line, causing an asymme-
try in medium resolution spectra (e.g. Vanzi et al. 2002).
While our spectral resolution allow us to disentangle these
two lines, the S/N is not sufficient to draw firm conclu-
sions. However, [SiVI]λ1.9628µm seems to be present in
the spectra of IRAS 00188-0856 and IRAS 00582-0258, the
same galaxies for which we have hints of a broad emission
component at Paα. In all other cases there is no evidence
of this coronal line.
3. The star formation rates.
In this section we compare the star formation rate derived
from the Paα or Brγ emission lines, with that deduced
from the IR, assuming that both originate in an ongoing
vigorous starburst.
In order to perform a proper analysis of the star for-
mation rate from the emission lines we need to correct the
measured fluxes for slit losses and for dust extinction.
3.1. Slit loss corrections
When comparing different observations, it is important
to evaluate any discrepancy caused by the different aper-
tures. If the source has a certain degree of symmetry,
it is possible to extrapolate the mono dimensional spa-
tial profile observed along the slit, to a two-dimensional
surface brightness distribution, and to estimate the flux
that would be recevied by a given aperture. We define as
"slit losses" the ratio between the flux corresponding to a
given aperture and the flux received within our rectangu-
lar aperture.
In our galaxies, the two-dimensional flux distribution
was obtained by rotating around the nucleus the best fit
Gaussian profile of the spatial distribution of the Paα or
Brγ line. If the slit covers more than one nucleus, dif-
ferent Gaussian profiles were fitted for each nucleus. As
an example, Figure 5 shows the case of IRAS20100-4156,
at two position angles P.A.=-24 and 56 respectively. Slit
loss coefficients,CHα and CIR, needed to compare our Paα
fluxes to Hα and IR fluxes taken from the literature are
reported in Table 4.
3.2. Attenuation
Correction for attenuation can be derived by comparing
the observed hydrogen line emission intensity ratios, Robs,
6
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
Fig. 3. Possible detection of a broad emission component in the IRAS 00188-0856 (left panel) and IRAS 00582-0258
(right panel). The width of the broad component is FWHM≃2339 km/s for IRAS 00188-0856, and FWHM≃2210 km/s
for IRAS 00582-0258.
Fig. 5. Example of calculation of slit losses for IRAS
20100-4156. A Gaussian fit of the Paα spatial profile of
IRAS 20100-4156 is performed along the slit, for the two
nuclei when needed. The 1′′ slit width is reported as
shaded. The Gaussian is thus rotated around each nucleus
and the contributions within the slit aperture and a suit-
able circular aperture around each nucleus, are separately
evaluated.
There exists a wide literature concerning predicted
intrinsic line ratios of Hydrogen recombination lines. In
general these values depend only very slightly on the as-
sumed density of the emitting region (usually between
102 − 104cm−3) but they depend more on the electronic
temperature. The latter may be assumed or may be the re-
sult of detailed modeling of HII regions, accounting for the
hardness of ionizing spectrum, geometry and metallicity
of the gas. To investigate this point, we have run a set of
CLOUDY (Ferland, 2003) models for different HII regions,
characterized by different ages of the ionizing stellar pop-
ulation, different metallicity and gas density. The intrinsic
ratio decreases at increasing metallicity (lower electronic
temperature) and increasing age (lower hardness, neglect-
ing a hardening of the spectrum due to the presence of
Wolf Rayet stars). For example, for the ratio R Hα,P aα
we
find values as high as 8.5 at Z=0.008 and young ages and
as low as 7.0 in metal rich HII regions and/or old clus-
ter ages. Furthermore, more realistic galaxy models must
take into account a law of star formation (almost contin-
int
Fig. 4. Relation between Paα and IR luminosity (upper
panel) and Brγ and IR luminosity (Lower panel). Dashed
line in the upper panel is our linear relation L(line)=
Const × L(IR), while that in the lower panel has ben
taken from the value quoted by Goldader et al. 1997.
with the intrinsic values predicted by models of nebular
emission, Rint.
However, since the only strong Hydrogen line present
in our spectra is either Paα or Brγ we need to compare our
fluxes with other observations taken from the literature.
Duc, Mirabel & Maza (1997), Veilleux, Kim & Sanders
(1999), and Veilleux et al. (1995) provide optical spec-
troscopy within slit apertures of 1.25′′, 2 kpc and 4 kpc re-
spectively, for almost all galaxies in our sample. Observed
intensities of the Hα and Hβ emission lines are reported
in columns two and three of Table 5.
Table 3. Observed Emission Lines Properties of Target Galaxies
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
7
EW FWHM FWHM
[km−1]
482.26
475.85
348.24
243.83
243.82
278.64
423.43
215.47
IRAS name
00085-1223
00188-0856
00582-0258
01077-1707
02411+0354NE
Line
HeI
Brγ
HeII
H2 1-0S(0)
H2 1-0S(5)
HeI
Paα
H2 1-0S(3)
HeI
Paα
H2 1-0S(3)
H2 1-0S(2)
HeI
H2 1-0S(1)
Brγ
HeI
H2 1-0S(5)
HeI
Paα
H2 1-0S(4)
Brδ
H2 1-0S(3)
02411+0354SW
06206-6315
Paα
Paα
19335-3632
H2 2-1S(6)
Paα
Brδ
H2 2-1S(3)
HeI
20100-4156S
H2 2-1S(5)
20100-4156N
22206-2715NE
22206-2715SW
22491-1808
23128-5919N
23230-6926
23389-6139
HeI
Paα
Brδ
H2 2-1S(3)
HeI
Paα
Paα
Paα
Paα
Brδ
H2 1-0S(3)
H2 1-0S(1)
HeI
H2 1-0S(2)
H2 1-0S(1)
Brγ
Paα
Brδ
H2 1-0S(3)
H2 1-0S(2)
Paα
H2 2-1S(6)
Brδ
H2 1-0S(3)
[FeII]
λ0
[µm]
2.1500
2.1661
2.2155
2.2235
1.8353
1.8689
1.8756
1.9570
1.8689
1.8756
1.9570
2.0338
2.0587
2.1213
2.1661
1.8150
1.8353
1.8689
1.8756
1.8920
1.9445
1.9570
1.8756
1.8756
1.8942
1.8756
1.9445
1.9570
2.0587
1.8353
1.8689
1.8756
1.9445
1.9570
1.8689
1.8756
1.8756
1.8756
1.8756
1.9445
1.9570
2.1213
2.0587
2.0338
2.1213
2.1661
1.8756
1.9445
1.9570
2.0338
1.8756
1.8942
1.9445
1.9570
1.9670
Flux
[erg s−1 cm−2]
4.99×10−15
1.25×10−14
4.34×10−15
6.15×10−15
4.21×10−16
3.86×10−16
5.97×10−15
5.01×10−16
9.34×10−17
1.63×10−15
2.05×10−16
1.29×10−15
2.22×10−15
2.16×10−15
5.61×10−15
3.32×10−15
1.56×10−16
5.25×10−16
8.30×10−15
2.52×10−16
3.29×10−16
3.73×10−16
2.18×10−15
5.05×10−15
5.66×10−16
6.15×10−15
5.19×10−16
4.07×10−16
4.71×10−15
3.1×10−16
8.8×10−16
6.09×10−15
6.00×10−16
1.66×10−15
3.29×10−16
1.70×10−15
2.81×10−15
6.58×10−16
4.94×10−15
1.21×10−16
7.04×10−16
8.31×10−16
3.75×10−15
8.28×10−15
5.85×10−15
1.37×10−14
7.21×10−15
5.93×10−16
1.30×10−15
2.87×10−16
7.34×10−15
1.71×10−16
9.30×10−17
2.38×10−15
2.63×10−16
[restA]
-1.04
-2.68
-0.99
-1.43
-1.81
-1.79
-29.4
-2.63
-1.89
-31.56
-3.18
-1.89
-3.27
-3.45
-9.38
-6.95
-2.95
-8.58
-151.5
-4.82
-6.03
-5.59
-92.3
-30.25
-3.60
-124.0
-7.43
-5.82
-8.78
-11.63
-18.61
-136.2
-13.91
-39.04
-14.08
-76.95
-37.78
-13.18
-90.96
-2.00
-11.93
-18.75
-8.30
-18.48
-13.99
-30.78
-111.8
-8.94
-21.94
-4.45
-77.48
-1.72
-0.90
-27.34
-2.89
[restA]
34.26
35.51
29.97
20.45
16.86
18.87
27.88
16.45
4.37
36.19
24.26
18.27
26.72
23.68
33.43
23.00
17.62
26.38
22.67
19.11
17.39
21.31
24.37
42.01
11.15
24.64
17.93
18.79
18.23
30.25
62.87
31.36
35.02
51.47
33.40
25.46
26.67
28.79
30.07
12.26
34.18
33.77
25.75
24.44
27.87
57.11
35.53
20.70
33.92
15.13
49.56
4.65
2.76
43.44
15.24
561.35
294.90
218.47
409.34
290.23
514.02
347.08
251.22
420.41
339.52
305.08
272.50
301.19
369.74
651.92
137.46
368.26
228.89
264.55
237.63
497.00
1008.76
501.56
540.18
788.84
535.97
407.21
408.40
439.67
449.54
142.29
513.09
459.81
427.81
328.09
375.23
811.43
501.97
303.25
497.54
174.93
676.14
639.7
150.4
8
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
uous in the case of a normal star forming galaxy) which,
in general will diminish the hardness of the ionizing spec-
trum. In all the following analysis we will make use of the
recent models for normal star forming galaxies presented
by Panuzzo et al. (2003). In these models the galaxy evo-
lution is followed with a chemical evolution code and the
spectro-photometric properties are computed with metal
dependent stellar population synthesis (Bressan, Chiosi
& Fagotto 1994). Line emission is also accounted for by
means of a large data set of HII region models computed
with CLOUDY, with the appropriate chemical composi-
tion. The code (GRASIL, Silva et al. 1998) also account
for attenuation of light by dust, both in molecular clouds
and in the more diffuse cirrus component. These models
have been successfully tested against a number of different
observations, from UV to the radio (Granato et al. 2000,
Bressan et al. 2002 and Panuzzo et al. 2003).
int
int
≃3.0, RHα,P aα
We adopt the values quoted by Panuzzo et al. (2003)
for a star-forming galaxy, i.e. RHα,Hβ
≃7.2
and R Hα,Brγ
≃93.5. The latter two values, in particular,
are lower than that quoted by Hummer & Storey (1987)
( ≃8.58, and ≃103.54), which however refer to a density
of 104cm −3 and a temperature of Te=104K. Our sin-
gle population CLOUDY models show that, already at
Z=0.02 (solar), the temperature is below 8000K and e.g.
R Hα,P aα
by Hummer & Storey
(1987), would increase the derived E(B-V) by about 0.09
mag.
≃8.0. Adopting RHα,P aα
int
int
int
To evaluate the extinction from the observed Paα or
Brγ fluxes, we have re-scaled our observations to the Hα
aperture by multiplying our values by the coefficients CHα
listed in Table 4. As for the extinction law we have used
that of Calzetti et al. (2000), which is suited for starburst
galaxies. The derived values of AV are shown in Table 5.
Table 4. Coefficients for slit corrections.
CHα CIR
2.08
1.67
1.74
1.44
3.44
1.90
1.88
2.44
3.56
1.70
3.35
3.03
IRAS name
00085-1223
00582-0258
1.18
00188-0856
1.28
01077-1707
02411+0354NE
1.05
02411+0354SW 1.05
06206-6315
1.32
19335-3632
20100-4156S
20100-4156N
22206-2715
22206-2715NE
22206-2715SW
22491-1808
23128-5919
23230-6926
23389-6139
0.83
1.24
1.16
1.22
4.79
2.04
3.02
1.76
0.41
1.21
3.3. The star formation rate.
To derive the star formation rate from the observed line
intensities we have adopted the calibrations provided by
Panuzzo et al. (2003):
SFR(Paα) = 5.06×10−41 L(Paα) M⊙yr−1/(erg s−1)
SFR(Brγ) = 6.60×10−40 L(Brγ) M⊙yr−1/(erg s−1)
For the IR luminosity (Table 1) we have adopted the
calibration provided by the same authors between the star
formation rate and the 8µm-1000µm infrared luminosity:
SFR(IR) = 4.63×10−44 L(IR) M⊙yr−1/(erg s−1)
All the above calibrations refer to a Salpeter IMF be-
tween 0.1M⊙ and 120M ⊙ and take into account an incre-
ment of 16% due to the lower IMF limit with respect to the
value of 0.15M⊙ adopted by Panuzzo et al. (2003). Paα
and Brγ fluxes were corrected for extinction and multi-
plied by the coefficients CIR listed in Table 4. Luminosities
have been computed adopting distances quoted in Table
1.
The values of SFR, obtained in this way from different
indicators, are reported in Table 6. The last column of
Table 6 indicates the ratio between the SFR derived from
the hydrogen line emission and that derived from the IR
luminosity.
On average the Paα luminosity, even corrected for slit
losses and attenuation, provides a SFR which is only 14%
of that derived from the far infrared luminosity (assuming
that the latter is entirely due to the starburst). In the case
of the objects showing the Brγ line, the average ratio is
60%.
4. Discussion
In all the observed galaxies the Paα or Brγ flux, even
corrected for aperture effects and extinction derived from
NIR optical recombination lines (Table 5), is significantly
less than that expected from a starburst of corresponding
bolometric luminosity. A similar "deficit" of recombina-
tion photons has already been noticed by Goldader et al.
(1995), based on the analysis of the Brγ line in a sample
of local ULIRGs.
Poggianti, Bressan & Franceschini (2001) have found
that the SFR derived from the extinction corrected inten-
sity of Hα in a sample of ULIRGs was a factor of three
less than that derived from the FIR luminosity. They at-
tributed the discrepancy to an age-selective extinction ef-
fect where a substantial young population was essentially
escaping the optical detection. On the contrary, in a recent
analysis of lower infrared luminosity starburst galaxies (10
≤log(LIR) ≤11) Mayya et al.(2004) have not found strong
evidence for a loss of optical photons.
We discuss below several possible explanations of the
differences between the SFR (and/or attenuation) derived
from the emission lines and from the IR emission.
Table 5. Dust Extinction
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
9
IRAS name
SHβ
SHα
AV
Hα/Hβ Paα/Hα Brγ/Hβ
4.08e-15a
00085-1223
2.45e-16b
00188-0856
1.16e-14a
01077-1707
8.00e-16b
02411+0354NE
02411+0354SW 1.10e-15b
2.60e-16c
06206-6315
1.86e-15c
20100-4156S
8.96e-16b
22206-2715
3.15e-15b
22491-1808
8.86e-15c
23128-5919
1.08e-15c
23230-6926
3.40e-16c
23389-6139
1.00e-13a
3.50e-15b
9.90e-14a
4.20e-15b
7.60e-15b
6.31e-15c
1.61e-14c
6.40e-15b
2.30e-14b
8.29e-14c
1.35e-14c
2.26e-14c
6.08
4.52
3.03
1.62
2.41
6.05
3.06
2.51
2.57
3.29
4.13
8.96
4.28
4.33
1.24
3.25
1.91
0.75
0.40
2.39
1.68
5.09
2.97
4.11
a) Veilleux et al. (1995); b) Veilleux, Kim & Sanders (1999); c) Duc, Mirabel & Maza (1997)
Table 6. SFR derived from the observed Paα (or Brγ) and IR luminosity. The spectral lines data have been corrected
for attenuation and for slit losses.
IRAS name
00085-1223
00188-0856
00582-0258
01077-1707
02411+0354
02411+0354NE
02411+0354SW
06206-6315
19335-3632
20100-4156
20100-4156S
20100-4156N
22206-2715
22491-1808
23128-5919
23230-6926
23389-6139
L(Paα)
[1041erg s−1]
SFR(Paα)
[M⊙ yr−1]
L(Brγ)
[1041erg s−1]
4.05
SFR(Brγ)
[M⊙ yr−1]
26.7
9.1
0.47
13.22
10.17
3.04
8.78
3.20
12.00
9.08
2.92
6.54
6.38
15.01
15.42
48.1
2.4
66.9
51.5
15.4
44.4
16.2
60.7
45.9
14.8
33.1
32.3
75.9
78.0
8.22
54.2
16.68
110.14
L(IR)
[1045erg s−1]
1.08
12.32
2.22
1.92
6.83
7.73
1.87
20.59
6.49
7.13
3.71
9.29
7.11
log
(LIR/L⊙)
11.45
12.51
11.76
11.70
12.25
12.30
11.69
12.73
12.23
12.27
11.99
12.38
12.27
SFR(IR)
[M⊙ yr−1]
49.9
570.5
102.7
89.1
316.2
357.9
86.6
953.1
300.4
330.1
171.8
430.0
329.0
Line/IR
0.54
0.08
0.02
0.61
0.21
0.12
0.19
0.06
0.11
0.10
0.64
0.18
0.24
Paα luminosities for IRAS00582-0258, IRAS19335-3632 and IRAS20100-4156N are not extinction corrected, and the SFRs are
only lower limits.
4.1. The AGN contribution
The first and most obvious cause for the photon deficit
is that a significant contribution to the infrared emis-
sion comes from an AGN. There is evidence that in the
most luminous nearby infrared galaxies a significant frac-
tion of the luminosity could arise in principle from an
AGN instead of the starburst, as suggested by some opti-
cal (Veilleux et al. 1995), NIR (Veilleux, Sanders & Kim,
1999) and ISO ( Tran et al. 2001) spectroscopic studies of
ULIRGs.
As already anticipated we have possibly detected
broad emission line components and [SiVI] emission only
in IRAS 00188-0856 and IRAS 00582-0258.
It is interesting to note that,
in the IR colour-
luminosity diagram (Neff & Hutchings 1992), both
IRAS 00188-0856 with log(S25µm/S60µm)=−0.63 and
log(L 60µm/L⊙)=12.05, and IRAS 00582-0258 with log(S
25µm/S60µm)=− 0.55 and log(L 60µm/L⊙)=11.24, fall in
the region of overlap between QSO and Seyfert 1 galaxies.
To explain the observed large discrepancy, the AGN
should provide a conspicuous fraction of the IR luminosity,
generally exceeding 80%.
The presence of a broad component (FWHM>2000
km/s) in the Paα line, the possible detection of the high
excitation [SiVI] line, and their position in the IR colour-
10
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
luminosity diagram, favour the existence of an obscured
AGN in IRAS 00188-0856 and IRAS 00582-0258.
raised by Farrah et al. (2003) (but see also Prouton et al.
2004).
4.1.1. Other diagnostics for unveiling the AGN
Hints on the nature of the IR luminosity may come also
from other wavelength observations.
In star forming galaxies, FIR and radio emissions are
tightly correlated over a wide range of IR luminosities.
Sanders & Mirabel (1996) report:
q = log
FF IR/(3.75 × 1012Hz)
Fν(1.49GHz)/(W m−2Hz−1)
≃ 2.35 ± 0.2
(1)
where FF IR=1.26×10−14(2.58S60µm + S100µm) Wm−2,
with S60 and S100 expressed in Jy. Radio observations
of all our target objects have been reported in Table 1.
After extrapolating a few 843MHz data to 1.4GHz by as-
suming a radio slope Fν ∝ ν−0.8, we have evaluated the q
parameter, reported in Table 1.
All
the objects,
including IRAS00188-0856 and
IRAS00582-0258, but IRAS22491-1808 and IRAS23389-
6139, fall on top of the FIR-radio relation of starburst
galaxies. Panuzzo et al. (2003) have shown that in normal
star-forming galaxies the radio luminosity is an excellent
indicator of star formation and that deviations from this
relation are due to variations of the corresponding FIR
luminosity as the latter depend more on the details of
the obscuration. In contrast, in a very young obscured
starburst there is an excess of FIR emission because core
collapsed supernovae, thought to provide the 90% of ra-
dio emission, have a delay of a few million years. For the
same reason, if the starburst is in a late phase and the
star formation decreased exponentially, there is an excess
of radio emission (Bressan, Silva & Granato 2001).
Based on different arguments, Smith et al. (1998),
Farrah et al. (2003) and Prouton et al. (2004), have shown
that the contribution of the putative AGN to the radio
luminosity in obscured starbursts (not harbouring radio-
loud sources) is low. Thus a significant contribution to the
IR from the AGN would significantly raise the value of q
above the FIR-radio correlation. This might be particu-
larly relevant for IRAS00188-0856 and IRAS 00582-0258
where, as we have already seen, there are hints for the
presence of the AGN from our own observations, but in
fact these galaxies show a q parameter typical of star-
bursts indicating that the contribution of the AGN in the
IR is not high.
Another diagnostic for the presence of the AGN is a
low value of the ratio of the line to continuum emission
of the 7.7 µm PAH feature, as measured by ISO spec-
troscopy (Lutz et al. 1998). Unfortunately, not all our
galaxies have this ratio measured. Following this indicator
(Table 1), three sources could harbour an obscured AGN,
IRAS 00188-0856, IRAS 23230-6926 and IRAS 23389-
6139. However only the first source shows evidence for an
AGN from our NIR observations. This may give further
support to criticism on the use of this indicator recently
There are also XMM-Newton hard X-ray observations
for IRAS 20100-4156 (Franceschini et al. 2003). IRAS
20100-4156 shows a faint X-ray flux of S(2-10 keV)≃ 1.9×
10−14 erg s−1cm−2, with a (2-10Kev)/(8-1000µm) flux ra-
tio ≃4.75 × 10−5. Unfortunately, its large distance and
consequent poor X-ray photon statistics prevented any
careful X-ray spectral analyses. Although the total source
flux is slightly above that expected from a starburst of
similar FIR power (Franceschini et al. 2003), there is no
definite evidence, from hard X-rays, that the galaxy hosts
an absorbed AGN. For instance MKN231, a typical Sy1
galaxy, is about 34 times brighter, between 2 and 10 keV,
than IRAS 20100-4156, in spite of having about the same
IR luminosity.
In summary, from our own observations and other col-
lected data in the literature we may conclude that IRAS
00188-0856 and IRAS 00582-0258 could harbour an AGN.
However in all cases the contribution of the AGN to the IR
seems not at the level required to explain the discrepancy
between line flux and FIR emission. Indeed the first evi-
dence of a buried AGN in the ultraluminous galaxy UGC
5101, provided by SPITZER, allows a contribution to the
total luminosity of only ≤10% (Armus et al. 2004).
4.2. High NIR extinction
Another alternative to explain the photon recombination
deficit is simply that the extinction is very high, even at
NIR wavelengths, and that what we see from optical lines
is only the external skin of the starburst.
To evaluate the possible range of that kind of extinc-
tion we have compared the observed values of the Paα/IR
(or Brγ/IR) ratio with that predicted by models of normal
star- forming galaxies. In this way we will obtain the ex-
cess extinction over that of a normal star forming galaxy.
From Panuzzo et al. (2003) we get the following average
relations between line and IR (8µm -- 1000µm) emission:
S(Hα)/S(IR)=6.56×10−3,
S(Paα)/S(IR)=7.9×10−4, and
S(Brγ)/S(IR)=6.32×10−5
The observed values of S(Paα)/S(IR) or S(Brγ)/S(IR)
are reported in Table 7. The observed intensities of the
emission lines have been corrected for slit losses by mul-
tiplying by the factor CIR of Table 4. The average value
of the observed S(Paα)/S(IR) ratios is ≃10% of that pre-
dicted by the models while, in the case of Brγ line, it is
≃40% of the predicted value.
Assuming that the difference with respect to the mod-
els is entirely due to an extinction larger than in the case
of a normal star forming galaxy, we have derived the corre-
sponding attenuation in magnitudes, in the line and in the
visual. The values of AV must be considered as in excess
of that of normal star forming galaxies which is typically
AV ≃1 mag. In this way we obtain an average attenuation
at Paα of about 2.9 mags, corresponding to an average vi-
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
11
sual attenuation of about 20 mags. These values are much
higher than those derived from optical emission lines (cfr.
Table 5). Values of AV ∼5-50 have been already found
by Genzel at al. (1998) from mid-infrared spectroscopy of
a small sample of ULIRGs and are confirmed by recent
SPITZER observations of selected ULIRGs ( Armus et al.
2004). High optical depths (≥20) at 1µm have been also
inferred from fitting the far infrared to radio SEDs of a
sample of compact ULIRGs (Prouton et al. 2004).
Table 7. Extinction derived from comparison of observed
and predicted line flux to IR ratios
IRAS name
Line/IR Obs/Modela A(line)
00188-0856
00582-0258
02411+0354
06206-6315
19335-3632
20100-4156
22206-2715
22491-1808
23230-6926
23389-6139
00085-1223
01077-1707
23128-5919
2.73E-05
2.13E-05
1.42E-04
3.97E-05
1.72E-04
4.01E-05
6.53E-05
5.73E-05
7.86E-05
4.56E-05
1.88E-05
3.02E-05
3.08E-05
Paα
3.45E-02
2.70E-02
1.80E-01
5.03E-02
2.18E-01
5.08E-02
8.27E-02
7.25E-02
9.95E-02
5.78E-02
Brγ
2.98E-01
4.77E-01
4.88E-01
3.65
3.92
1.86
3.25
1.65
3.23
2.71
2.85
2.51
3.10
1.31
0.80
0.78
Ab
V
24.9
26.7
12.7
22.1
11.2
22.0
18.4
19.4
17.0
21.1
13.5
8.3
8.0
a) Observed L(Paα)/IR = 7.9E-4 and L(Brγ)/IR = 6.3E-5, for
a normal star forming galaxy (Panuzzo et al. 2003)
b) AP aα/AV =0.147 and ABrγ /AV =0.097 from Calzetti et al.
(2000)
We also notice that, in the case of Paα galaxies, even
assuming an AGN contribution to the IR of the order of
50% (which is large, following Prouton et al. 2004), the
visual attenuation would still remain between 15 and 25
mags.
As a second point, we notice that in the galaxies ob-
served in the Brγ domain the extinction is lower than that
obtained for the sample with Paα in agreement with a
slab attenuated model. In fact, for such a model we expect
A(Brγ)≃0.66 ×A(Paα) (Calzetti et al. 2000). The average
extinction of the Brγ sample is A(Brγ)≃0.96, while taht
of the Paα sample is A(Paα)≃2.9, with a ratio of 0.33.
Though we are comparing different galaxies and the statis-
tics is low (but notice that our Brγ galaxies lie on the re-
lation defined by the more exhaustive sample of Goldader
et al. 1997), this may indicate that the photon deficit is
wavelength dependent, an thus favour the effects of a com-
plex extinction geometry over those of absorption by dust
within HII regions, discussed below.
4.3. Dust within the HII regions
In presence of dust within the ionized regions of the star-
burst, only a fraction f of Ly-continuum photons may
effectively ionize the gas, while the remaining (1-f ) is ab-
sorbed. This effect causes a deficit of recombination pho-
tons, with respect to the FIR emission. To explain the
observed discrepancy between NIR line emission and IR
flux, dust within HII regions should absorb about 80% of
the ionizing flux.
Dust absorption within HII regions has been invoked
by Hirashita et al. (2003) to explain the anomalous Hα to
UV flux, observed in a sample of IUE selected star forming
galaxies. Hirashita et al. (2003) have estimated an aver-
age value of (1-f )≃50%, with some objects reaching (1-f )
≃80%. However, it is worth recalling that the conclusions
reached by Hirashita et al. (2003) are based on a sin-
gle screen extinction model. By analyzing the same data,
Panuzzo et al. (2003) have instead found that this effect
may result from "age-selective" extinction, namely that
younger populations are more extinguished than older
ones, the latter still contributing to the UV flux but not
to nebular emission.
Dust within ionized regions has been indicated also
by Luhman et al. (2003) as a possibility to explain the
[CII] deficit relative to FIR, observed in their sample of
ULIRGs. In fact dust would not only affect the pool of
ionizing photons, but would also inhibit the penetration
of 13.6eV−6eV photons (thought to be responsible of the
[CII] excitation) in the photo-dissociation regions, while
preserving the overall FIR emission.
If confirmed by further studies, a 80% depression of the
ionizing flux would deeply challenge any determination of
SFR from even the less extinguished emission lines.
4.4. Age of the starbursts
Estimators of the SFR are usually derived assuming a
continuous star formation rate. This is essentially correct
when dealing with normal galaxies, but starburst galaxies
are, by definition, currently dominated by a single episode,
with a duration which is generally comparable to the char-
acteristic times of the star formation indicators. In such
circumstances the use of different calibrators may lead to
significant discrepancies. For instance, this may be the
case of the Radio and IR SFR indicators, as discussed by
Bressan, Silva & Granato (2002).
To get hints on the evolutionary status of our galaxies
we have analyzed the equivalent widths (EW) of hydrogen
emission lines (see e.g. Terlevich et al. 2004). These indica-
tors are relatively unaffected by dust extinction in at least
two cases, namely when both continuum and line emis-
sion arise from the same population, or when attenuation
can be approximated by an intervening absorbing slab. In
the latter case however, the equivalent width may be af-
fected by the presence of the old population, outside the
starburst region. Figure 6 compares the observed Paα and
Brγ equivalent widths of our galaxies with those predicted
12
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
alent width of the lines and the corresponding SFR show
a comparable distribution in the two samples (though the
Brγ sample is not statistically significant). This not only
suggests that the line emission originates from the cen-
tral regions, but also indicates that the equivalent widths
are dominated by sources that share the same spatial dis-
tribution. For instance, they should not be significantly
affected by an underlying old population, as expected in
vigorous starbursts (Mayya et al. 2004).
We conclude that the observed distribution of EWs,
skewed toward old ages, is actually another aspect of the
photon deficit problem, independent from bolometric lu-
minosity considerations. Unfortunately a statistical argu-
ment based on the distribution of EWs cannot help to
discriminate between the different causes. In fact, strong
age selective extinction could make the starburst to ap-
pear older than it is; absorption of ionizing photons could
diminish the equivalent width, because the K band contin-
uum of the ionizing cluster would suffer much less atten-
uation; and finally the K-band continuum could be dom-
inated by the presence of an AGN, still compatible with
the FIR/Radio correlation (Prouton et al. 2004).
5. Conclusions and Perspectives
We have obtained NIR medium dispersion long slit spec-
troscopy with SOFI at NTT of ten luminous infrared
galaxies having suitable redshift to push their Paα emis-
sion in the Ks band. We included also three objects with
lower redshift, for which the Brγ line was in the Ks band.
We have found that the Paα emission, even corrected
for slit losses and for extinction estimated by compar-
ing our data with optical spectroscopy, is significantly less
than that expected from a starburst of corresponding bolo-
metric luminosity. The discrepancy is lower for the galax-
ies observed in the Brγ region, but in general we confirm
the existence of a deficit of recombination photons first
pointed out in ULIRGs by Goldader et al. (1995).
Furthermore, in IRAS 00188-0856 and IRAS 00582-
0258 we find evidence for significant broadening of the Paα
line and for the presence of [SiVI] coronal line. However
these two galaxies fall on top of the FIR/Radio correla-
tion, indicating that, though present, the AGN does not
dominate the far infrared emission. For all other sources
we do not find evidence of AGN contribution and we ar-
gue that the studied galaxies appear to be predominantly
powered by a nuclear starburst.
Based on current data alone it is impossible to disclose
the origin of the recombination photon deficit, and we may
only advance the following hints.
The galaxies may harbour a highly attenuated star
forming region. In this case an estimate of the attenua-
tion can be obtained from the comparison the observed
line/IR emission ratio with the predictions of models for
normal star-forming galaxies. Then the average attenua-
tion would be AV ≃10-25mag and AP aα ≃2-4mag. These
figures are slightly lower when derived from the Brγ/IR ra-
tio (A Brγ ≃1mags) and are consistent with the decrement
Fig. 6. Starburst ages derived from the equivalent width
of Paα and Brγ emission. The SFR of the models decreases
exponentially with e-folding time of τSF R=25Myr. The
EW of the models is plotted against the ratio between
the current age and the SFR e-folding time so that the
figure does not change significantly by considering other
plausible values of τSF R. Observed values are plotted as
horizontal segments.
by starburst models. The models have been constructed
by adding nebular emission to the spectral energy distri-
bution of simple stellar populations of solar metallicity,
as described in Bressan, Poggianti & Franceschini (2001).
The adopted star formation e-folding time is τSF R=25,
but since we are interested in the ratio between the age
of the starburst and the e-folding time, our conclusions
do not change significantly by assuming other plausible
values for τSF R.
From Figure 6, we observe a lack of young objects, with
only one source, IRAS 23128-5919, appearing younger
than twice the SFR e-folding time. The other galaxies
have ages that are either between two and three times
the SFR e-folding time, or around four times the e-folding
time. Taken at face value, these estimates would place
IRAS 00188-0856, IRAS 00582-0258, IRAS 06206-6315
and IRAS 22206-2715NE in a post-starburst phase. Part
of the photon deficit discrepancy could then originate sim-
ply from our use of stationary models. However, given the
high current IR luminosity of the latter galaxies, their
bolometric luminosity at the peak SFR would have been
unreasonably large, between 3×1013L⊙ and 2×1014L⊙.
It is worth noticing that, being the average distance of
the Paα sample about four times that of the Brγ sample,
our slit is sampling an intrinsic galaxy area which increases
by at least one order of magnitude going from the Brγ
sample to the Paα sample. In spite of that, both the equiv-
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
13
of attenuation going from Paα to the Brγ wavelength re-
gion. A significant attenuation of the nuclear star forming
region even in the NIR comes out to be in agreement with
the large molecular cloud optical depths ( τ1µm ≥ 20), de-
rived by Prouton et al. (2004) from SED fitting of compact
ULIRGs. It is also in agreement with recent SPITZER
observations of ultra luminous galaxies with buried AGN
(Armus et al. 2004). Assuming that the AGN may con-
tribute about 50% of the IR flux (a quite extreme figure,
Prouton et al. 2004), would imply only a slightly smaller
attenuation.
Our finding is also compatible with a scenario where
a large fraction (≃80%) of the ionizing flux is absorbed
by dust within the HII regions. The required fraction is
large but comparable with some extreme values found by
Hirashita et al. (2003) in their analysis of a sample of UV
selected starbursts. This effect has been recently invoked
as one of the possible causes of the [CII] emission deficit
relative to FIR observed in some ULIRGs (Luhman et al.
2003 ). Based on MIR to radio SED fitting, Prouton et al.
(2004) have estimated that the dust sublimation radius
within molecular clouds in compact ULIRGs, is generally
a fraction of a parsec. Dust may thus survive even in the
innermost regions of an HII region and, if such a strong
effect is confirmed, it will challenge our ability to derive
SFR and/or to study inner environmental conditions, from
even the most un-extinguished emission lines.
Though the above alternative scenarios are both com-
patible with the present data, their predictions at longer
wavelengths are markedly different and suggest that a de-
cisive test to disentangle between high nuclear obscuration
and dust absorption within HII regions, would be to look
at the Brα 4.05 µm line emission. In fact, assuming the
Calzetti et al. (2000) extinction law and an intrinsic ratio
L( Paα)/L(Brα)≃3.9 (Panuzzo et al. 2003) we get
L(Brα)/L(P aα) ≃ 0.26 × 100.04AV .
Thus if the Paα deficit is due to absorption of ionizing pho-
tons by dust within HII region and AV ≃3 mag, as derived
from optical and NIR-optical emission line ratios, then
the expected ratio L(Brα)/L(Paα) is ≃0.34. On the other
hand, if the Paα emission comes from a dust enshrouded
region with say AV ≃20 mag, then L(Brα)/L(Paα) ≃1.6,
a factor about five times larger than in the previous case.
referee
Acknowledgements. We tank the anonymous
for
her/his comments and suggestions and M. Clemens for dis-
cussions and careful reading of the manuscript. A.B. ac-
knowledges warm hospitality by INAOE and J.R.V. acknowl-
edges warm hospitality by INAF, Osservatorio Astronomico
di Padova. S.B. acknowledges support by ASI research grant
no. I/R/062/02. This research was partially supported by the
European Commission Research Training Network 'POE' un-
der contract HPRN-CT-2000-00138 and by MURST under
COFIN n. 2001/021149
References
Barger, A. J., Cowie, L. L. & Richards, E. A. 2000, AJ, 119,
209.
Berta, S., Fritz, J., Franceschini, A., Bressan, A. & Pernechele,
C.: 2003, A&A, 403, 119.
Bressan, A., Silva, L., & Granato, G. L. 2002, A&A, 392, 377
Bressan, A., Poggianti, B., & Franceschini, A. 2001, QSO Hosts
and Their Environments, 171
Bressan, A., Chiosi, C., Fagotto, F. 1994, ApJS, 94, 63
Bushouse, H.A., Borne, K.D., Colina, L., Lucas, R.A., Rowan-
Robinson, M., Baker, A.C., Clements, D.L., Lawrence, A.,
Oliver, S., 2002, ApJSS, 138, 1
Calzetti, D., Armus, L., Bohlin, R. C., Kinney, A. L.,
Koornneef, J., & Storchi-Bergmann, T. 2000, ApJ, 533,
682
Condon, J. J., Helou, G., Sanders, D. B., & Soifer, B. T. 1996,
ApJS, 103, 81
Condon, J. J.; Cotton, W. D.; Greisen, E. W.; Yin, Q. F.;
Perley, R. A.; Taylor, G. B.; Broderick, J. J., 1998, AJ,
115, 1693
Duc, P.-A., Mirabel, I. F. & Maza, J. 1997, A&AS, 124, 533
Elbaz, D., Cesarsky, C.J., Fadda, D., et al. 1999, A&A, 351,
37
Farrah, D., Afonso, V., Efstathiou, A., Rowan-Robinson, M.
,Fox, M., Clements, D., 2003, MNRAS, 343, 585
Ferland, G. J. 2003, ARA&A, 41, 517
Fisher, K.B., et al. , 1995, ApJS, 100, 69
Franceschini A., Aussel H., Cesarsky C., Elbaz D., Fadda D.:
A&A, 2001, 378, 1.
Franceschini, A., Braito, V.,Persic, M. et al. 2003, MNRAS,
343, 1181
Gear, W. K., Lilly, S. J., Stevens, J. A., Clements, D. L., Webb,
T. M., Eales, S. A., & Dunne, L. 2000, MNRAS, 316, L51
Genzel, R., Lutz, D., Sturm, E., et al. 1998, ApJ, 498, 579
Goldader, J. D.; Joseph, R. D.; Doyon, R.; Sanders, D. B. 1997,
ApJS, 108, 449
Goldader, J. D., Joseph, R. D., Doyon, R.; Sanders, D. B. 1995,
ApJ, 444, 97
Granato, G.L., Lacey, C.G., Silva, L., Bressan, A., Baugh,
C.M., Cole, S., Frenk, C.S. 2000, ApJ, 542, 710
Hirashita, H., Buat, V., Inoue, A. K., 2003, A&A, 410, 83
Hummer, D. G.; Storey, P. J. 1987, MNRAS, 224, 801
Ivison, R. J. et al. 2002, MNRAS, 337, 1
Kennicutt, R.C. 1998, ARAA, 36, 189
Kim, A. G., Gabi, S., Goldhaber, G., et al. 1997, ApJ, 476,
L63
Luhman, M.L., et al. , 2003, ApJ, 594, 758
Lutz, D., Spoon, H. W. W., Rigopoulou, D., Moorwood,
A. F. M., & Genzel, R. 1998, ApJ, 505, L103
Mauch, T.; Murphy, T.; Buttery, H. J.; Curran, J.; Hunstead,
R. W.; Piestrzynski, B.; Robertson, J. G.; Sadler, E. M.,
2003, MNRAS, 342, 1117
Mayya, Y.D., Bressan, A., Rodriguez, M., Valdes, J.R.,
Chavez, M., 2004, ApJ, 600, 188
Moorwood A., Cuby J.G. & Lidman C. 1998, The Messenger
91, 9
Neff, S. G., Hutchings, J. B., 1992, AJ, 103, 1746
Panuzzo, P., Bressan, A., Granato, G. L., Silva, L., & Danese,
L. 2003, A&A, 409, 99
Pernechele, C., Berta, S., Marconi, A., Bonoli, C., Bressan, A.,
Franceschini, A., Fritz, J., & Giro, E. 2003, MNRAS, 338,
L13
Pickles, A.J. 1998, PASP, 110, 863
Poggianti, B. M., Bressan, A., & Franceschini, A., 2001, ApJ,
Armus, L.; et al. 2004, AAS, 204, 3319
550, 195
14
J.R. Vald´es et al.: NIR Spectroscopy of LIRGs
Prouton, O. R.; Bressan, A.; Clemens, M.; Franceschini, A.;
Granato, G. L.; Silva, L. 2004, A&A, 421, 115
Rigopoulou, D., Spoon, H. W. W., Genzel, R., Lutz, D.,
Moorwood, A. F. M., & Tran, Q. D. 1999, AJ, 118, 2625
Sanders, D. B. & Mirabel, I.F. 1996, ARAA, 34, 7479
Silva, L., Granto, G.L., Bressan, A., Danese, L. 1998, ApJ, 509,
103
Smail, I., Ivison, R.J., Owen, F.N., et al. 2000, ApJ, 528, 612
Smail, I., Ivison, R. J., Gilbank, D. G., Dunlop, J. S., Keel,
W. C., Motohara, K., & Stevens, J. A. 2003, ApJ, 583, 551
Smith, H.E.; Lonsdale, C.J.; Lonsdale, C.J. 1998, A&A,
492,137
Stevens, J. A. et al. 2003, NATURE, 425, 264
Tran Q. D., et al, 2001, ApJ, 552, 527
Terlevich, R., Silich, S., Rosa-Gonz´alez, D., & Terlevich, E.
2004, MNRAS, 348, 1191
Vanzi, L., Bagnulo, S., Le Floc'h, E., Maiolino, R., Pompei, E.,
Walsh, W., 2002, A&A, 386, 464
Veilleux, S., Kim, D.-C., & Sanders, D. B., 1999, ApJ, 522, 113
Veilleux, S., Sanders, D. B., & Kim, D.-C., 1999, ApJ, 522, 139
Veilleux S., Kim, D.-C., Sanders D. B., Mazzarella J. M., Soifer
B. T., 1995, ApJS, 98, 171
Zenner, S., Lenzen, R., 1993, A&ASS, 101, 363
|
astro-ph/0504242 | 1 | 0504 | 2005-04-11T14:16:46 | Constraining the mass transfer in massive binaries through progenitor evolution models of Wolf-Rayet+O binaries | [
"astro-ph"
] | Since close WR+O binaries are the result of a strong interaction of both stars in massive close binary systems, they can be used to constrain the highly uncertain mass and angular momentum budget during the major mass transfer phase. We explore the progenitor evolution of the three best suited WR+O binaries HD 90657, HD 186943 and HD 211853, which are characterized by a WR/O mass ratio of $\sim$0.5 and periods of 6..10 days. We are doing so at three different levels of approximation: predicting the massive binary evolution through simple mass loss and angular momentum loss estimates, through full binary evolution models with parametrized mass transfer efficiency, and through binary evolution models including rotation of both components and a physical model which allows to compute mass and angular momentum loss from the binary system as function of time during the mass transfer process. All three methods give consistently the same answers. Our results show that, if these systems formed through stable mass transfer, their initial periods were smaller than their current ones, which implies that mass transfer has started during the core hydrogen burning phase of the initially more massive star. Furthermore, the mass transfer in all three cases must have been highly non-conservative, with on average only $\sim$10% of the transferred mass being retained by the mass receiving star. This result gives support to our system mass and angular momentum loss model, which predicts that, in the considered systems, about 90% of the overflowing matter is expelled by the rapid rotation of the mass receiver close to the $\Omega$-limit, which is reached through the accretion of the remaining 10%. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. 2368wro
(DOI: will be inserted by hand later)
November 20, 2018
5
0
0
2
r
p
A
1
1
1
v
2
4
2
4
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Constraining the mass transfer in massive binaries through
progenitor evolution models of Wolf-Rayet+O binaries
Jelena Petrovic1
,
2, Norbert Langer1, and Karel A. van der Hucht3
,
4
1 Sterrenkundig Instituut, Universiteit Utrecht, Princetonplein 5, NL -- 3584 CC Utrecht, The Netherlands
2 Astronomical Institute, Radboud Universiteit Nijmegen, Toernooiveld 1, NL -- 6525 ED, Nijmegen, The
Netherlands
3 SRON, Nationaal Instituut voor Ruimte Onderzoek, Sorbonnelaan 2, NL -- 3584 CA Utrecht, The Netherlands
4 Sterrenkundig Instituut Anton Pannekoek, Universiteit van Amsterdam, Kruislaan 403, NL -- 1098 SJ
Amsterdam, The Netherlands
Received; accepted
Abstract. Since close WR+O binaries are the result of a strong interaction of both stars in massive close binary
systems, they can be used to constrain the highly uncertain mass and angular momentum budget during the major
mass transfer phase. We explore the progenitor evolution of the three best suited WR+O binaries HD 90657,
HD 186943 and HD 211853, which are characterized by a WR/O mass ratio of ∼0.5 and periods of 6..10 days. We
are doing so at three different levels of approximation: predicting the massive binary evolution through simple
mass loss and angular momentum loss estimates, through full binary evolution models with parametrized mass
transfer efficiency, and through binary evolution models including rotation of both components and a physical
model which allows to compute mass and angular momentum loss from the binary system as function of time
during the mass transfer process. All three methods give consistently the same answers. Our results show that,
if these systems formed through stable mass transfer, their initial periods were smaller than their current ones,
which implies that mass transfer has started during the core hydrogen burning phase of the initially more massive
star. Furthermore, the mass transfer in all three cases must have been highly non-conservative, with on average
only ∼10% of the transferred mass being retained by the mass receiving star. This result gives support to our
system mass and angular momentum loss model, which predicts that, in the considered systems, about 90% of
the overflowing matter is expelled by the rapid rotation of the mass receiver close to the Ω-limit, which is reached
through the accretion of the remaining 10%.
Key words. Stars:binaries:close, stars:evolution, stars:fundamental parameters, stars:rotation, stars:Wolf-Rayet
1. Introduction
The evolution of a star in a binary system can differ sig-
nificantly from that of an isolated one with the same mass
and chemical composition. The physical processes that en-
ter binary evolution are the gravitational and radiation
field from the companion, as well as the centrifugal force
arising from the rotation of the system. But, most impor-
tant, it is the evolution of the more massive component
that will influence dramatically the evolution of the sys-
tem. In certain evolutionary phases, mass transfer from
one star to another can occur, changing the fundamental
properties of both stars as well as their future evolution.
The rotational properties of binary components may
play a key role in this respect. The evolution of mas-
sive single stars can be strongly influenced by rotation
Send offprint requests to: J.Petrovic,
e-mail: [email protected]
(Heger & Langer 2000; Meynet & Maeder 2000), and evo-
lutionary models of rotating stars are now available for
many masses and metallicities. While the treatment of the
rotational processes in these models is not yet in a final
stage (magnetic dynamo processes are just being included
Heger et al. 2004; Maeder & Meynet 2003), they provide
first ideas of what rotation can really do to a star. Effects
of rotation, as important they are in single stars, can be
much stronger in the components of close binary systems:
Estimates of the angular momentum gain of the accreting
star in mass transferring binaries show that critical rota-
tion may be reached quickly (Packet 1981; Langer et al.
2000; Yoon & Langer 2004b). In order to investigate this,
we need binary evolution models which include a detailed
treatment of rotation in the stellar interior, as in recent
single star models. However, in binaries, tidal processes
as well as angular momentum and accretion need to be
2
Petrovic et al.: WR+O progenitors
considered at the same time. Some first such models are
now available and are discussed below.
Angular momentum accretion and the subsequent
rapid rotation of the mass gainer may be essential for
some of the most exciting cosmic phenomena, which may
occur exclusively in binaries: Type Ia supernovae, the
main producers of iron and cosmic yardsticks to measure
the accelerated expansion of the universe (Yoon & Langer
2004a,b), and gamma-ray bursts from collapsars, which
the most recent stellar models with rotation and mag-
netic fields preclude to occur in single stars (Petrovic et al.
2004; Heger et al. 2004; Woosley 2004). For both, the
Type Ia supernova progenitors and the gamma-ray burst
progenitors, it is essential to understand how efficient the
mass transfer process is and on which physical properties
it depends. Further exciting astrophysical objects whose
understanding is affected by our understanding of mass
transfer comprise X-ray binaries (Chevalier & Ilovaisky
1998) and Type Ib and Ic supernovae (Podsiadlowski et al.
1992).
How much matter can stars accrete from a binary com-
panion? As mentioned above, non-magnetic accretion, i.e.
accretion via a viscous disk or via ballistic impact, trans-
ports angular momentum and can lead to a strong spin-
up of the mass gaining star. For disk accretion, it appears
plausible that the specific angular momentum of the ac-
creted matter corresponds to Kepler-rotation at the stellar
equator; this leads to a spin-up of the whole star to crit-
ical rotation when its initial mass is increased by about
20% (Packet 1981). It appears possible that mass accretion
continues in this situation, as viscous processes may trans-
port angular momentum outward through the star, the
boundary layer, and the accretion disk (Paczynski 1991).
However, as the star is rotating very rapidly, its wind mass
loss may dramatically increase (Langer 1997, 1998), which
may render the mass transfer process inefficient.
Observations of massive post-mass transfer binary
systems constrain this effect. Langer et al. (2003) and
Langer et al. (2004) points out that there is evidence for
both extremes occurring in massive close binaries, i.e. for
quasi-conservative evolution as well as for highly non-
conservative evolution. In the present study, we are in-
terested in those binaries that contain a Wolf-Rayet and
a main sequence O star. We have chosen to focus on three
WN+O systems (HD 186943, HD 90657 and HD 211853)
which have similar mass ratios (≈0.5) and orbital peri-
ods (6..10 days). As clearly the two stars in these systems
must have undergone a strong interaction in the past, an
understanding of their progenitor evolution may be the
key to constrain the mass transfer efficiency in massive
binaries: which fraction of the mass leaving the primary
star is accumulated by the secondary star during a mass
transfer event?
Evolutionary calculations of massive close bina-
ries were performed by various authors. General ideas
about the formation of WR+O binary systems were
given by Paczy´nski (1967), Kippenhahn et al. (1967),
van den Heuvel & Heise (1972). Vanbeveren et al. (1979)
modelled the evolution of massive Case B binaries with
different assumptions for mass and angular momentum
loss from the binary system. Vanbeveren (1982) computed
evolutionary models of massive close Case B binaries with
primary masses between 20 M⊙ and 160 M⊙. He concluded
that most of the WR primaries are remnants of stars
initially larger than 40 M⊙ and that the accretion effi-
ciency in these systems should be very below 0.3 in or-
der to fit the observations. de Loore & de Greve (1992)
computed detailed models of massive Case B binary sys-
tems for initial mass ratios of 0.6 and 0.9, assuming an
accretion efficiency of 0.5. Wellstein & Langer (1999) and
Wellstein et al. (2001) modelled massive binary systems
mass range 12..60 M⊙ assuming conservative evolution,
and Wellstein (2001) presented the first rotating binary
evolution models for initial masses of ≈15 M⊙ and initial
mass ratios q≈1.
While it was realized through these models that dif-
ferent mass accretion may be needed to explain different
observations, these efforts did not have the potential to ex-
plore the physical reasons for non-conservative evolution.
I.e., there is no reason to expect that the mass transfer
efficiency remains constant during the mass transfer pro-
cess in a given binary system, nor that its time-averaged
value is constant for whole binary populations.
It is not yet known which physical processes can ex-
pel matter from a binary system. Vanbeveren (1991) pro-
posed that if a binary component is more massive than
≈40-50 M⊙ it will go through an LBV phase of enhanced
mass loss, which will prevent the occurrence of RLOF.
Dessart et al. (2003) investigated the possibility that ra-
diation pressure from the secondary prevents the accre-
tion. They found that even for moderate mass transfer
rates (5·10−6 M⊙ yr−1) the wind and photon momenta
which emerge from the accretion star can not alter the
dynamics of the accretion stream. Here, we follow the
suggestion that the effective mass accretion rate can be
significantly decreased due to the spin-up of the mass
receiving star (Wellstein 2001; Langer et al. 2003, 2004;
Petrovic & Langer 2004).
The remainder of this paper is organized as follows.
In Sect. 2 we briefly discuss the observational data avail-
able for WR+O binary systems. In Sect. 3 we derive esti-
mates for the masses of both stars in WR+O systems for
given initial masses and accretion efficiencies. In Sect. 4
we present the physics used to compute our detailed evo-
lutionary models. Non-rotating binary evolution models
with an adopted constant mass accretion efficiency are
presented in Sect. 5. Our rotating models in which the
mass accretion efficiency is obtained selfconsistently are
discussed in Sect. 6. We briefly compare our models with
observations in Sect. 7. Conclusions are given in Sect. 8.
2. Observational data
There are about 20 observed Wolf-Rayet+O binary sys-
tems with known masses of components in the catalogue
of van der Hucht (2001). We have chosen to model three
Petrovic et al.: WR+O progenitors
3
spectroscopic double-lined systems: HD 186943 (WN3),
HD 90657 (WN5) and GP Cep (WN6/WCE)have, since
they have similar mass ratios (q=MWR/MO≈0.5) and or-
bital periods (6..10 days).
WN+O systems that also have short orbital periods
are V444 Cyg, CX Cep, CQ Cep, HD 94546, HD 320102
and HD 311884. V444 Cyg has period of 4.2 days and
can be result of stable mass transfer evolution, but since
mass ratio of this system is ∼0.3 we did not include it in
this paper. Orbital periods of CX Cep and CQ Cep are
very short (∼2 days) and these systems are probably the
result of a contact evolution. HD 94546 and HD 320102
are systems with very low masses of WR and O compo-
nents (4 M⊙+9 M⊙ and 2.3 M⊙+4.1 M⊙ respectively) and
HD 311884 is extremely massive WR+O binary system
(51 M⊙+60 M⊙). Recently, an even more massive WR+O
system has been observed 83 M⊙+82 M⊙ (Rauw et al.
2004; Bonanos et al. 2004).
The mass ratio of a binary system is determined
from its radial velocity solution, with an error of 5-10%.
However, to determine the exact value of the masses of
the binary components, the value of the inclination of the
system has to be known. Without knowledge of the incli-
nation, only minimum masses of the components can be
determined, i.e., M sin3i. Massey (1981) determined the
minimum mass for the WR star in HD 186943 to be 9-
11 M⊙. Niemela & Moffat (1982) determined the masses
of the components of HD 90657 in the range 11-14 M⊙ for
the WN4 component and 21-28 M⊙ for the O-type com-
ponent. The masses of the WR components in HD 186943
and HD 90657 given in Table 1 have been determined by
Lamontagne et al. (1996) on the basis of improved values
for the inclination of these systems. Demers et al. (2002)
determined minimum masses of the components of the sys-
tem GP Cep. Previously, Lamontagne et al. (1996) sug-
gested values of MWR=15 M⊙ and MO=27 M⊙ for this
system.
There is no obvious hydrogen contribution in the
WR spectrum in any of these systems (Massey 1981;
Niemela & Moffat 1982). Massey (1981) showed that hy-
drogen absorption lines are fairly broad in the spectrum
of HD 186943, equivalent to vsini≃250 km s−1, thus the
O-type star is rotating much faster than synchronously.
Beside the fact that the binary system GP Cep has a
similar mass ratio and period as the other two systems,
it has some very different properties as well. The spec-
tral type of the WR component in GP Cep is a combi-
nation of WN and WC (WN6/WCE Demers et al. 2002).
Also, Massey (1981) showed that, next to the main pe-
riod of ∼6.69 days of the binary system GP Cep, ra-
dial velocities of absorption lines vary also with a period
of 3.4698 days. He proposed that GP Cep is a quadru-
ple system, consisting of two pairs of stars, WR+O and
O+O. Panov & Seggewiss (1990) suggested that in both
pairs one component is a WR star. However, Demers et al.
(2002) showed that there is only one WR star in this
quadruple system.
3. The simple approach
If the initial binary system is very close (an initial period is
of the order of few days), RLOF occurs while the primary
is still in the core hydrogen burning phase and Case A
mass transfer takes place (fast and slow phase). When the
primary expands due to shell hydrogen burning, it fills its
Roche lobe and Case AB mass transfer starts. During this
mass transfer the primary star loses the major part of its
hydrogen envelope. After Case AB mass transfer, the pri-
mary is a helium core burning Wolf-Rayet star. During all
this time, the secondary is still a main sequence star, but
with an increased mass due to mass transfer. When the
initial binary period is of the order of one to few weeks,
the primary fills its Roche lobe for the first time during
shell hydrogen burning and Case B mass transfer takes
place. The primary loses most of its hydrogen envelope,
becomes a WR star and the secondary is an O star with
an increased mass. Case C mass transfer occurs when ini-
tial period is of the order of years. The primary fills its
Roche lobe during helium shell burning and mass trans-
fer takes place on the dynamical time scale. This scenario
is not likely for chosen systems, since some of the sec-
ondary stars in WR+O systems have been observed to ro-
tate faster than synchronously. This means that they have
accreted some matter which increased their spin angular
momentum.
We constructed a simple method to quickly estimate
the post-mass transfer parameters for a large number of
binary systems for a given accretion efficiency β. This al-
lows us to narrow the space of possible initial parameters
(primary mass, secondary mass and orbital period) that
allows the evolution into a specific observed WR+O sys-
tems.
We
considered binary systems with initial pri-
mary masses M1,in=25..100 M⊙ and secondaries masses
M2,in=25/1.7..100 M⊙ with an initial period of 3 days.
We assumed that the primary is transferring matter to
the secondary until it reaches the mass of its initial he-
lium core (Eq. 1).
Matter that is not accreted on the secondary leaves
the system with the specific angular momentum which
corresponds to the secondary's orbital angular momentum
(King et al. 2001), which is consistent with our approach
for mass loss from the binary system (cf. Sect. 4). Stellar
wind mass loss is neglected.
More massive initial primaries produce more massive
WR stars (helium cores) in general, but if the star is in
a binary system that goes through mass transfer during
hydrogen core burning of the primary (Case A), this de-
pends also on other parameters:
-If the initial period is longer, mass transfer starts later in
the primary evolution and the initial helium core of the
primary is more massive.
-If the initial mass ratio is further from unity, the mass
transfer rate from the primary reaches higher values and
the initial helium core mass is smaller.
4
Petrovic et al.: WR+O progenitors
Table 1. Basic parameters of selected WN+O SB2 binaries
WR number
HD number
spectral type
p (days)
e
q
a sin i (R⊙)
M sin3i (M⊙)
i (◦)
MWR (M⊙)
MO (M⊙)
WR 21
HD 90657
WN5+O4-6
8.2546 ± 0.0001
0.04 ± 0.03
0.52
37 ± 3
8.4
50 ± 4
19
37
WR 127
HD 186943
WN3+O9.5V
9.5550
0.07 ± 0.04
0.47
39 ± 6
9.3
55 ± 8
17
36
Notes:
a: all parameters from compilation of van der Hucht (2001), unless noted otherwise.
b: Demers et al. (2002).
WR153 b
HD 211853
WN6/WCE+O3-6I
6.6887
0 + 0
0.54
> 35.2
73
> 6
> 21
We initially want to restrict ourselves to systems that
undergo stable mass transfer,
i.e. avoid contact situa-
tions. Wellstein et al. (2001) found that the limiting ini-
tial mass ratio for conservative Case A binary system
is M1,in/M2,in∼1.55 and for conservative Case B sys-
tems ∼1.25. Since we allow non-conservative evolution,
we consider initial mass ratio q≤1.7 for Case A and q≤1.4
for Case B. The observed WR+O systems (HD 186943,
HD 90657 and HD 211853) all have very short orbital pe-
riods, between 6 and 10 days. Since, the net effect of the
Case A+Case AB, or Case B is a widening of the orbit
(if there is no contact), we have to assume that the initial
periods need to be shorter than, or approximatively equal
to the observed ones. We adopted a minimum initial or-
bital period of 3 days to avoid that the primary fills its
Roche lobe on the ZAMS.
We estimated the minimum initial helium core masses,
which are obtained by the earliest Case A systems, for sys-
tems with a mass ratio of M1,in/M2,in=1.7 and an initial
period of 3 days (M1,in>
∼41 M⊙) from the detailed evolu-
tionary models shown later in this paper (Sect. 5):
(1)
MWR,in = 0.24 ∗ M1,in + 0.27.
In this linear approximation, we neglected the influence of
the initial mass ratio on the initial WR mass. It is shown
in Sect. 5.2 that this dependence becomes important only
for initial mass ratios above q≃2.
For Case B binaries, the initial WR mass does not
depend on the initial period and the initial mass ratio
of the system, since during core hydrogen burning, the
primary evolves as a single star, without any interaction
with the secondary. We estimated the relation between
initial main sequence mass and initial WR mass as a linear
fit from the Case B binary systems with initial primaries
M1,in>
∼18 M⊙ (Wellstein & Langer 1999):
MWR,in = 0.53 ∗ M1,in − 4.92.
(2)
The minimum initial period for a system to evolve
through Case B mass transfer depends on the initial pri-
mary mass and the mass ratio. We can estimate, based
on the radii of the primaries at the end of the main se-
quence evolution, what would be the initial orbital separa-
tion necessary to avoid the primary filling its Roche lobe
before shell hydrogen burning. From Kepler's law follows
that the orbital separation is proportional to the mass ra-
1/3. Since
tio a∼q−1/3 and the initial primary mass a∼M1
q∼1 we can neglect this dependence and estimate the ini-
tial period for which the radius of the primary at the
end of MS is equal to its Roche radius. For this estimate
we do not take into account stellar wind that will widen
the orbit and decrease the masses. We conclude that the
Case B limiting initial orbital period for 40 M⊙ is ∼10
days, for 45 M⊙∼15 days and for 75 M⊙∼30 days. Since
the result of stable Case B mass transfer is widening of the
orbit, it follows that (stable mass transfer) Case B binary
systems can not be progenitors of the observed systems
HD 186943, HD 90657 and HD 211853 whose orbital peri-
ods are shorter than 10 days. However, WR star masses
resulting from Case B evolution are practically the same
as those from very late Case A evolution, which is still
considered in our analysis.
We calculate binary systems for early Case A (pin=3
days) and for late Case A (pin≈plimit).
The results are shown in Fig. 1 for four different accre-
tion efficiencies (β=0.0,0.1,0.5,1.0 respectively) for early
Case A evolution (p=3 days). Fig. 2 shows the results for
early Case A systems (p=3 days) for all values of β=0..1
and for Case B/late Case A, also for all β.
We notice from Fig. 1, that when the assumed β is
larger, the resulting WR+O systems lie further from the
line defined by q=0.5. The reason is clear: if the accre-
tion efficiency is higher, the secondary will become more
massive while the initial mass of the WR star stays the
same. Conservative evolution (Fig. 1d) produces WR+O
systems that have small mass ratios, q=1/5..1/6.
Petrovic et al.: WR+O progenitors
5
The orbital period of WR+O systems depends on their
initial orbital period, their initial mass ratio and on the
parameter β. If the initial period increases and there is
no contact during the evolution, the orbital period in the
WR+O stage will also increase. However, the orbital pe-
riod of WR+O systems will be shorter if the initial mass
ratio is larger. If the initial masses are very similar, the
primary will become less massive than the secondary very
early during the mass transfer, and afterward matter is
transfered from the less to the more massive star, which
results in a widening of the orbit. Conversely, the final pe-
riod is shorter for a larger difference in initial masses in
the binary system.
We can draw the following conclusions:
-The accretion efficiency during the major mass transfer
phase in the progenitor evolution of the three observed
WR+O binaries is small, i.e. β=0..0.1, as for larger β the
O stars during the WR+O phase are more massive and
the WR/O-mass rations smaller than observed. However,
we note that it is unlikely that the secondaries did not ac-
crete at all (β=0), since some O stars are found to rotate
faster than synchronously.
-The initial orbital period needs to be larger than ∼3 days,
to avoid contact at the beginning of hydrogen burning.
-The initial orbital period should be larger than ∼3 days,
in order to obtain massive enough WR stars.
-The initial orbital periods should be shorter than the ob-
served orbital periods in the three WR+O systems, i.e.
shorter than ∼10 days. This excludes Case B mass trans-
fer.
-While the initial mass ratio M1,in/M2,in should not be
too far from unity so contact is avoided, it should be close
to the contact limit, since this leads to the shortest orbital
periods and largest WR/O mass ratios in WR+O systems,
as needed for the three observed systems.
4. Numerical code and physical assumptions
We showed in Sect. 3 that we can roughly estimate the
parameters of the progenitor systems of observed WR+O
binaries HD 186943, HD 90657 and HD 211853. However,
detailed numerical models are required in order to verify
that the assumption of contact-free evolution can in fact
be justified. And finally, we want to check whether the
required mass and angular momentum loss can be repro-
duced by our detailed selfconsistent approach.
We are using a binary evolutionary code which was
originally developed by Braun (1998) on the basis of an im-
plicit hydrodynamic stellar evolution code for single stars
(Langer 1991, 1998). It calculates simultaneous evolution
of the two stellar components of a binary system in a cir-
cular orbit and the mass transfer within the Roche ap-
proximation (Kopal 1978). Mass loss from the Roche lobe
filling component through the first Lagrangian point is
given by Ritter (1988) as:
M = M0 exp(R − Rl)/Hp
(3)
Fig. 1. Masses of both components of post-Case A mass
transfer WR+O binary systems resulting from our sim-
ple approach, for initial primary masses in the range
25..100 M⊙ and an initial period of pin=3 days, for four
different assumed accretion efficiencies β (a-d). The solid
line represents a mass ratio of q=MWR/MO=0.5. For an
increasing β, the O stars in WR+O systems become more
massive and the WR/O-mass ratio decreases.
We conclude that if the considered three observed
WR+O binary systems evolved through a stable mass
transfer, a large amount of matter must have left the sys-
tem. On the other hand, since some of the secondary stars
in WR+O binaries have been observed to rotate faster
than synchronously (Massey 1981; Underhill et al. 1988),
a certain amount of accretion may be required.
Fig. 2 shows the resulting WR+O masses in Case A
and Case B (latest Case A) for accretion efficiency β=0..1.
If the primary star does not lose mass in a mass transfer
during core hydrogen burning (pin≥plimit), it will form a
more massive WR star, as we already explained. There
will be less mass to transfer from the primary to the sec-
ondary, and for fixed β the corresponding O star will be-
come less massive. However, since the observed periods of
HD 186943, HD 90657 and HD 211853 are shorter than 10
days and Case A+Case AB widens the binary orbit, the
initial orbital period should be shorter than observed, so
we can conclude roughly that pin is between 3 and 10 days.
6
Petrovic et al.: WR+O progenitors
given in Wellstein & Langer (1999) and Wellstein et al.
(2001). We use the OPAL Rosseland-mean opacities of
(Iglesias & Rogers 1996). For all models, a metallicity of
Z=0.02 is adopted. The abundance ratios of the isotopes
for a given element are chosen to have the solar meteoritic
abundance ratios according to Grevesse & Noels (1993).
The change of the orbital period (orbital angular momen-
tum loss) due to the mass transfer and stellar wind mass
loss is computed according to Podsiadlowski et al. (1992),
with the specific angular momentum of the stellar wind
material calculated by Brookshaw & Tavani (1993).
the
The
centrifugal
implemented
force
according
influence of
is
in the
rotating models
to
Kippenhahn & Thomas (1970). The stellar spin vec-
tors are assumed to be perpendicular to the orbital
plane. Synchronization due to tidal spin-orbit coupling
is included with a time scale given by Zahn (1977).
Rotationally enhanced mass loss is included as follows:
M / M (vrot = 0) = 1/(1 − Ω)ξ,
(5)
where ξ=0.43, Ω=vrot/vcrit and v2
crit=GM (1 − Γ)/R with
Γ=L/LEdd=κL/(4πcGM ) is Eddington factor, G is grav-
itational constant, M is mass, R radius, κ opacity, vrot ro-
tating velocity and vcrit critical rotational velocity (Langer
1998).
When the star approaches Ω=1, the mass loss rate is
increased according to the previous equation. However,
mass loss also causes a spin-down of the star and equilib-
rium mass loss rate Ωeq results (Langer 1998). If Ω > Ωeq,
the corresponding angular momentum loss is so large that
the star evolves away from the Ω-limit.
The transport of angular momentum through the stel-
lar interiour is formulated as a diffusive process:
∂t (cid:19)m
(cid:18) ∂ω
=
1
i (cid:18) ∂
∂m(cid:19)t(cid:20)(cid:0)4πr2ρ(cid:1)2
r (cid:18) ∂r
iν(cid:18) ∂ω
∂t(cid:19)m
−
2w
1
2
∂m(cid:19)t(cid:21)
dlni
dlnr
,
(6)
where ν is the turbulent viscosity and i is the specific
angular momentum of a shell at mass coordinate m.
The specific angular momentum of the accreted mat-
ter is determined by integrating the equation of motion of
a test particle in the Roche potential in case the accretion
stream impacts directly on the secondary star, and is as-
sumed Keplerian otherwise Wellstein (2001). Rotationally
induced mixing processes and angular momentum trans-
port through stellar interior are described by Heger et al.
(2000). Magnetic fields generated due to differential rota-
tion in the stellar interior (Spruit 2002) are not included
here (however, see Petrovic et al. 2004).
We calculated the evolution of the binary systems in
detail until Case AB mass transfer starts. Then we es-
timated the outcome of this mass transfer by assuming
that it ends when WR star has ∼5% of the hydrogen left
at the surface. For this purpose we calculate the Kelvin-
Helmholtz time scale of the primary:
tKH = 2 · 107M1
2/(L1Rl1)yr
(7)
Fig. 2. Masses of both components of post-Case A mass
transfer WR+O binary systems resulting from our sim-
ple approach, for initial primary masses in the range
25..100 M⊙ and for early (pin=3 days) Case A and late
Case A respective Case B evolution. The assumed accre-
tion efficiency is β=0..1. The solid line represents a mass
ratio of q=MWR/MO=0.5.
with M0=ρvsQ/√e, where Hp is the photospheric pressure
scale height, ρ is the density, vs the velocity of sound and
Q the effective cross-section of the stream through the first
Lagrangian point according to Meyer & Meyer-Hofmeister
(1983).
Stellar wind mass loss for O stars on the main se-
quence is calculated according to Kudritzki et al. (1989).
For hydrogen-poor stars (Xs<0.4) we assume mass loss
based on the empirical mass loss rates for Wolf-Rayet stars
derived by Hamann et al. (1995):
log( MWR/ M⊙ yr−1) = −11.95+1.5 log L/L⊙−2.85Xs.(4)
Since Hamann & Koesterke (1998) suggested that these
mass loss rates may be overestimated, we calculated evo-
lutionary models with mass loss rate given by Eq. 4 mul-
tiplied by factors 1/2, 1/3 and 1/6.
The treatment of a convection and a semiconvection
which is applied here is described in Langer (1991) and
Braun & Langer (1995). Changes in chemical composi-
tion are computed using a nuclear network including
pp chains, the CNO-cycle, and the major helium, car-
bon, neon and oxygen burning reactions. More details are
Petrovic et al.: WR+O progenitors
7
where M1, L1 and Rl1 are mass, luminosity and Roche
radius (in Solar units) of the primary star at the onset
of Case AB mass transfer. The mass transfer rate is then
assumed as:
Mtr = (M1 − MWR,in)/tKH
where MWR,in is the mass of the WR star that has a hy-
drogen surface abundance of 5%; all quantities are taken
at the beginning of the mass transfer. We calculate the
change of the orbital period orbit using constant value
of β=0.1 for non-rotating and β=0.0 for rotating mod-
els (Wellstein 2001). Matter that is not retained by the
secondary is assumed to leave the system with a specific
angular momentum which corresponds to the secondary's
orbital angular momentum (King et al. 2001).
(8)
5. Non-rotating models
We concluded in Sect. 3 that massive O+O binaries
can result in WR+O systems similar to observed the
(HD 186943, HD 90657 and HD 211853) if accretion effi-
ciency β is low. Since some O stars in WR+O binaries
have been observed to rotate faster than synchronously,
we concluded that β>0.0 and assumed a constant value
of β=0.1 in our detailed evolutionary models. We already
mentioned that the orbital periods of the observed sys-
tems are between 6 and 10 days. Since the net effect of
Case A+Case AB mass transfer is a widening of the orbit,
the initial periods should be shorter than the observed
ones, so we modelled binary systems with initial orbital
periods of 3 and 6 days.
We chose initial primary masses to be in the range
41..75 M⊙. The masses of the secondaries are chosen so
that the initial mass ratio (M1,in/M2,in) is q≈1.7-2.0. An
initial mass ratio of ≈1.55 is estimated to be the limit-
ing value for the occurrence of contact between the com-
ponents in Case A systems by Wellstein et al. (2001) for
conservative mass transfer. Contact occurs when the ac-
cretion time scale of the secondary ( M2,acc/M2) is much
longer than the thermal (Kelvin-Helmholtz) time scale of
the primary ( M =M1/tKH), so the secondary expands and
fills its Roche lobe. In our models, only 10% of matter
lost by the primary is accreted on the secondary star, so
it reaches hydrostatic equilibrium faster and expands less
than in the case of larger β. This is the reason why we
adopted a weaker condition for contact formation and cal-
culate models with mass ratios q≈1.7..2.0.
All modelled systems (except the ones that enter con-
tact) go through Case A and Case AB mass transfer.
Details of the evolution of all calculated binary systems
are given in Table 2. We discuss the details of the bi-
nary evolution taking the system number 11 as an exam-
ple. Fig. 3 shows the evolutionary tracks of the primary
and the secondary in the HR diagram until the onset of
Case AB mass transfer. This system begins its evolution
with the initial parameters M1,in=56 M⊙, M2,in=33 M⊙,
pin=6 days. Both stars are core hydrogen burning stars
(dashed line, Fig. 3), but since the primary is more mas-
sive, it evolves faster and fills its Roche lobe, so the system
enters Case A mass transfer (solid line, Fig. 3) ∼5.6·106
years after the beginning of core hydrogen burning. The
first phase of Case A is fast process and takes place on the
Kelvin-Helmholtz (thermal) time scale (∼3.1·104 years).
The primary loses matter quickly and continuously with
a high mass transfer rate ( M max
tr ∼3.1·10−3 M⊙ yr−1). In
order to retain hydrostatic equilibrium, the envelope ex-
pands, which requires energy and causes a decrease in
luminosity (Fig. 3). At the same time the secondary is
accreting matter and is expanding. Due to this, its lumi-
nosity increases and the effective temperature decreases
(Fig. 3). During fast phase of Case A mass transfer the
primary loses ∼19 M⊙ and the secondary accretes 1/10 of
that matter. After the fast process of mass transfer, the
primary is still burning hydrogen in its core and is still
expanding, so slow phase of Case A mass transfer takes
place on a nuclear time scale (0.46·106 years) with a mass
Mtr∼10−6 M⊙ yr−1. After this, the pri-
transfer rate of
mary is the less massive star, with decreased hydrogen sur-
face abundance. Stellar wind mass loss of the primary in-
creases when its surface becomes hydrogen poor (Xs<0.4).
At the end of core hydrogen burning the primary con-
tracts (effective temperature increases) and thus RLOF
stops (Fig. 3 dotted line). When the primary starts shell
hydrogen burning it expands (dash-dotted line, Fig. 3),
fills its Roche lobe and Case AB mass transfer starts.
Fig. 4 and Fig. 5 show the evolution of the interior of
the primary and the secondary until Case AB mass trans-
fer. The primary loses huge amounts of matter during fast
Case A mass transfer and its convective core becomes less
than a half of its original mass. At the same time, the sec-
ondary accretes matter from the primary and the heav-
ier elements are being relocated by thermohaline mixing.
In Fig. 6 and Fig. 7 we see the mass transfer rate and
the surface abundances of hydrogen, carbon, nitrogen and
oxygen.
During Case AB mass transfer the primary star loses
the major part of its hydrogen envelope. After Case AB
mass transfer, the primary is a helium core burning star
(WR) and the secondary is still a core hydrogen burning
O star. The masses of the modelled WR stars are in the
range from ∼8..18.5 M⊙ The orbital periods of the mod-
elled WR+O systems vary from ∼9.5 to ∼20 days, and
the mass ratios are between 0.33 and 0.53.
5.1. Relation between initial and WR mass
The initial mass of helium core of the primary in the bi-
nary system depends on a few parameters: initial primary
mass, initial period, initial mass ratio and stellar wind
mass loss rate. If the primary loses matter due to the mass
transfer or stellar wind during core hydrogen burning, it
will form a helium core that is less massive than if there
was no mass loss. If the initial period is very short, Case A
mass transfer will take place very early in the evolution
M max
Table 2. Non-rotating WR+O progenitor models for β=0.1. N is the number of the model, M1,in and M2,in are initial masses of the primary and the secondary,
pin is the initial orbital period and qin is the initial mass ratio of the binary system. tA is time when Case A mass transfer starts, ∆tf is the duration of the
fast phase of Case A mass transfer,
is the maximum mass transfer rate, ∆M1,f and ∆M2,f are mass loss of the primary and mass gain of the secondary
(respectively) during fast Case A, ∆ts is the duration of slow Case A mass transfer, ∆M1,s and ∆M2,s are mass loss of the primary and mass gain of the
secondary (respectively) during the slow Case A, pAB is the orbital period at the onset of Case AB, ∆M1,AB is the mass loss of the primary during Case AB
(mass gain of the secondary is 1/10 of this, see Sect. 4), MWR,5 is the WR mass when the hydrogen surface abundance is Xs=0.05, the WR mass at Xs ≤0.01
is given in brackets, MO is the mass of the corresponding O star, q is the mass ratio MWR/MO, and p is the orbital period of the WR+O system. The models
are computed with a stellar wind mass loss of Hamann/6, except ∗ Hamann/3, ∗∗ Hamann/2.
c indicates a contact phase that occurs for low masses due to a mass ratio too far from unity, for high masses due to the secondary expansion during slow phase
of Case A.
tr
N r
M1,in M2,in
pin
qin
tA
∆tf
M max
tr
∆M1,f ∆M2,f
∆ts
∆M1,s ∆M2,s
pAB ∆M1,AB MWR,5(1) MO
q
M⊙
M⊙
d
106yr
104yr M⊙/yr
M⊙
M⊙
106yr
M⊙
M⊙
M⊙
M⊙
s
r
o
t
i
n
e
g
o
r
p
O
+
R
W
:
.
l
a
t
e
c
i
v
o
r
t
e
P
8
N 1
N 2
N 3
N 4
N 5
N 6
N 7
N 8
N 9
N 10
N 11∗
N 12∗∗
N 13
N 14
41
41
41
41
41
41
41
45
56
56
56
56
65
75
20
20
20.5
24
24
27
30
27
33
33
33
33
37
45
3
6
3
3
6
3
3
3
3
6
6
6
3
3
2.05
2.05
2.00
1.71
1.71
2.8
3.6
2.8
2.8
3.6
1.52
2.75
1.37
1.67
1.70
1.70
1.70
1.70
1.76
1.67
2.7
2.5
1.9
2.8
2.8
2.8
1.6
1.3
c
3.9
2.2
3.1
4.3
5.8
6.7
3.6
5.0
5.8
5.8
5.8
3.2
4.2
p
d
d
−
M⊙
−
7.1
−
−
−
−
11.8(11.2)
22.5
0.52
12.6
−
−
−
−
−
−
5.4
18.82
1.85
0.39
0.97
0.06
5.9
18.0
21.13
2.11
−
−
−
2.85
8.17
7.7(7.2)
23.2
0.33
12.5
3.6
3.2
1.9
1.1
3.3
4.1
3.1
3.1
3.1
4.7
3.1
15.18
1.51
1.51
5.13
0.20
3.87
9.05
10.1(9.3)
26.4
0.38
13.5
17.31
1.72
0.42
1.88
0.09
8.92
7.53
12.1(11.4)
26.3
0.46
21.5
13.82
1.37
1.51
5.86
0.17
4.38
9.59
10.3(9.8)
29.1
0.35
16.6
12.60
1.24
1.51
6.72
0.08
5.20
9.76
10.5(10.0)
31.8
0.33
20.8
15.41
1.53
1.57
7.48
0.25
3.88
8.81
11.5(10.7)
29.4
0.39
12.0
17.2
1.70
1.86
15.66
0.44
4.07
7.14
13.6(12.7)
35.4
0.38
9.8
19.35
19.35
19.35
1.9
1.9
1.9
0.60
4.77
0.02
7.77
9.18
18.6(17.5)
35.1
0.53
15.2
0.46
3.63
0.06
7.91
7.15
18.6(17.2)
34.9
0.53
13.8
0.43
3.43
0.07
8.89
3.5
18.3(16.4)
34.5
0.53
12.1
18.81
1.87
18.57
1.79
c
c
−
−
−
−
−
−
−
−
16.2(14.8)
18.5(16.9)
−
−
−
−
−
−
Petrovic et al.: WR+O progenitors
9
Fig. 3. HR diagram of the initial system M1,in=56 M⊙,
M2,in=33 M⊙, pin=6 days. Both stars are core hydrogen
burning (dashed line) until Case A mass transfer starts
(solid line). The primary is losing mass and its luminosity
and effective temperature decrease. At the same time the
secondary is accreting matter and expanding, becoming
more luminous and cooler. After Case A mass transfer is
finished, the primary is losing mass by stellar wind and
contracting at the end of core hydrogen burning (dotted
line). After this the primary starts with shell hydrogen
burning and expands (dash-dotted line).
the internal
Fig. 4. The evolution of
structure of
the 56 M⊙ primary during the core hydrogen burning.
Convection is indicated with diagonal hatching and semi-
convection with crossed hatching. The hatched area at the
bottom indicates nuclear burning. The topmost solid line
corresponds to the surface of the star.
systems with an initial mass ratio of q≈1.7 and an initial
period p = 3 days (Table 2: N 4, 8, 9, 13, 14). Large 'star'
symbols represent WR stars with 5% of hydrogen at the
surface and small symbols indicate WR stars that have a
hydrogen surface abundance of less than 1%.
of the primary, so the star will not have time to develop
a larger core before it starts losing mass due to Roche
lobe overflow. If the initial mass ratio (q=M1,in/M2,in)
increases, the mass transfer rate from the primary star in-
creases too and this results in less massive primaries that
will evolve into less massive WR stars.
In our models the primary starts losing mass by stel-
lar wind as WR star when its hydrogen surface abundance
goes below Xs=0.4. However, the observed WR stars in
HD 186943, HD 90657 and HD 211853 do not have obvi-
ous hydrogen on the surface, so we assume that these WR
stars are the result of Case AB mass transfer, with a hy-
drogen surface abundance of Xs≈0.05. We also calculated
the corresponding WR masses with Xs≤0.01. We plotted
in Fig. 8 the initial WR masses (Xs=0.05 and Xs ≤0.01)
versus the initial primary (progenitor) masses. With 'star'
symbols we indicated WR stars that originate from binary
We derive a relation between the initial primary mass
and the initial WR mass (derived as a linear fit) for p=3
days and q≈1.7, (Xs=0.05):
MWR = 0.24 ∗ M1,in + 0.27.
We use this relation to estimate the initial parameters of
the possible progenitors of the observed WR+O binary
systems, as already explained in Sect. 3. In the same way,
the relation between the initial primary mass and the ini-
tial WR mass (Xs<0.01) for the same systems is:
MWR = 0.22 ∗ M1,in + 0.56.
(10)
(9)
We also show in Fig. 8 the initial WR masses (Xs=0.05
for binary systems N 5, 10, 11, 12, Table 2) for an ini-
tial mass ratio of ∼1.7 and an initial orbital period of 6
days (diamond symbols). We notice that the resulting WR
masses are higher than the ones that come out from sys-
tems with an initial orbital period of 3 days (see Sect. 5.3).
10
Petrovic et al.: WR+O progenitors
Fig. 5. The evolution of the internal structure of the
33 M⊙ secondary during core hydrogen burning of the
primary. Convection is indicated with diagonal hatching,
semiconvection with crossed hatching and thermohaline
mixing with straight crossed hatching. The hatched area
at the bottom indicates nuclear burning. The topmost
solid line corresponds to the surface of the star.
Different 'diamond' symbols for the initial primary 56 M⊙
are for different mass loss rates (see Sect. 5.4). Triangle
symbols in Fig. 8 show the initial WR masses for con-
stant initial primary mass, M1,in=41 M⊙, but for different
initial mass ratios (see Sect. 5.2)
Note that the WR masses that are the result of
early Case A progenitor evolution are significantly lower
than ones that are the result of Case B evolution
(Wellstein & Langer 1999), because of the mass transfer
from the primary during the core hydrogen burning phase.
5.2. Influence of the initial mass ratio on the WR mass
and orbital period
During the mass transfer phase, the mass transfer rate in-
creases roughly until the masses of both components are
equal. The maximum mass transfer rate during Case A
increases with the increase of the initial mass ratio
(M1,in/M2,in) and the resulting WR star is less massive.
To analyse the influence of the initial mass ratio on the
evolution of the binary system, we compared systems with
Fig. 6. Upper plot: Mass transfer rate during Case A
mass transfer in the binary system with M1,in=56 M⊙,
M2,in=33 M⊙, pin=6 days. Lower plot: The hydrogen sur-
face abundance of the primary (solid line) is decreasing
during mass transfer and further due to stellar wind mass
loss. The secondary (dashed line) recovered its original
surface hydrogen abundance through thermohaline mix-
ing. The primary starts losing mass with WR stellar wind
mass loss when its hydrogen surface abundance falls be-
neath Xs=0.4, represented by the dotted line.
an initial primary mass of M1,in=41 M⊙, an initial or-
bital period of pin=3 days for five different initial mass
ratios: 2.05, 2.00, 1.71, 1.52 and 1.37. (Table 2 N 1, 3,
4, 6, 7, Fig. 9). The system with qin=2.05 enters con-
tact during fast Case A mass transfer. The mass transfer
rate in this case is very high ( M≈6·10−2 M⊙ yr−1), the
secondary expands, fills its Roche lobe and the system
enters a contact phase. The system with an initial mass
ratio of qin=2.00 loses ∼21M⊙ during the fast phase of
Case A. The maximum mass transfer rate of this system
is M≈1.8·10−2 M⊙ yr−1. The helium surface abundance of
the primary after this mass transfer is 65%, so the primary
shrinks, loses mass through a WR stellar wind and there
is no slow phase of Case A mass transfer (R1< 9 R⊙). For
the other three models q =1.71,1.52,1.37, the primaries
lose less mass (∼15,14,13 M⊙ respectively) during the fast
phase of Case A mass transfer. The helium surface abun-
dances in these systems after fast Case A mass transfer
Petrovic et al.: WR+O progenitors
11
Fig. 7. Surface abundance of carbon (solid line), nitro-
gen (dotted line) and oxygen (dashed line) of the primary
(upper plot) and the secondary (lower plot) in the system
with M1,in=56 M⊙, M2,in=33 M⊙, pin=6 days.
are ∼30-35%. The primaries expand (R≈12-15 R⊙) on a
nuclear time scale and transfer mass to the secondaries
(slow phase of Case A).
We can conclude the following: First, if the initial mass
ratio is larger, the mass transfer rate from the primary
during fast phase Case A mass transfer is higher. Second,
if the mass transfer rate is higher, the helium surface abun-
dance of the star increases faster and if it reaches ≈58%,
the primary starts losing mass with a higher (WR) mass
loss rate and slow Case A mass transfer can be avoided.
We also show in Fig. 9 (lower plot) how the period
changes during Case A mass transfer for binary systems
N 3, 4, 6, 7. Roughly, when the mass is transfered from
the more to the less massive star, the binary orbit shrinks,
and when the mass is transfered from the less to the more
massive star, the orbit widens. If the initial period is close
to unity, the absolute difference between stellar masses
is small, and more mass is transfered from the less to the
more massive star during the evolution of the system. This
results in a longer final period after Case A mass transfer.
Systems with initial mass ratios of 2.00, 1.71, 1.52 and
1.37 enter Case AB mass transfer with orbital periods of
2.9, 3.9, 4.4 and 5.2 days respectively. However, the final
period is also (more significantly) influenced by the stellar
Fig. 8. Initial WR mass as a function of initial (pro-
genitor) mass. Large and small symbols indicate WR
stars with hydrogen surface abundance of Xs=0.05 and
Xs ≤0.01, respectively. Systems with an initial orbital pe-
riod of pin=3 days and a mass ratio of q ∼1.7 are indicated
with star symbols, systems with an initial period of 6 days
and q ∼1.7 with diamond symbols, systems with an initial
primary 41 M⊙ and initial period of 3 days with triangle
symbols. The dashed line represents a linear fit for sys-
tems with an initial period of 3 days and with Xs=0.05,
and the dotted line represents linear fit for systems with
an initial period of 3 days and Xs ≤0.01.
wind mass loss rate and the amount of matter lost from the
primary during Case AB mass transfer (see Section 5.4).
5.3. Influence of the initial period on the WR mass
Depending on the initial orbital period of a binary sys-
tem, Case A mass transfer phase will start earlier or later
in the evolution. If the period is larger, the primary will
develop a larger core, before it starts transferring mass
onto the secondary, and the resulting WR star will be
more massive. To investigate the influence of the initial
period, we compare binary systems 41 M⊙+24 M⊙ and
56 M⊙+33 M⊙ with p=3 days and p=6 days (Table 2: N
4, 5, 9, 10).
If the initial orbital period increases for 3 days, a 41 M⊙
star will enter Case A mass transfer ∼8·105 yr later and
12
Petrovic et al.: WR+O progenitors
Fig. 9. Mass transfer rate (upper plot) and orbital period
(lower plot) during Case A mass transfer as a function
of the change of the primary mass for systems with the
initial primary M1,in=41 M⊙, initial orbital period pin=3
days and four different initial mass ratios: 2.00 (dotted
line), 1.71 (solid line), 1.52 (dashed line) and 1.37 (dash-
dotted line). (Table 2 N 3, 4, 6, 7)
transfer
Fig. 10. Mass
rate during Case A mass
transfer for systems 41 M⊙+24 M⊙ (upper plot) and
56 M⊙+33 M⊙ (lower plot) and an initial orbital period
of pin=3 days(solid line) and pin=6 days (dashed line).
Case A mass transfer starts later in initially wider binary
systems, the primary has more time to increase mass of
its core and the initial WR star is more massive (Table 2:
N 4, 5, 9, 10).
a 56 M⊙ star ∼9·105 yr later. So, there are two things
to point out: first, the more massive star (56 M⊙) evolves
faster, and second, a 3 days longer initial period postpones
Case A mass transfer, for this star, by about 105 yr more
than for a 41 M⊙ star. The net effect is a more significant
increase of the convective core (i.e. initial helium core, i.e
initial WR mass), for more massive star, due to the initial
orbit widening.
5.4. Influence of WR mass loss rate on masses and
final period
In our models, we assume that when the star has less than
40% of hydrogen at the surface, it starts losing mass ac-
cording to the Hamann et al. (1995) WR mass loss rate,
multiplied by factors: 1/6, 1/3 and 1/2 (Table 2: N 10, 11,
12). If the primary is losing more mass by stellar wind dur-
ing core hydrogen burning, it will develop a less massive
helium core. At the same time there will be less matter to
be transfered during Roche lobe overflow, so the secondary
will accrete less.
We show in Fig. 11 the influence of the stellar wind
mass loss (Plot c) on the primary mass (Plot a) and
mass transfer rate (Plot b) of systems with M1,in=56 M⊙,
M2,in=33 M⊙, pin=3 and three different stellar wind mass
loss rates (from Xs≤0.4): 1/6 (solid line), 1/3 (dotted
line) and 1/2 (dashed line) of the mass loss proposed by
Hamann et al. (1995). We notice that for higher mass loss
rates, the slow phase of Case A stops earlier, due to the
decrease of the stellar radius. The orbit is widening due
to the stellar wind mass loss and the final period increases
with the increasing mass loss rate. However, the orbit is
more significantly widening during Case AB mass trans-
fer. The more mass there is to transfer from the primary to
the secondary during Case AB mass transfer, the larger
the final orbital period. So, if the stellar wind removes
most of the hydrogen envelope of the primary, there will
be less mass to transfer during Case AB and the net effect
of a higher mass loss rate is a shorter orbital period of the
WR+O system.
Petrovic et al.: WR+O progenitors
13
Fig. 11. Primary mass (first plot), mass transfer rate (sec-
ond plot) and stellar wind mass loss rate from the primary
(third plot) until the onset of Case AB mass transfer for
the system M1,in=56 M⊙, M2,in=33 M⊙, pin=6 days and
three different stellar wind mass loss rates (from Xs ≤0.4):
1/6 (solid line), 1/3 (dotted line) and 1/2 (dashed line) of
mass loss rate proposed by Hamann et al. (1995) (Table 2:
N 10, 11, 12).
6. Rotating models
When mass transfer in a binary system starts, the pri-
mary loses matter through the first Lagrangian point (L1).
This matter carries a certain angular momentum that will
be transfered to the secondary. If there is an accretion
disk, the angular momentum of the transfered matter is
assumed to be Keplerian. If there is a direct impact accre-
tion, like in our models, we calculate the angular momen-
tum following a test particle moving through L1. This an-
gular momentum spins up the top layers of the secondary
star, and angular momentum is transfered further into the
star due to rotationally induced mixing processes. Every
time the secondary spins up to close to critical rotation it
starts losing more mass due to the influence of centrifugal
force (Eq. 5). High mass loss decreases the net accretion
efficiency and also removes angular momentum from the
secondary star. The secondary star is also spun down by
tidal forces that tend to synchronize it with the orbital
motion. Wellstein (2001) investigated these processes in
binary systems with initial mass ratios close to unity and
Fig. 12. HR diagram of the initial system M1,in=56 M⊙,
M2,in=33 M⊙, pin=6 days with rotation. Both stars are
core hydrogen burning (dashed line) until Case A mass
transfer starts (solid line). The primary is losing mass and
its luminosity decreases. At the same time the secondary
is accreting matter and expanding, becoming more lumi-
nous. After Case A mass transfer is finished, the primary
is losing mass by stellar wind and contracting at the end
of core hydrogen burning (dotted line). After this the pri-
mary starts shell hydrogen burning and expands (dash-
dotted line).
concluded that the accretion efficiency does not decrease
significantly for Case A mass transfer, but in the Case B
the parameter β can be significantly decreased by rota-
tion. We present Case A rotating models with larger mass
ratio q=M1,in/M2,in=1.7..2 and find that accretion can be
significantly decreased during Case A mass transfer. The
reason is the following: if the initial mass ratio increases,
so does the maximum mass transfer rate ( Mmtr increases
roughly until the masses in binary system are equal). If
there is more mass transfered from the primary to the sec-
ondary, the rotational velocity of the secondary is higher
as well as its mass loss, which leads to a smaller accretion
efficiency.
We compare the evolution of non-rotating and ro-
tating binary systems on the example M1,in=56 M⊙
M2,in=33 M⊙, an initial orbital period of p=6 days, and
Hamann/3 WR mass loss stellar wind rate (Table 3: N
Table 3. Rotating WR+O progenitor models. N is the number of the model, M1,in and M2,in are initial masses of the primary and the secondary, pin is the
initial orbital period and qin is the initial mass ratio of the binary system. tA is the time when Case A mass transfer starts, ∆tf is the duration of the the
is the maximum mass transfer rate, ∆M1,f and ∆M2,f are mass loss of the primary and mass gain of the secondary
fast phase of Case A mass transfer,
(respectively) during the fast Case A, ∆ts is the duration of slow Case A mass transfer, ∆M1,s and ∆M2,s are mass loss of the primary and mass gain of the
secondary (respectively) during the slow Case A, pAB is the orbital period at the onset of Case AB, ∆M1,AB is the mass loss of the primary during Case AB
(mass gain of the secondary is 1/10 of this, see Sect. 4), MWR,5 is the WR mass when the hydrogen surface abundance is Xs = 0.05, MO is the mass of the
corresponding O star, q is the mass ratio MWR/MO, p is the orbital period of the WR+O system and MWR,1 is WR mass with Xs/le0.01. The models are
computed with a stellar wind mass loss of Hamann/6 except :∗ Hamann/3, ∗∗ Hamann/2.
c indicates a contact phase
M max
tr
N r
M1,in M2,in
pin
qin
tA
∆tf
M max
tr
∆M1,f , ∆M2,f
∆ts
∆M1,s, ∆M2,s
pAB ∆M1,AB MWR,5(1), MO
q
p
d
M⊙
M⊙
R1
R2∗∗
R3
R4
R5
R6∗
R7∗∗
R8
R9∗
41
41
41
41
56
56
56
60
60
20
20
24
24
33
33
33
35
35
d
6
6
3
6
6
6
6
6
6
106yr
104yr M⊙/yr
M⊙
106yr
M⊙
d
M⊙
M⊙
2.05
2.05
1.71
1.71
1.70
1.70
1.70
1.71
1.71
3.4
3.4
2.6
3.4
2.4
2.4
2.4
2.3
2.3
1.5
1.5
1.5
2.6
3.7
3.7
3.7
2.2
2.2
6.5
6.5
3.9
3.8
3.2
3.2
3.2
3.0
3.0
18.67, 3.33
0.58
2.38(1.37), 0.81
3.97
6.61
11.0(10.2), 23.98
0.46
9.78
18.67, 3.33
0.10
0.32(0.11), 0.20
4.77
2.96
10.4(9.0), 23.20
0.45
7.92
15.47, 5.04
1.34
9.38(1.00), 7.54
4.27
6.32
8.2(7.6), 36.17
0.23
17.86
17.75, 4.06
0.68
2.55(0.9), 1.53
5.66
7.25
11.2(10.5), 29.27
0.38
16.42
19.32, 2.91
0.98
11.93(4.13), 6.98
6.09
4.88
14.9(13.6), 42.09
0.35
11.59
19.32, 2.91
0.90
10.93(6.42), 3.91
6.64
19.32, 2.91
0.45
3.24(0.65), 2.27
8.43
1.8
0.0
14.8(12.8), 38.99
0.38
8.53
11.2(8.8), 37.04
0.30
8.43
19.97, 3.98
0.92
12.32(4.20), 7.21
6.58
5.27
15.7(14.6), 45.13
0.35
12.75
19.97, 3.98
0.84
11.42(6.26), 4.58
7.64
0.0
14.9(12.2), 42.43
0.35
7.64
s
r
o
t
i
n
e
g
o
r
p
O
+
R
W
:
.
l
a
t
e
c
i
v
o
r
t
e
P
4
1
Petrovic et al.: WR+O progenitors
15
Fig. 13. The evolution of the internal structure of the
rotating 56 M⊙ primary during core hydrogen burning.
Convection is indicated with diagonal hatching and semi-
convection with crossed hatching. The hatched area at the
bottom indicates nuclear burning. Gray shaded areas rep-
resent regions with rotationally induced mixing (intensity
is indicated with different shades, the darker the colour,
the stronger rotational mixing). The topmost solid line
corresponds to the surface of the star.
Fig. 14. The evolution of the internal structure of the ro-
tating 33 M⊙ secondary during core hydrogen burning of
the primary. Convection is indicated with diagonal hatch-
ing, semiconvection with crossed hatching and thermoha-
line mixing with straight crossed hatching. The hatched
area at the bottom indicates nuclear burning. Gray shaded
areas represent regions with rotationally induced mixing
(intensity is indicated with different shades, the darker the
colour, the stronger rotational mixing). The topmost solid
line corresponds to the surface of the star.
6). The rotating binary system is synchronized as it starts
core hydrogen burning and it stays that way until mass
transfer starts. The radius of the primary increases during
the main sequence phase (from ∼10 to ∼25 R⊙, Fig. 19b),
but the rotation of the primary stays synchronized with
the orbital period. This is why the rotational velocity of
the primary also increases from ∼100 to ∼200 km s−1
(Fig. 19d). The radius of the rotating primary increases
faster than the radius of the non-rotating primary due to
the influence of the centrifugal force. The result is that
Case A mass transfer starts earlier for the rotating binary
system (Xc,non≈80%) then for the corresponding nonro-
tating one(Xc,rot≈71%, Fig. 15).
When the fast phase of Case A starts, the secondary
spins up (Fig. 20d) and stellar wind mass loss rapidly in-
creases ( Msw∼10−3 M⊙ yr−1, Fig. 20c). The accretion effi-
ciency during this phase in the rotating system is β=0.15
(Table 3). We see in Fig. 15 that the orbital period af-
ter Case A mass transfer of the rotating binary system is
shorter than for the non-rotating system (4.5 compared
with 6.6 days). The orbital angular momentum of the bi-
nary is changing due to mass transfer, mass loss from the
system and spin-orbit coupling. The rotating binary sys-
tem loses more angular momentum and the final orbital
period is shorter than in the corresponding non-rotating
system. Angular momentum loss in our systems is calcu-
lated according to Podsiadlowski et al. (1992) as already
mentioned in Sect 4, and parameter α that determines the
efficiency of angular momentum loss is calculated accord-
ing to Brookshaw & Tavani (1993). It increases with the
mass ratio M2/M1 and the ratio between the secondary ra-
dius and its Roche radius R2/Rl2. In rotating system the
secondary accretes slightly more matter ( ¯β=0.15) com-
pared to β=0.1 in non-rotating systems, so the mass ratio
M2/M1 is larger in the rotating system. Second, the sec-
ondary is spinning fast and its radius is larger than in the
non-rotating case, and so is the ratio R2/Rl2. The result
16
Petrovic et al.: WR+O progenitors
Fig. 15. Upper plot: The mass transfer rate during Case A
mass transfer in the binary systems with M1,in=56 M⊙,
M2,in=33 M⊙, pin=6 days with (solid line) and without
rotation (dotted line). Lower plot: Orbital period evolu-
tion in rotating and non-rotating system.
is that the angular momentum is more efficiently removed
from the system in the rotating binary system. After the
fast phase of Case A mass transfer, the two primaries, non-
rotating and rotating, have almost the same mass ∼34 M⊙
and helium surface abundance Ys∼44%. However, since
the orbital periods are different, so are the radii of the
primaries (∼18 R⊙ for the rotating and ∼23 R⊙ for the
non-rotating case).
When the fast phase of Case A is finished, the non-
rotating primary has still ∼20% of hydrogen to burn
(∼7·105 yr), and the rotating primary has ∼10% more
than that (∼1.2·106 yr). When the surface hydrogen abun-
dance is less than 40%, the primaries start losing mass as
WR stars, i.e., their stellar wind mass loss rate increases.
Since the rotating primary has more time to spend on
the main sequence, it also has more time to lose mass
by WR stellar wind mass loss (7.2·105 yr compared with
2.5·105 yr for non-rotating system).The result is that the
non-rotating primary enters Case AB mass transfer as a
∼26 M⊙ star with Ys=0.75, while the rotating one is a
∼17 M⊙ star with Ys=0.90. Clearly, the rotating primary
has less hydrogen in its envelope, i.e. less mass to trans-
fer to the secondary during Case AB mass transfer, and
Fig. 16. The hydrogen surface abundance (solid line) in
the primary in system with M1,in=56 M⊙, M2,in=33 M⊙,
pin=6 days is decreasing during mass transfer and further
due to stellar wind mass loss. The secondary (dashed line)
decreases its hydrogen surface abundance due to mass
transfer. The dotted line indicates a hydrogen abundance
of 0.4, where the primary starts losing mass with a WR
stellar wind.
the orbit widens less than in the non-rotating system. We
can draw the conclusion that if rotation is included in our
calculations, the initial WR mass is smaller and the or-
bital period of the WR+O system is shorter than in the
corresponding non-rotating system (Table 4).
We present in Fig. 12 the evolutionary tracks of
the rotating primary and secondary in the HR dia-
gram. Both stars are core hydrogen burning stars (dashed
line, Fig. 12), but since the primary is more massive, it
evolves faster and fills its Roche lobe, so the system en-
ters Case A mass transfer (solid line, Fig. 12). The pri-
mary loses matter quickly with a high mass transfer rate
( M max
tr ≈3.2·10−3 M⊙ yr−1) and its luminosity decreases
(Fig. 12). At the same time the secondary accretes matter
and its luminosity increases, but due to change in rota-
tional velocity (Fig. 20d) its radius and effective temper-
ature are changing as well (Fig. 12d, Fig. 20a,b). During
fast Case A mass transfer the primary lost ∼19 M⊙ and
the secondary accreted 15% of that matter. After the fast
mass transfer, the primary is still burning hydrogen in its
core and is still expanding, so slow Case A mass transfer
Petrovic et al.: WR+O progenitors
17
Fig. 17. Surface abundance of carbon (solid line), nitro-
gen (dotted line) and oxygen (dashed line) in the primary
(upper plot) and the secondary (lower plot), in the sys-
tem with M1,in=56 M⊙, M2,in=33 M⊙, pin=6 days. The
secondary abundances are changed due to mass transfer
of matter from the primary, thermohaline mixing and ro-
tational mixing.
Fig. 18. The orbital angular momentum (upper plot) of
the non-rotating (dotted line) and the rotating (solid line)
binary systems with M1,in=56 M⊙, M2,in=33 M⊙, pin=6
days, decreases rapidly due to mass loss from the system
during fast Case A mass transfer and then further due to
stellar wind mass loss. Spin period (lower plot) of the pri-
mary (dashed line) and the secondary (dash-dotted line)
in the above mentioned rotating binary system.
takes place. After the primary starts losing mass with a
WR stellar wind mass loss rate (Xs<0.4) its radius will
decrease and the slow phase of Case A stops (Fig. 19c).
However, the primary continues expanding on the nuclear
time scale (Fig. 19b) and it fills its Roche lobe once again
(Fig. 15, upper plot). At the end of core hydrogen burn-
ing the primary contracts (effective temperature increases)
and thus RLOF stops. This phase is presented in Fig. 12
with a dotted line. When hydrogen starts burning in a
shell, the primary star expands (dash-dotted line, Fig. 12),
fills its Roche lobe and Case AB mass transfer starts.
The initial helium core masses are 18.6 M⊙ for the
non-rotating and 14.8 M⊙ for the rotating primary. When
Case AB mass transfer starts, the orbital periods are 7.9 d
and 6.6 d for the non-rotating and the rotating system re-
spectively (Fig. 15, lower plot). The non-rotating primary
loses ∼7 M⊙ and the rotating one ∼2 M⊙ during Case AB.
When there is more mass to be transfered from the less to
the more massive star in a binary system, the orbit widens
more and the final orbital period is longer.
Fig. 13 and Fig. 14 show the structure of the primary
and the secondary before Case AB mass transfer. The pri-
mary loses large amounts of matter during the fast phase
of Case A mass transfer (∼20 M⊙), and its convective core
becomes less than half of its original mass. At the same
time, the secondary accretes matter from the primary and
the heavier elements are being relocated by thermohaline
mixing. Fig. 16 and Fig. 17 show surface abundances of
the primary and the secondary. The secondary is accret-
ing material from the primary and its surface abundances
change due to this, but also due to thermohaline and ro-
tational mixing.
Fig. 18 shows the orbital angular momentum of the
system and the spin periods of both components. The or-
bital angular momentum of the system decreases rapidly
due to mass loss from the system during fast Case A
mass transfer, and then further due to stellar wind mass
loss. The primary slows down rapidly during fast Case A
and further due to stellar wind mass loss. The secondary
spins up due to the accretion from the primary during fast
18
Petrovic et al.: WR+O progenitors
Fig. 19. Effective temperature (plot a), stellar radius (plot
b), stellar wind mass loss rate (plot c) and rotational ve-
locity (plot d) of the primary star in the non-rotating
(dotted line) and rotating (solid line) binary system with
M1,in=56 M⊙, M2,in=33 M⊙, pin=6 days.
Fig. 20. Effective temperature (plot a), stellar radius (plot
b), stellar wind mass loss rate (plot c) and rotational ve-
locity (plot d) of the secondary star in the non-rotating
(dotted line) and rotating (solid line) binary system with
M1,in=56 M⊙, M2,in=33 M⊙, pin=6 days.
Case A mass transfer and then slows down due to stellar
wind mass loss. It spins up again during slow Case A mass
transfer.
The masses of modelled WR stars are in the range
from ∼11 M⊙ to ∼15.7 M⊙. Period of modelled WR+O
systems vary from ∼7.6 to ∼12.7 days and mass ratios
are between 0.35 and 0.46 (Table 3).
6.1. Influence of rotation on the accretion efficiency
We show in Table 5 average accretion efficiencies of rotat-
ing binary systems during different mass transfer phases,
and total average values with and without stellar wind
mass loss from the primary included. During fast Case A
mass transfer, the primary stars are losing matter with
very high mass transfer rates (3..6.5·10−3 M⊙ yr−1). The
angular momentum of surface layers in the secondary
increases fast, they spin up to close to the critical ro-
tation and start losing mass with high mass loss rate
(∼10−3 M⊙). The average accretion efficiency during fast
Case A in our models is 15-20%. Since this phase takes
place on the thermal time scale, stellar wind mass loss
from the primary is negligible during this phase.
Table 4. Comparison of resulting WR masses and or-
bital periods from non-rotating and rotating binary sys-
tems with the same initial parameters.M1,in, M2,in are ini-
tial primary and secondary mass, pin is the initial orbital
period, MWR,5, MWR,1 are WR masses at Xs = 0.05 and
Xs ≤ 0.01 respectively and p is the orbital period in the
initial WR+O system where the hydrogen surface abun-
dance of WR star is Xs = 0.05. Systems are modelled with
WR stellar wind mass loss H/6 except ∗ which are done
with H/3, R indicates rotating models.
M1,in + M2,in
[ M⊙]
pin
[d]
MWR,5
[ M⊙]
MWR,1
[ M⊙]
41 + 20
41 + 20R
41 + 24
41 + 24R
56 + 33∗
56 + 33∗,R
6
6
6
6
6
6
11.8
11.0
12.1
11.2
18.6
14.9
11.2
10.2
11.4
10.5
17.5
13.6
p
[d]
12.6
9.8
21.5
16.4
13.8
8.5
Petrovic et al.: WR+O progenitors
19
Slow Case A mass transfer takes place on the nuclear
time scale. The primary stars start losing their mass due
to a WR stellar wind when their surfaces become hydro-
gen deficient (Xs<0.4). The WR stellar wind mass loss
M ∼10−5 M⊙ yr−1, and we have
rates are of the order of :
to take into account stellar wind mass loss of the primary
during slow Case A. We calculate the mass loss of the
primary only due to mass transfer and total mass loss in-
cluding stellar wind mass loss, and the two corresponding
average accretion efficiencies. If we calculate ¯βs only for
mass transfer, we notice that the slow Case A is almost
a conservative process. The average mass transfer rates
are ∼10−6 M⊙ yr−1 and the secondary stars are able to
accrete almost everything without spinning up to critical
rotation.
Fig. 21 shows how mass transfer rate, accretion rate
and β change in the rotating model with 56 M⊙+33 M⊙,
pin=6 days (WR mass loss Hamann/3) depending on the
amount of matter lost by the primary. We also see in this
figure the mass transfer rate from the primary in the non-
rotating case. We can notice in the upper plot, what we
previously discussed, that during most of the fast Case A
mass transfer, the mass accretion rate of the secondary is
about one order of magnitude lower than the mass loss
rate of the primary. The primary loses ∼19.3 M⊙ during
the fast phase and the secondary gains ∼2.9 M⊙, which
means that on average ∼15% of the mass has been ac-
creted. However, the mass loss of the primary due to mass
transfer during the slow phase is ∼4.5 M⊙, and the sec-
ondary accretes ∼3.9 M⊙ which means that ¯βs ∼0.87. If
we take into account stellar wind mass loss of the pri-
mary stars, the average accretion efficiencies are lower.
For example, the total mass loss of the primary during
slow Case A mass transfer, including the stellar wind,
in the previous example is ∼10.9 M⊙, which means that
¯βw
s ≈0.36. We neglected accretion during Case AB mass
transfer since Wellstein (2001) showed that it is ineffi-
cient, and since the primary stars in the modelled systems
have relatively low mass hydrogen envelopes, and masses
of secondary stars will not significantly change due to this
mass transfer. (However, let us not forget that even the
accretion of very small amounts of matter can be impor-
tant for spinning up the secondary's surface layers and
making it rotate faster than synchronously in a WR+O
binary system.) Also, since this mass transfer takes place
on the thermal time scale, stellar wind mass loss can be
neglected.
Finally, we can estimate the total mass loss from the
binary systems including stellar wind, or only due to mass
transfer, and calculate corresponding values of β. In the
binary systems we modelled, the primary stars lose be-
tween 30 M⊙ and 45 M⊙ due to mass transfer and stellar
wind, until they ignite helium in their core. The amount
of lost mass increases with initial mass. At the same time
the secondaries accrete 3..10 M⊙. This means that in most
cases 80..90% of the mass lost by the primary leaves the
binary system. On the other hand, the primary stars lose
∼20..30 M⊙ only due to mass transfer, so the average ac-
Table 5. Mass loss from binary systems. N is number of
the model corresponding to Table 3. ¯βfast is the average
accretion efficiency of the secondary during the fast phase
of Case A mass transfer. ¯βslow is the accretion efficiency
of the secondary during the slow phase of Case A mass
transfer taking into account matter lost by the primary
only due to the mass transfer. ¯βwind
slow is the average accre-
tion efficiency of the secondary during the slow phase of
Case A mass transfer taking into account matter lost by
the primary due to the mass transfer and stellar wind. ¯β
is the average accretion efficiency of the secondary during
the progenitor evolution of WR+O binary system taking
into account matter lost by the primary only due to the
mass transfer and ¯βwind
taking also into account stellar
mtr
wind mass loss of the primary.
N r
R1
R2
R3
R4
R5
R6
R7
R8
R9
¯βfast
0.18
0.18
0.33
0.23
0.15
0.15
0.15
0.20
0.20
¯βslow
¯βwind
slow
0.80
0.95
0.90
0.94
0.90
0.87
0.88
0.88
0.89
0.35
0.65
0.80
0.60
0.58
0.36
0.70
0.58
0.40
¯β
0.15
0.15
0.40
0.20
0.28
0.23
0.18
0.30
0.30
¯βwind
0.13
0.10
0.37
0.18
0.22
0.14
0.09
0.23
0.16
cretion of secondary stars in our models is between 15 and
30%.
7. Comparison with observations
Our rotating models give generally similar results as our
non-rotating models for β = 0.1.
The rotating binary systems R6 (56 M⊙+33 M⊙,
p=6 days, Hamann/3 WR mass loss) and R1 and R2
(41 M⊙+20 M⊙, p=6 days) agree quite well with the ob-
served systems HD 186943 and HD 90657, as well as the
non-rotating systems N11 and N12 (56 M⊙+33 M⊙, p=6
days; WR mass loss rate Hamann/2 and Hamann/3).
The system R6 evolves into a WR+O configuration with
15 M⊙+39 M⊙ and p=8.5 days. I.e., its masses and pe-
riod are close to those found in HD 186943 and HD 90657,
even though its mass ratio of 0.38 is somewhat smaller
than what is observed. Systems R1 and R2 evolve into a
11 M⊙+24 M⊙ WR+O system with a 9.8 day orbital pe-
riod. I.e., period and mass ratio (0.46) agree well with the
observed systems, but the stellar masses are somewhat
smaller than observed (cf. Sect. 2). Systems N11 and N12
evolve into a WR+O system of 19 M⊙+35 M⊙ with an or-
bital period of 12..14 days. In this case, both masses and
the mass ratio (0.53) agree well with the observed ones,
but the orbital period is slightly too large. I.e., although
none of our models is a perfect match of HD 186943 or
HD 90657 -- which to find would require many more mod-
20
Petrovic et al.: WR+O progenitors
have already lost several solar masses of helium-rich mat-
ter, which causes the orbit to widen. For example, sys-
tem R6, which evolved into a 14.8 M⊙+39.0 M⊙ WR+O
system with p=8.53 days, evolves into a WC+O system
after losing ∼5 M⊙ more from the Wolf-Rayet star, which
increases its orbital period by ∼3 days. I.e., HD 211853
might have entered the WR+O stage with an orbital pe-
riod of about 4 days, which would put it together with
the shortest period WR binaries like CX Cep or CQ Cep
whose periods are 2.1 and 1.64 days respectively.
During the evolution of WR+O binary system, the pri-
mary loses mass due to WR stellar wind mass loss. WR
stellar wind mass loss of the primary decreases mass ra-
tio of the system and increases the orbital period, which
means that, for example, WC+O binary system HD63099
(MWR=9 M⊙, MO=32 M⊙ and p=14 days) could have
evolved into present state through a WN+O binary sys-
tem with q=0.5.
8. Conclusions
In an effort
to constrain the progenitor evolution
of the three WN+O binaries HD 186943, HD 90657,
and HD 211853, we calculated the evolution of non-
conservative Case A binary systems with primaries
M1,i=41..65 M⊙ and initial mass ratios between 1.7 and
2 until the WN+O stage. We performed binary evolution
calculations neglecting rotational processes in the two stel-
lar components, and assuming a constant mass accretion
efficiency of 10% for all three phases of the mass transfer,
fast Case A, slow Case A, and Case AB. Those models
could match two of the three systems reasonably well,
while HD 211853, which has the shortest orbital period,
the largest mass ratio, and a WN/WC Wolf-Rayet com-
ponent, was found to be not well explained by contact-free
evolutionary models: While models with shorter initial or-
bital periods result in short periods during the WR+O
stage, the initial WR mass is decreasing at the same time,
which leads to smaller initial WR/O mass ratios.
We then computed binary evolution models including
the physics of rotation in both stellar components as well
as the spin-up process of the mass gainer due to angular
momentum accretion. In these models, the surface of the
accreting star is continuously spun-up by accretion, while
at the same time angular momentum is transported from
the outer layers into the stellar interior by rotationally
induced mixing processes. By employing a simple model
for the mass loss of rapidly rotating luminous stars -- the
so called Ω-limit, which was actually worked out to de-
scribe the mass loss processes in Luminous Blue Variables
(Langer 1997) -- accretion is drastically reduced once the
star reaches critical rotation at its surface. The mass ac-
cretion rate is then controlled by the time scale of internal
angular momentum transport.
Some first such model for Case A and early Case B
have been computed by Wellstein (Langer et al. 2003,
2004) for a primary mass of 15 M⊙ and a mass ratio close
to one. The result was that rather high mass accretion
Fig. 21. Upper plot: Mass transfer (solid line) and ac-
cretion rate (dotted line) of the rotating initial system
56 M⊙+33 M⊙, p=6 days. Dashed line represent mass
transfer rate in the corresponding non-rotating binary.
Lower plot: accretion efficiency of the secondary taking
into account matter lost by the primary only due to mass
transfer.
els, however, might not teach us very much -- it is clear
form these results that both systems can in fact be well
explained through highly inefficient Case A mass transfer.
The situation is more difficult with HD 211853 (GP
Cep): neither the models with nor those without rota-
tion reproduce it satisfactory. HD 211853 has the shortest
period (6.7 d) and largest mass ratio (0.54) of the three
chosen Galactic WR+O binaries. While we can not ex-
clude that a Case A model of the kind presented here can
reproduce this systems, especially the small period makes
it appear more likely that this system has gone through
a contact phase: contact would reduce the orbital angular
momentum, and increase the mass loss from the system,
i.e. result in a larger WR/O mass ratio (Wellstein et al.
2001). This reasoning is strengthened by the considera-
tion that, in contrast to HD 186943 or HD 90657, the WR
star in HD 211853 is of spectral type WN6/WNC. I.e., as
this spectroscopic signature is not interpreted in terms of
a binary nature of the WR component, but rather by as-
suming that the WR star is in the transition phase from
the WN to the WC stage (Massey & Grove 1989; Langer
1991). This implies that the WR star in HD 211853 must
Petrovic et al.: WR+O progenitors
21
efficiencies (β ≃0.7) could be obtained for initial peri-
ods shorter than about 8 days. Here we find that, with
the same physical assumptions although at higher system
mass, the accretion efficiency drops to about 10% at an
initial mass ratio of 1.7. As Wellstein (2001) computed one
early Case A model for a 26 M⊙+25 M⊙ system which gave
β=0.63, it is like the high initial mass ratio in our models
which is responsible for the low accretion efficiency: Larger
initial mass ratios lead to larger mass transfer rates and,
as the time scale of internal angular momentum trans-
port in the accreting star is rather unaffected, to smaller
accretion efficiencies.
Our rotating models -- in which the accretion effi-
ciency is no free parameter any more but is computed self-
consistent and time-dependent -- reproduce the observed
WR+O binaries quite well, i.e. as good as our models
without rotation physics, where the accretion efficiency is
a free parameter. Our simplified considerations in Sect. 3
have shown that this is unlikely attributable to the free-
dom in the choice of the initial parameter of the binary
system, i.e. initial masses and period -- at least under the
assumption that contact was avoided. In case of contact,
various new parameters enter the model, similar to the
case of common envelope evolution. And indeed, also our
rotating models can not reproduce HD 211853 very well,
mostly because it currently has a too short orbital period,
which was likely even significantly shorter at the beginning
of its WR+O stage. However, this of course only confirms
the result of the simpler approaches that a contact-free
approach does not work well for this system.
In summary we can say that the system mass and an-
gular momentum loss model used here -- which is the
first detailed approach to tackle the long-standing angu-
lar momentum problem in mass transferring binaries --
has passed the test of WR+O binaries. However, it still
needs to be explored over which part of the space spanned
by the initial binary parameters this model works well,
and to what extent its results are sensitive to future im-
provements in the stellar interior physics. The inclusion of
magnetic fields generated by differential rotation (Spruit
2002) will be the next step in this direction (Petrovic et al.
2004).
References
Hamann, W.-R., Koesterke, L., & Wessolowski, U. 1995,
A&A, 299, 151
Heger, A. & Langer, N. 2000, ApJ, 544, 1016
Heger, A., Langer, N., & Woosley, S. E. 2000, ApJ, 528,
368
Heger, A., Woosley, S. E., Langer, N., & Spruit, H. C.
2004, in IAU Symposium 215, 591
Iglesias, C. A. & Rogers, F. J. 1996, ApJ, 464, 943
King, A. R., Schenker, K., Kolb, U., & Davies, M. B. 2001,
MNRAS, 321, 327
Kippenhahn, R., Kohl, K., & Weigert, A. 1967, Zeitschrift
fur Astrophysics, 66, 58
Kippenhahn, R. & Thomas, H.-C. 1970, in IAU Colloq. 4:
Stellar Rotation, 20
Kopal, Z., ed. 1978, Dynamics of Close Binary Systems
Kudritzki, R. P., Pauldrach, A., Puls, J., & Abbott, D. C.
1989, A&A, 219, 205
Lamontagne, R., Moffat, A. F. J., Drissen, L., Robert, C.,
& Matthews, J. M. 1996, AJ, 112, 2227
Langer, N. 1991, A&A, 252, 669
Langer, N. 1997, in ASP Conf. Ser. 120: Luminous Blue
Variables: Massive Stars in Transition, 83
-- . 1998, A&A, 329, 551
Langer, N., Deutschmann, A., Wellstein, S., & Hoflich, P.
2000, A&A, 362, 1046
Langer, N., Wellstein, S., & Petrovic, J. 2003, in IAU
Symposium 212, 275
Langer, N., Yoon, S.-C., Petrovic, J., & Heger, A. 2004,
in IAU Symposium 215, 535
Maeder, A. & Meynet, G. 2003, A&A, 411, 543
Massey, P. 1981, ApJ, 244, 157
Massey, P. & Grove, K. 1989, ApJ, 344, 870
Meyer, F. & Meyer-Hofmeister, E. 1983, A&A, 121, 29
Meynet, G. & Maeder, A. 2000, A&A, 361, 101
Niemela, V. S. & Moffat, A. F. J. 1982, ApJ, 259, 213
Packet, W. 1981, A&A, 102, 17
Paczy´nski, B. 1967, Acta Astronomica, 17, 355
Paczynski, B. 1991, ApJ, 370, 597
Panov, K. P. & Seggewiss, W. 1990, A&A, 227, 117
Petrovic, J. & Langer, N. 2004, in Revista Mexicana de
Astronomia y Astrofisica Conference Series, 231 -- 231
Petrovic, J., Langer, N., Yoon, S.-C., & Heger, A. 2004,
A&A, accepted
Podsiadlowski, P., Joss, P. C., & Hsu, J. J. L. 1992, ApJ,
391, 246
Rauw, G., De Becker, M., Naz´e, Y., et al. 2004, A&A, 420,
Bonanos, A. Z., Stanek, K. Z., Udalski, A., et al. 2004,
L9
ApJ, 611, L33
Braun, A. 1998, Ph.D. Thesis
Braun, H. & Langer, N. 1995, A&A, 297, 483
Brookshaw, L. & Tavani, M. 1993, ApJ, 410, 719
Chevalier, C. & Ilovaisky, S. A. 1998, A&A, 330, 201
de Loore, C. & de Greve, J. P. 1992, A&AS, 94, 453
Demers, H., Moffat, A. F. J., Marchenko, S. V., Gayley,
Ritter, H. 1988, A&A, 202, 93
Spruit, H. C. 2002, A&A, 381, 923
Underhill, A. B., Yang, S., & Hill, G. M. 1988, PASP, 100,
1256
van den Heuvel, E. P. J. & Heise, J. 1972, Nature Physical
Science, 239, 67
van der Hucht, K. A. 2001, New Astronomy Reviews, 45,
K. G., & Morel, T. 2002, ApJ, 577, 409
135
Dessart, L., Langer, N., & Petrovic, J. 2003, A&A, 404,
991
Hamann, W.-R. & Koesterke, L. 1998, A&A, 335, 1003
Vanbeveren, D. 1982, A&A, 105, 260
-- . 1991, A&A, 252, 159
22
Petrovic et al.: WR+O progenitors
Vanbeveren, D., de Greve, J. P., de Loore, C., & van
Dessel, E. L. 1979, A&A, 73, 19
Wellstein, S. 2001, Ph.D. Thesis
Wellstein, S. & Langer, N. 1999, A&A, 350, 148
Wellstein, S., Langer, N., & Braun, H. 2001, A&A, 369,
939
Woosley, S. E. Heger, A. 2004, in IAU Symposium 215,
601
Yoon, S.-C. & Langer, N. 2004a, A&A, 419, 645
-- . 2004b, A&A, 419, 623
Zahn, J.-P. 1977, A&A, 57, 383
|
astro-ph/9702155 | 1 | 9702 | 1997-02-18T13:38:25 | Neutrino pair emission due to scattering of electrons off fluxoids in superfluid neutron star cores | [
"astro-ph"
] | We study the emission of neutrinos, resulting from the scattering of electrons off magnetic flux tubes (fluxoids) in the neutron star cores with superfluid (superconducting) protons. In the absence of proton superfluidity (T> T_{cp}), this process transforms into the well known electron synchrotron emission of neutrino pairs in a locally uniform magnetic field B, with the neutrino energy loss rate Q proportional to B^2 T^5. For temperatures T not much below T_{cp}, the synchrotron regime (Q \propto T^5) persists and the emissivity Q can be amplified by several orders of magnitude due to the appearance of the fluxoids and associated enhancement of the field within them. For lower T, the synchrotron regime transforms into the bremsstrahlung regime (Q \propto T^6) similar to the ordinary neutrino-pair bremsstrahlung of electrons which scatter off atomic nuclei. We calculate Q numerically and represent our results through a suitable analytic fit. In addition, we estimate the emissivities of two other neutrino-production mechanisms which are usually neglected -- neutrino-pair bremsstrahlung processes due to electron-proton and electron-electron collisions. We show that the electron-fluxoid and electron-electron scattering can provide the main neutrino production mechanisms in the neutron star cores with highly superfluid protons and neutrons at T < 5 10^8 K. The electron-fluxoid scattering is significant if the initial, locally uniform magnetic field B > 10^{13} G. | astro-ph | astro-ph | A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
02.04.1 - 08.09.3 - 08.14.1
ASTRONOMY
AND
ASTROPHYSICS
5.12.2017
Neutrino pair emission due to scattering of electrons off
fluxoids in superfluid neutron star cores
A.D. Kaminker1, D.G. Yakovlev1, and P. Haensel2
1 A.F. Ioffe Institute of Physics and Technology, 194021 St.Petersburg, Russia
2 N. Copernicus Astronomical Center, Polish Academy of Sciences, Bartycka 18, 00-716 Warszawa, Poland
7
9
9
1
b
e
F
8
1
1
v
5
5
1
2
0
7
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
December 5, 2017
Abstract. We study the emission of neutrinos, resulting
from the scattering of electrons off magnetic flux tubes
(fluxoids) in the neutron star cores with superfluid (su-
perconducting) protons. In the absence of proton super-
fluidity (T ≥ Tcp), this process transforms into the well
known electron synchrotron emission of neutrino pairs in
a locally uniform magnetic field B, with the neutrino
energy loss rate Q proportional to B2T 5. For temper-
atures T not much below Tcp, the synchrotron regime
(Q ∝ T 5) persists and the emissivity Q can be ampli-
fied by several orders of magnitude due to the appearance
of the fluxoids and associated enhancement of the field
within them. For lower T , the synchrotron regime trans-
forms into the bremsstrahlung regime (Q ∝ T 6) similar
to the ordinary neutrino-pair bremsstrahlung of electrons
which scatter off atomic nuclei. We calculate Q numer-
ically and represent our results through a suitable ana-
lytic fit. In addition, we estimate the emissivities of two
other neutrino-production mechanisms which are usually
neglected – neutrino-pair bremsstrahlung processes due to
electron-proton and electron-electron collisions. We show
that the electron-fluxoid and electron-electron scattering
can provide the main neutrino production mechanisms in
the neutron star cores with highly superfluid protons and
is significant if the initial, locally uniform magnetic field
neutrons at T <∼ 5×108 K. The electron-fluxoid scattering
B >∼ 1013 G.
1. Introduction
Neutrino energy losses play a decisive role in cooling of
young neutron stars (NSs). During first 105 − 106 yrs af-
ter the NS birth, the cooling is dominated by the neutrino
emission from the NS cores, of density exceeding the nor-
mal nuclear matter density, ρ0 = 2.7×1014 g cm−3 (corre-
sponding to the nucleon number density n0 = 0.16 fm−3).
Send offprint requests to: P. Haensel
The emission of neutrinos results from the weak in-
teraction processes involving baryons and/or leptons in
the NS cores. The presence of the strong internal mag-
netic field allows an additional process - the neutrino
synchrotron radiation of electrons. This process was con-
sidered by many authors (e.g, Kaminker et al. 1991, 1992,
and references therein) for a spatially homogeneous mag-
netic field.
It is widely accepted (see, e.g., Takatsuka & Tama-
gaki 1993, and references therein), that neutrons and/or
protons in the NS core can be superfluid. Let Tcp and
Tcn be the critical temperatures of the proton and neu-
tron superfluidities, respectively. Theoretical values of Tcp
and Tcn are strongly model dependent and range from
108 to 1010 K. For T ≪ Tc, the superfluidity of nucleons
strongly suppresses the neutrino luminosity produced by
the Urca and bremsstrahlung processes in the NS cores
(e.g., Pethick 1992, Yakovlev & Levenfish 1995). This en-
hances the relative importance of the neutrino synchrotron
process.
It is commonly thought that superfluid protons form
the type II superconductor (see Sauls 1989, and references
therein). Let us assume that the core of a newly born, hot
and initially nonsuperfluid NS contains a magnetic field
(e.g., primordial one, or generated at the NS formation
stage). The transition to a superconducting state in the
course of the stellar cooling would then be accompanied by
a dramatic change in the spatial structure of the magnetic
field. Initially homogeneous field splits into an ensemble of
fluxoids which contain a superstrong magnetic field, em-
bedded in the field-free superconducting medium. So far,
the neutrino synchrotron radiation has been studied only
for a homogeneous magnetic field. In the present paper,
we show that after the superfluidity onset, the neutrino
synchrotron mechanism transforms into the neutrino pair
emission due to the electron-fluxoid scattering.
Physical conditions in the superconducting NS cores
are described in Sect. 2. In Sect. 3, we present general for-
malism of the electron-fluxoid (ef ) scattering. The calcu-
2
A.D. Kaminker et al: Neutrino pair emission
lations of the neutrino emissivity are performed in Sect. 4.
In Sect. 5, we estimate the emissivities of two additional
neutrino production mechanisms which are usually ne-
glected: the bremsstrahlung processes due to electron-
proton (ep) and electron-electron (ee) scattering. No cal-
culation of these emissivities has been done earlier, to our
knowledge, although the importance of the ep scattering
has been mentioned by Schaab et al. (1996). In Sect. 6
we show that the new mechanisms (especially ef and ee
scattering) can be important in the NS cores with highly
superfluid protons and neutrons since the superfluidity re-
duces strongly the familiar neutrino generation reactions
(Urca processes, nucleon-nucleon bremsstrahlung).
2. Superconducting neutron star cores
The density within a NS core is expected to range from
about 0.5ρ0 at its outer edge (Lorenz et al. 1993), to (5–
10)ρ0 at the star center. For densities ρ <∼ 2ρ0, matter con-
sists of neutrons, with a few percent admixture of protons,
electrons, and possibly muons. The fraction of muons is
usually much smaller than that of electrons, and a simple
npe model is a good approximation. At higher densities,
ρ >∼ 2ρ0, other particles may appear, such as hyperons,
condensed pions and/or kaons, or even free quarks. De-
pendence of the neutrino emissivity on the composition of
dense matter is reviewed by Pethick (1992). For simplicity,
we will neglect exotic matter constituents, restricting our-
selves to the simplest npe model. All constituents of the
npe matter are degenerate. The electrons form an ultra-
relativistic and almost ideal gas, with the electron Fermi
energy µe ≈ ¯hc(3π2ne)1/3 ∼ (100–300) MeV, where ne is
the electron number density. The neutrons and protons
form strongly interacting Fermi liquids. The Fermi energy
of neutrons is of the order of that of the electrons, while
the Fermi energy of protons is much smaller.
Both nucleon species, n and p, can be in a superfluid
state (for a review, see Takatsuka & Tamagaki 1993). The
superfluidity is widely accepted to be of the BCS–type; the
nn and/or pp pairing occurs due to nuclear attraction. Let
us focus on the proton superfluidity. In view of a relatively
low proton number density, the inter-proton distance is
rather high, and the pp interaction is attractive in the 1S0
state. Thus, the proton pairing is most likely to be in the
1S0 state. Recent reviews on the properties of supercon-
ducting protons in the NS cores are given by Sauls (1989)
and Bhattacharya & Srinivasan (1995). The proton super-
conductivity is characterized by an isotropic energy gap
in the single-(quasi)particle spectrum ∆p, which depends
on temperature and density. The critical temperature Tcp
is related to the zero-temperature energy gap ∆p(0) by
Tcp = ∆p(0)/(1.76 kB) (Lifshitz & Pitaevskii 1980), where
kB is the Boltzmann constant. An important parameter of
superconducting protons is the pp correlation length ξ; it
measures the size of a pp Cooper pair. In the BCS model,
ξ is related to ∆p and the proton Fermi velocity vFp by
ξ = ¯hvFp/(π∆p). The zero temperature value of ξ will be
denoted by ξ0. Another important parameter is the pene-
tration depth λ of the magnetic field into a proton super-
conductor. In the case of λ ≫ ξ (which corresponds to the
conditions prevailing in the NS cores), the zero tempera-
ture penetration depth, λ0, is determined by the proton
plasma frequency, ωp (e.g., Lifshitz & Pitaevskii 1980),
λ0 =
c
ωp
,
ωp =(cid:18) 4πnpe2
p (cid:19)1/2
m∗
,
(1)
where np is the proton number density, and m∗
p is the
proton effective mass, that can differ from the bare mass
mp due to the many-body effects in dense matter. Simple
estimates yield typical values of ξ0 of a few fm, and λ0 of
a few ten fm. If so, λ0 ≫ ξ0, which means that protons
constitute a type II superconductor. Notice that the values
of λ0 and ξ0 are model dependent, and one cannot exclude
the case of λ0 ∼ ξ0, in which the proton superconductivity
could be of type I, although we will not consider this case
in the present article (some comments on the case of a
type I proton superconductor are given at the end of Sect.
6).
For T → Tcp, both ξ and λ diverge as (Tcp − T )−1/2,
while for T ≪ Tcp they can be replaced by their zero tem-
perature values, ξ0 and λ0. However, we need to know the
temperature dependence of λ for all temperatures below
Tcp. In what follows, we will approximate its tempera-
ture dependence by the Gorter–Casimir formula (Tilley &
Tilley 1990),
.
(2)
λ =
λ0
p1 − (T /Tcp)4
The transition to the type II superconductivity dur-
ing the NS cooling is accompanied by the formation of
quantized flux tubes (Abrikosov fluxoids), parallel to the
initial local magnetic field ¯B. Each fluxoid carries an ele-
mentary magnetic flux φ0 = π¯hc/e. The number of flux-
oids per unit area perpendicular to the initial field is
NF = ¯B/φ0, and the mean distance between the flux-
oids is dF = [2φ0/(√3 ¯B)]1/2. Simple estimate yields dF ≈
1500/( ¯B13)1/2 fm, where ¯B13 ≡ ¯B/(1013 G). A fluxoid has
a small central core of radius ∼ ξ containing normal pro-
tons. A typical fluxoid radius is λ. Just after the supercon-
ductivity onset (T < Tcp), this radius is large and the flux-
oids fill all the available space. When temperature drops
to about 0.8Tcp, λ reduces nearly to its zero-temperature
value λ0. Thus, λ becomes much smaller than the inter-
fluxoid distance dF, for the magnetic fields ¯B < 1015 G
which we will consider in the present article. The maxi-
mum value of B is reached at the fluxoid axis, Bmax ≃
[φ0/(πλ2)] log(λ/ξ). In our case λ ≫ ξ, and the magnetic
field profile at r ≫ ξ is given by (e.g., Lifshitz & Pitaevskii
1980):
B(r) =
φ0
2πλ2 K0(cid:16) r
λ(cid:17) ,
(3)
A.D. Kaminker et al: Neutrino pair emission
3
where K0(x) is a McDonald function. In particular, for
r ≫ λ, one has B(r) ≈ φ0(8πrλ3)−1/2 exp(−r/λ).
The superconducting state is destroyed when dF < ξ,
which corresponds to magnetic fields ¯B > Bc2 = φ0/(πξ2),
whose typical values are ∼ 1018 G. We do not consider
such strong fields.
Let us mention that the fluxoids may migrate slowly
outward the stellar core due to the buoyancy forces (Mus-
limov & Tsygan 1985, Jones 1987, Srinivasan et al. 1990).
3. General formalism
Consider the neutrino-pair emission due to scattering
of strongly degenerate, relativistic electrons off fluxoids
(Sect. 2),
(4)
e + f → e + f + ν + ¯ν.
We will treat this process using the standard perturba-
tion theory with free electrons in nonperturbed states.
The process (4) is similar to the well known neutrino-pair
bremsstrahlung due to scattering of electrons by atomic
nuclei (see, e.g., Festa and Ruderman 1969, Soyeur and
Brown 1979, Haensel et al. 1996, and references therein).
It is described by two second-order diagrams, where one
(electromagnetic) vertex is associated with electron scat-
tering by the fluxoid magnetic fields, while the other (four-
tail) vertex is due to the neutrino-pair emission.
In what follows, we will mainly use the units in which
c = ¯h = kB = 1 although we will insert the ordinary
physical units in the final expressions.
The neutrino energy loss rate (emissivity) Qflux (ergs
cm −3 s −1) of process (4) can be written as
Qflux = NF
(2π)10 Z dpZ dp′Z dkνZ dk′
× δ(ε − ε′ − ω) δ(pz − p′
ν
z − kz) ωf (1 − f ′)W,
(5)
where NF is the fluxoid surface number density defined
in Sect. 2, P = (ε, p) and P ′ = (ε′, p′) are, respec-
tively, the electron 4-momenta in the initial and final
states, Kν = (ων, kν ) and K ′
ν ) are the 4-
momenta of the neutrino and of the antineutrino, and
K = Kν + K ′
ν = (ω, k) is the 4-momentum of the neutrino
pair (ω = ων + ω ′
ν and k = kν + k′
ν ). Furthermore,
ν = (ω ′
ν, k′
f =(cid:20)1 + exp(cid:18) ε − µ
T (cid:19)(cid:21)−1
(6)
is the Fermi-Dirac function for the initial electron state,
and f ′ ≡ f (ε′) is the same function for the final electron
state. The δ-functions describe energy conservation and
momentum conservation along the fluxoid axis (the axis
z). Finally, W is the differential transition rate
W =
G2
F
2
1
(2ων)(2ω ′
ν)(2ε)(2ε′) Xσ,ν
M2,
(7)
where GF = 1.436 × 10−49 erg cm3 is the Fermi weak
interaction constant, and M2 is the squared transition
matrix element. Summation is over the electron spin states
σ before and after scattering and over the neutrino flavors
(νe, νµ, ντ ). The neutrino energies are assumed to be much
lower than the intermediate boson mass (∼80 GeV). Then
the standard approach yields
Xσ
M2 = e2 Z dq δ(p − q − p′ − k)AiAj∗
νOβ )
× Tr( KνOα K ′
× Trh ¯Lβj( P ′ + me)Lαi( P + me)i ,
Oα = γα(1 + γ5),
Lαi = ΓαG(P − Q)γi + γiG(P ′ + Q)Γα,
, Γα = CV γα + CAγαγ5.
G(P ) =
P + me
P 2 − m2
e
(8)
(9)
(10)
Here, q = p−p′−k is a momentum transfer (in the (x, y)-
plane) from an electron to a fluxoid, Q = P − P ′ − K =
(Ω, q) is an appropriate 4-momentum transfer (with no
energy transfer, Ω = 0), A ≡ A(q) is a 2-dimensional
Fourier transform of the magnetic-field vector-potential,
which lies in the (x, y) plane, and me is the electron mass.
Greek indices α and β run over (0,1,2,3) and Latin ones i
and j refer to the spatial components (1,2,3); G(P ) is the
free-electron propagator, γα is a Dirac matrix, upper bar
denotes Dirac conjugate, and P ≡ Pαγα (Berestetskii et
al. 1982). Furthermore, CV and CA are the vector and the
axial vector weak interaction constants, respectively. For
the emission of the electron neutrinos (charged + neutral
currents), one has CV = 2 sin2 θW + 0.5 and CA = 0.5,
while for the emission of the muonic or the tauonic neu-
V = 2 sin2 θW − 0.5 and
trinos (neutral currents only), C ′
A = −0.5; θW is the Weinberg angle, sin2 θW ≃ 0.23.
C ′
Using the identity (Berestetskii et al. 1982)
ν δ(4)(K − Kν − K ′
ν)
ν K ′β
K α
ωνω ′
ν
ν
Z dkνZ dk′
π
6
=
(K 2gαβ + 2K αK β),
× Trh( P ′ + me)Lαi( P + me) ¯Lβji ,
where gαβ = diag(1,−1,−1,−1) is the metric tensor. The
integration in Eq. (12) is to be carried out over the domain
in which K 2 ≥ 0. Notice that Eq. (12) reproduces the well-
known expression for the neutrino-pair bremsstrahlung of
12(2π)9 Z dpZ dp′Z dk δ(pz − p′
z − kz)
we obtain
Qflux =
e2 G2
F NF
× AiAj∗ ω
Jij = Xν
εε′ Jij f (1 − f ′),
(K αK β − K 2gαβ)
(11)
(12)
(13)
4
A.D. Kaminker et al: Neutrino pair emission
electrons which scatter off atomic nuclei (e.g., Haensel
et al. 1996). For this purpose, one needs only to replace
2πNFe2δ(pz − p′
z − kz)AiAj∗ → niU (q)2δijδi0, where ni
is the number density of nuclei and U (q) is the Fourier
transform of the electron-nucleus Coulomb potential.
The 2-dimensional Fourier transform B(q) = Bz(q) of
the fluxoid magnetic field (3) in cylindrical coordinates is
B(q) =
Φ0
λ2 Z ∞
0
dr rK0(cid:16) r
λ(cid:17) J0(qr)
Φ0q2
0
q2 + q2
0
,
(14)
where J0(x) is a Bessel function and q0 = 1/λ. Then,
using cylindrical gauge, we have A(q) = −ieA B(q)/q,
eA = (B × q)/(Bq) being a unit vector. Accordingly, in
Eq. (12) AiAj∗ = (B(q)/q)2 ei
A, where i, j = 1 or 2, for
nonvanishing components.
Aej
Equations (12) and (13) determine the neutrino emis-
sivity for any degree of electron degeneracy and relativism.
We are interested in the case of ultrarelativistic, strongly
degenerate electrons (Sect. 2). In the relativistic limit, te-
dious but straightforward calculations yield:
9 (cid:18) np mp
n0 m∗
p(cid:19)1/2
≈ 2.66 × 1016 ¯B13 T 6
Tcp(cid:19)4#1/2
× "1 −(cid:18) T
L erg s−1 cm−3 .
(19)
Here, n0 = 0.16 fm−3 is the standard (nuclear) number
density, T9 is temperature in the units of 109 K, and ¯B13
is defined in Sect. 2. Numerical expression for Qflux takes
into account emission of νe, νµ, ντ , so that C2
+ ≈ 1.675.
In Eq. (19), we introduced the dimensionless quantity L,
defined by
L =
189
(2π)9T 6q0 Z dp dp′ dk δ(pz − p′
qφ0 (cid:19)2 ω
× (cid:18) B(q)
εε′ J+ f (1 − f ′),
z − kz)
(20)
with J+ given by Eq. (16). Notice that Eq. (20) is valid
for any axially symmetric distribution of the fluxoid mag-
netic field B(r) although we will use a specific distribution,
Eqs. (3) and (14). An analogous quantity L was introduced
in the case of the neutrino-pair bremsstrahlung due to the
electron-nucleus scattering. In that case, it had a meaning
of the Coulomb logarithm (Haensel et al. 1996).
4. Practical evaluation of L
Let us calculate L from Eq. (20). Since the electrons are
strongly degenerate (Sect. 2), the main contribution to L
comes from those electron transitions, in which the elec-
tron momenta p and p′ lie in the narrow thermal shell
around the Fermi surface, ε − µ <∼ T and ε′ − µ <∼ T .
In Eq. (20) we may set dp = ε2 dε dΩ, where dΩ is a solid
angle element in the direction of p. We also put ε = p
since the electrons are ultrarelativistic (Sect. 2). Then we
can set ε = pFe in all smooth functions of ε. Afterwards,
the integration over ε is standard:
Z dε f (1 − f ′) =
ω
eω/T − 1
.
(21)
The integrand in Eq. (20) depends obviously on the rela-
tive azimuthal positions of p, p′ and k. Thus we can place
p in the (xz)-plane without any loss of generality.
(17)
We shall mainly consider the case of q0 ≪ pFe typi-
cal for the NS cores (Sect. 2). It is convenient to replace
the integration over p′ in Eq. (20) by the integration over
q. The integration is simplified because the main contri-
bution comes from the values of q ≪ pFe. Thus we can
use the approximation of small-angle scattering. Since k
can be comparable to q, we should keep two first terms
in Eq. (16). Using the inequalities k ≪ p and q ≪ p we
obtain
J+ ≈ 8(ω2 − k2)(cid:20) q2
p2 + 1(cid:21) ,
(eAp)2
(22)
q2
r
(15)
(16)
Aej
ei
A Jij ≈ C2
+J+, C2
+ =Xν
(C2
V + C2
A),
J+ =
4K 2
uv
4K 2
uv
uv
q2 [ 2(eAp)(eAp′) + εε′ − pp′ ] + 8K 2
(εε′ − pp′)(cid:2)K 2 − 2(eAk)2(cid:3)
+ 4 K 4 (eAk)2
−
+ 4K 2 (qk)2
uv
− 2K 4 [ 2(eAp)(eAp′) + εε′ − pp′ ](cid:18) 1
u −
−
v2(cid:19) (2qk − K 2)
+ K 4(cid:18) 1
u2 +
v2 (cid:19) ,
− 4K 4(eAk)(cid:18) eAp
u2 −
(qk)(eAk)2
8K 2
uv
eAp′
1
v(cid:19)2
1
where we introduce the notations
u =
(P ′ + K)2 = P ′K +
1
2
v = −
Using Eqs. (12), (15), and (16), we get
(P − K)2 = P K −
1
2
1
2
K 2 .
K 2 ,
1
2
Qflux =
G2
Fe2C2
+
12 (2π)9 NFZ dpZ dp′Z dk δ(pz − p′
q (cid:19)2 ω
× (cid:18) B(q)
εε′ J+ f (1 − f ′) .
z − kz)
(18)
For practical applications, it is convenient to express
Qflux in the form
G2
Fe2φ2
0C2
+
2268 ¯h9c8
Qflux =
(kBT )6q0NF L
A.D. Kaminker et al: Neutrino pair emission
5
4.1. Low-temperature, bremsstrahlung regime, T ≪ T0
In the limiting case of t0 ≪ 1, Eq. (26) gives L = π/4.
Then the neutrino emissivity is then given by
Qflux =
πG2
Fe2φ2
0C2
+
9072 ¯h9c8
(kBT )6q0NF
9 (cid:18) np mp
n0 m∗
p(cid:19)1/2
≈ 2.09 × 1016 ¯B13 T 6
Tcp(cid:19)4#1/2
× "1 −(cid:18) T
erg s−1 cm−3.
(28)
The neutrino emission in the low–temperature regime
is very similar to that due to the electron–nucleus
bremsstrahlung (e.g., Haensel et al. 1996). In particular,
the emissivity Qflux is proportional to T 6. Therefore we
will call this regime the bremsstrahlung regime.
For the sake of completeness, we have also examined a
not too realistic case of q0 ∼ pFe at T ≪ T0. The analysis
is based on Eq. (20). In this case, the integration can be
simplified, and reduces to
L =Z 1
0
dy
y3
0
(y2 + y2
0)2
1 + y2
1 − y2 (cid:20)1 +
2y2
1 − y2 log(y)(cid:21) ,
(29)
where y = q/(2pFe). Now, L is a function of the only
dimensionless parameter (see Sect.2)
3π¯hc
pFe
m∗
¯hq0
2pFe
pc(cid:19)1/2"1 −(cid:18) T
n0(cid:19)1/6"1 −(cid:18) T
=(cid:18) e2
y0 ≡
≈ 0.0165(cid:18)mp
In the limit of y0 ≪ 1 we reproduce our basic result L =
π/4. For a finite y0, the function L given by Eq. (29) can be
fitted (with relative error < 2.4 %) by a simple expression
Tcp(cid:19)4#1/2
Tcp(cid:19)4#1/2
p(cid:19)1/2(cid:18) ne
(30)
m∗
.
where we have neglected the quantity εε′ − pp′ = [(q +
k)2 − ω2]/2 in the first term of (16), and have used the
approximate expressions u ≈ v ≈ pqr, which follow from
Eqs. (17). The subscript r denotes the vector component
along p; qr = qx sin θ, where θ is an angle between p and
the z-axis (electron pitch-angle).
Using energy-momentum conservation,
in analogy
with the results of Haensel et al. (1996) we obtain
ω = ε − ε′ ≈ qr + kr,
r − k2
K 2 = ω2 − k2
t ≈ k2
k2
0 = qr(2ω − qr),
q2
J+ = 8(ω2 − k2)
q2
r
where kt is a component of k transverse to p.
0 − k2
t ,
sin2 θ,
(23)
L =
k2
t < k2
obtain
189
The condition ω2 ≥ k2 requires qr > 0, ω > qr/2 and
0 = qr(2ω − qr). Taking into account Eq. (21), we
φ0 qr(cid:19)2
24π7 T 6 q0 Z π
dqy(cid:18) B(q)
× Z ∞
dqxZ ∞
0 − k2
t ).
eω/T − 1 Z k0
dθ sin3 θZ ∞
dkt kt (k2
(24)
ω2
dω
qr /2
−∞
0
0
0
The integrals converge rapidly due to sharp decrease of
B(q), Eq. (14). In view of this, we could extend the inte-
gration over qx and qy up to ∞, and the lower integration
limit over qy down to −∞.
Integrating over kt, we have
dqy
ω2
eω/T − 1
L =
which can be transformed as
0
0
0
qr
qr /2
189
8π7 T 6 q0 Z π
φ0 (cid:19)2 Z ∞
× (cid:18) B(q)
32π6 T 6 Z π
× Z ∞
dω (cid:16)ω −
dθ sin3 θZ ∞
dqxZ ∞
2 (cid:17)2
dω (cid:16)ω −
dθ sin3 θZ ∞
2 (cid:17)2
eω/T − 1
dqx
x + q2
189 q3
0
0
ω2
(q2
qr /2
qr
0
.
L =
0)3/2
,
(25)
These equations take into account weak inelastic-
ity of the ef scattering, in analogy with the neutrino
bremsstrahlung due to the electron-nucleus collisions
(Haensel et al. 1996).
As can be seen from Eq. (26), L depends on the dimen-
sionless parameter t0 = T /T0:
p n0
Tcp(cid:19)4#−1/2
mp np(cid:19)1/2"1 −(cid:18) T
t0 ≈ 0.00786 T9(cid:18) m∗
where T0 = Tpp1 − (T /Tcp)4, and Tp = ¯hωp/kB is the
proton plasma temperature corresponding to the proton
plasma frequency ωp (Sect. 2). Typically, Tp ∼ 1 MeV,
under the conditions in the NS cores. Let us analyze two
extreme cases: T ≪ T0 and T ≫ T0.
(27)
,
(26)
L =
π
4p1 + 0.7057 y2
0
.
(31)
4.2. High-temperature, synchrotron regime, T ≫ T0
In the case of t0 ≫ 1, from Eq. (26) we obtain an asymp-
totic form L ≈ (189/π6) ζ(5)/t0, where ζ(5) ≈ 1.037 is the
value of the Riemann zeta-function. This yields
Qflux =
ζ(5) G2
0C2
+
Fe2φ2
12π6 ¯h8c7
np mp
n0 m∗
0 (kBT )5NF ≈ 6.89 × 1017 ¯B13
q2
Tcp(cid:19)4# erg s−1 cm−3.
(32)
p "1 −(cid:18) T
× T 5
9
The temperature dependence (Q ∝ T 5) is the same as for
the synchrotron emission of neutrinos by electrons (e.g.,
6
A.D. Kaminker et al: Neutrino pair emission
Kaminker et al. 1991). Thus we will refer to the high-
temperature regime as the synchrotron regime.
Let us compare the ef -scattering emissivity (32) with
the emissivity Qsyn of the "purely synchrotron process"
for the most important case of ω∗
B ( µe is
the electron chemical potential and ω∗
B = eBc/µe is the
electron gyrofrequency). For the uniform magnetic field
Kaminker et al. (1991) obtained:
B ≪ T ≪ µ3ω∗
Qsyn =
2ζ(5)
9π5
e2G2
FC2
+(kBT )5 B2
¯h8c7
≈ 9.04 × 1014B2
13T 5
9
ergs cm−3 s−1 .
(33)
(The numerical factor 9.04 differs from a factor 8.97 in
Eq. (17) of Kaminker et al. 1991, because in the present
paper we use more accurate value of the Fermi constant
GF.)
We start with a brief discussion of the ef scattering at
t0 ≫ 1 for an arbitrary distribution of the fluxoid mag-
netic field B(q). In this case, we can set qr → 0 in Eq. (25).
Then the integrations over θ and ω are decoupled from
the integration over q, and can be done analytically. This
yields the emissivity
Qflux =
1
12π7 ζ(5) G2
FC2
+T 5 NFZ dq B2(q),
(34)
where the integration should be done over the entire
(qx, qy) plane. The emissivity is seen to be indepen-
dent of the electron Fermi momentum. In particular, in
the limit of a (quasi) uniform magnetic field B(r) =
B0, B(q) = 4π2 B0 δ(2)(q), we can replace B(q)2 →
4π2 B2
0 δ(2)(q)/NF, and obtain
+(kBT )5 B2
0
¯h8c7
ζ(5)
3π5
e2G2
FC2
Qflux =
.
(35)
Comparing (35) and (33), we have Qsyn = (2/3) Qflux.
The same factor 2/3 is obtained, if we treat B in Qsyn as
a local magnetic field of a fluxoid, average Qsyn ∝ B2
over a lattice of fluxoids with the magnetic field (3), and
compare the result with Qflux, Eq. (32). Let us empha-
size, that this emissivity ¯Qsyn, averaged over a nonuniform
magnetic field, can differ from the synchrotron emissivity
Qsyn in the initial uniform field ¯B by a large factor of
B2/ ¯B2 = φ0q2
0/(4π ¯B) ∼ (dF/λ)2, where dF is an inter-
fluxoid distance (Sect. 2). This increase of ¯Qsyn is caused
by the magnetic field enhancement within the fluxoids due
to magnetic flux conservation. The enhancement is much
stronger than the reduction of neutrino-emission space oc-
cupied by the fluxoids.
A detailed analysis shows, that the difference by a
factor of 2/3 comes from momentum space available for
neutrino-pair momentum k. The space is different in the
case of synchrotron radiation in a uniform magnetic field
and in the case of ef scattering by magnetic inhomo-
geneities (in our case - fluxoids). Let us fix the electron
pitch angle θ and average over the electron Larmor rota-
tion. In the both cases k is concentrated to the cone of
a given θ (neutrinos are emitted predominantly within a
narrow angle interval δθ in the direction of the electron
pitch angle θ), and typically, k ∼ T . In the case of the
ef scattering, our estimates yield δθflux ∼ pq0/T . The
synchrotron radiation consists of many discrete cyclotron
B ≫ 1, for the
harmonics s with typical values s ∼ T /ω∗
quasiclassical regime of interest. For a fixed θ, the radia-
tion in each harmonics is peaked at the pitch-angle cone,
and maximum neutrino momenta k are restricted by a
parabolic surface (Kaminker et al. 1991). The maximum
B/(sin θ)2, and,
value of k in the pitch-angle direction is sω∗
typically, δθsyn ∼ s−1/3. Replacing a sum over harmonics
by an integral we obtain a smoothed synchrotron k space.
It fills only 2/3 of the available momentum space, due to
the parabolic momentum space restriction for each har-
monics (e.g., the surface area below the parabolic curve
y = 1 − x2 at 0 ≤ x ≤ 1 is exactly 2/3 of the surface
area below a straight segment y = 1). A transition from
the synchrotron formula (33) to the fluxoid formula (32)
is expected to occur at δθflux ∼ δθsyn. This is equivalent
B)s−1/3, which takes place when
to the condition λ ∼ (c/ω∗
the scale of the magnetic field variation λ becomes compa-
rable to a distance along which a neutrino pair is emitted
in a uniform magnetic field, i.e., to the s1/3-th part of the
Larmor radius c/ω∗
B.
4.3. Overall fit and overview
Combining the results of Sects. 4.1 and 4.2, we can pro-
pose the following fitting expression for the quantity L in
Eq. (19) at y0 ≪ 1:
L = L0 U V,
(36)
L0 =
π
4
0.260 t0 + 0.0133
t2
0 + 0.25 t0 + 0.0133
,
U =
2γ + 1
3γ + 1
,
B
γ = (cid:18) ¯hω∗
× (cid:18) n0
µe (cid:19)2 t0
ne(cid:19)13/6 (cid:18) m∗
0 ≈ 8.38 × 10−12 T9 ¯B2
y2
mp(cid:19)3/2 "1 −(cid:18) T
p
Tcp(cid:19)4#−3/2
13
V = 1 +
4π ¯Bλ2
φ0
≈ 1 + 0.00210 ¯B13
m∗
p n0
mp np "1 −(cid:18) T
Tcp(cid:19)4#−1
.
.
(37)
(38)
Here, L0 is the analytic expression which fits the results
of our numerical calculations of L from Eq. (26) with er-
ror < 1 % for any t0. The factor U ensures, somewhat
arbitrarily, the difference by a factor of 2/3 between the
cases of weakly and strongly nonuniform magnetic fields
(Sect. 4.2). The factor V provides a smooth transition
A.D. Kaminker et al: Neutrino pair emission
7
from the ef scattering to the pure synchrotron emission
in the case when a fluxoid radius λ becomes very large
(T → Tcp), the neighboring fluxoids overlap and the over-
all magnetic field becomes nearly uniform.
Using our results, we can follow the evolution of Qflux
in the course of the superconductivity onset. If T ≥ Tcp,
the emissivity Qflux is given by Eq. (33) since it is es-
sentially the same as the synchrotron emissivity Q(0)
syn =
Qsyn( ¯B) in a locally uniform 'primordial' magnetic field
B = ¯B. After T falls only slightly below Tcp, the factor U
transforms from U = 2/3 ('pure synchrotron') to U = 1
(ef scattering in the synchrotron regime). The fluxoid
structure is still not very pronounced, i.e., V ≈ 4π ¯Bλ2/φ0,
Qflux ∼ Q(0)
syn. When T decreases to about 0.8 Tcp, the flux-
oid radius λ nearly achieves its zero-temperature value λ0,
and V ≃ 1. Accordingly, Qflux/Q(0)
syn grows by a factor of
(dF/λ0)2 due to the magnetic field confinement within the
fluxoids (Sect. 4.2). Simultaneously, t0 decreases from very
large values at T → Tcp to t0 ≃ Tcp/Tp.
Let us first consider the case in which the proton
plasma temperature is Tp ≪ Tcp. Then the emissiv-
ity Qflux is enhanced with respect to Q(0)
syn by a factor
of (dF/λ0)2 at T ≈ 0.8Tcp, and this enhancement re-
mains nearly constant over a wide temperature range
down to Tp. Within this range, the ef scattering oper-
ates in the synchrotron regime. At lower temperatures,
T <∼ Tp, the synchrotron regime transforms into the
bremsstrahlung regime (Sect. 4.1) and we have Qflux ∼
Q(0)
syn(dF/λ0)2 (T /Tp). Thus, for T <∼ Tp the emissivity
Qflux decreases with respect to Q(0)
lower than Q(0)
syn.
syn, and may become
In the opposite case of Tp >∼ Tcp, the ef scatter-
ing operates in the synchrotron regime only in a narrow
temperature range below Tcp, and transforms into the
bremsstrahlung regime at lower T . In the latter regime,
we have Qflux ∼ Q(0)
syn(dF/λ)2(T /Tp).
5. Neutrino bremsstrahlung due to ep and ee scat-
tering
As will be shown in Sect. 6, the ef scattering mechanism
can be important, if protons and neutrons in a NS core
are highly superfluid (have high critical temperatures Tcp,
Tcn), so that the rates of the "standard" Urca and nucleon
bremsstrahlung processes are strongly suppressed. Under
such conditions, one should carefully take into account
all the neutrino processes, even if they are negligible for
relatively low critical temperatures.
In this section, we consider two additional neutrino
production mechanisms, the neutrino-pair bremsstrahlung
due to the ep and ee scattering:
e + p → e + p + ν + ¯ν,
e + e → e + e + ν + ¯ν.
(39)
As far as we know, these mechanisms were neglected in
the neutron-star cooling simulations.
Let us restrict ourselves to simple estimates of the
emissivities Qep and Qee; detailed analysis will be pub-
lished elsewhere (Kaminker et al.
in preparation). We
start from the well-known emissivity QeZ of the electron-
nucleus bremsstrahlung (e.g., Haensel et al. 1996). In the
particular case of the ep collisions of relativistic, strongly
degenerate electrons in a nonsuperfluid matter, we obtain
Q(0)
ep =
+e4G2
8πC2
F
567 ¯h9c8
(kBT )6np Lep ,
(40)
where Lep is the Coulomb logarithm. The above formula
has been derived assuming, that an energy transfer to a
proton in a neutrino-emission act is much smaller than
T . As applied to our case of strongly degenerate protons,
the latter means that the Pauli principle imposes no re-
strictions on the states of protons. This is true as long as
T ≫ qs, where qs is an inverse length of plasma screening
of the Coulomb ep interaction (see below). However, in
reality we have just an opposite regime, T ≪ qs, and Eq.
(40) does not hold.
To estimate Qep at T ≪ qs, we assume that Qep ∝ νep,
where νep is some effective ep collision frequency. If our as-
sumption is correct, we have Qep ≈ Q(0)
ep , where
ν(0)
ep and νep refer to the high- and low- temperature
cases, respectively. For high temperatures, we will use the
effective frequency of nearly elastic collisions ν(0)
(e.g.,
ep
Yakovlev and Urpin 1980) which determines the electric
and thermal conductivities of degenerate electrons. For
T ≪ qs, we employ the collision frequency νep that de-
termines the electron thermal conductivity (e.g., Gnedin
and Yakovlev 1995). Performing the above rescaling, we
obtain
ep νep/ν(0)
Qep =
np (kBT )8Rep
+m∗2
π5G2
Fe4C2
p
945¯h9c8y3
s p4
Fe
3.42 × 1014
mp(cid:19)2(cid:18) n0
(cid:18) m∗
np(cid:19)1/3
p
≈
y3
s
9 ergs cm−3 s−1,
× Rep T 8
(41)
where ys = ¯hqs/(2pFe) is the plasma screening parameter.
We additionally introduce the factor Rep that describes
suppression of the emissivity by the proton superfluid-
ity. According to Gnedin & Yakovlev (1995) the plasma
screening in npe matter of the NS cores is defined by
y2
s =
e2
π¯hc(cid:18)1 +
m∗
m∗
p pFp
e pFe
Zp(cid:19) ,
(42)
e = µe/c2. The first term in brackets comes from
where m∗
the electron screening while the second is due to the pro-
ton screening. The latter contains the factor Zp which de-
scribes reduction of the proton screening by the proton
8
A.D. Kaminker et al: Neutrino pair emission
superfluidity. According to Gnedin & Yakovlev (1995) this
factor is fitted as
with those produced by the neutron and proton branches
of the modified Urca reactions (Friman & Maxwell 1979,
Yakovlev & Levenfish 1995),
Zp = (cid:16)0.9443 +p(0.0557)2 + (0.1886v)2(cid:17)1/2
× exp(cid:16)1.753 −p(1.753)2 + v2(cid:17) ,
where v is the proton superfluid gap parameter
v =
∆p(T )
kBT
=s1 −
T
Tcp
T
Tcp
(44)
+ 1.764
T ! .
× 1.456 − 0.157rTcp
If protons are normal (T ≥ Tcp), one has Rep = Zp = 1
in Eqs. (41) and (42). The factor Rep, which damps Qep
at T < Tcp, should be the same as the factor Rnp that
describes reduction of the neutrino emission in np colli-
sions by the proton superfluidity (Yakovlev and Levenfish
1995):
Rep =
1
2.732ha exp(cid:16)1.306 −p(1.306)2 + v2(cid:17)
+ 1.732 b7 exp(cid:16)3.303 −p(3.303)2 + 4v2(cid:17)i ,
with a = 0.9982 +p(0.0018)2 + (0.3815v)2,
b = 0.3949 +p(0.6051)2 + (0.2666v)2.
Now consider the neutrino emissivity Qee due to ee
scattering. Under the same assumption, we obtain Qee =
Qep νee/νep, where νee is the effective ee collision fre-
quency that determines the electron thermal conductivity
(e.g., Gnedin & Yakovlev 1995). Then
(45)
ne (kBT )8
Qee =
≈
π5G2
Fe4C2
+
378¯h9c10y3
s p2
Fe
1.07 × 1014
y3
s
n0(cid:19)1/3
(cid:18) ne
T 8
9 ergs cm−3 s−1.
(46)
Although we do not calculate Qpe and Qee exactly, we
believe that Eqs. (41) and (46) give reliable estimates. The
emissivity Qep is reduced exponentially by the strong pro-
ton superfluidity, while Qee is affected by the superfluidity
in a much weaker manner, only through the plasma screen-
ing parameter (42). If protons are normal they provide the
major contribution into the plasma screening. If they are
strongly superfluid, a weaker electron screening becomes
important, which enhances Qee, but not to a great extent
(see below).
6. Discussion
Figures 1 and 2 illustrate the efficiency of various neutrino
production mechanisms in npe matter of the NS cores. We
compare the neutrino emissivity due to the ef scatter-
ing or due to the electron synchrotron radiation (Sect. 4)
(43)
n + n → n + p + e + ¯νe, n + p + e → n + n + νe ;
n + p → p + p + e + ¯νe, p + p + e → n + p + νe ;
by the nucleon-nucleon bremsstrahlung processes (Friman
& Maxwell 1979, Yakovlev & Levenfish 1995)
(47)
(48)
n + n → n + n + ν + ¯ν,
n + p → n + p + ν + ¯ν,
p + p → p + p + ν + ¯ν;
and also by the ep and ee bremsstrahlung processes (39)
considered in Sect. 5. For better presentation, the total
emissivity of the both branches of the modified Urca re-
actions (47) is represented by a single curve 'Murca',
and the total emissivity from the three nucleon-nucleon
bremsstrahlung reactions (48) are represented by a sin-
gle curve 'NN'. We adopted a moderately stiff equation of
state of Prakash et al. (1988) (the same version as used
by Page & Applegate 1992) and choose matter density
ρ = 5.6×1014 g cm−3 (np = ne ≈ 0.0207 fm−3, nn ≈ 0.323
fm−3), well below the threshold density 1.30×1015 g cm−3
at which the direct Urca process becomes allowed (Lat-
timer et al. 1991). Therefore our curves correspond to the
standard neutrino emission (no exotic cooling agents, see,
e.g., Pethick 1992). We show temperature dependence of
various neutrino emissivities in the range from 108 K to
5 × 109 K, which is the most important for the cooling
theories of NS. The nucleon effective masses are set equal
to 0.7 of the masses of bare particles. The neutrino emis-
sivities vary slowly with density (below the direct Urca
threshold), i.e., we present a typical situation within the
NS cores. In both figures, we assume the presence of the
'primordial' magnetic field ¯B = 1012, 1013 or 1014 G.
Figure 1 corresponds to a nonsuperfluid matter. If the
magnetic field is absent, the Murca processes are seen to
be the most important ones, in the displayed temperature
range. The N N bremsstrahlung is much weaker, and the
ep and ee bremsstrahlung processes are even weaker. The
emissivities of all these mechanisms vary as T 8, whereas
the emissivity of the synchrotron radiation varies as T 5.
Thus the synchrotron radiation is negligible for T >∼ 109
K but becomes more important with decreasing T . At
T ∼ 108 K and B = 1014 G the synchrotron dominates
over all other mechanisms except the Murca processes,
and at slightly lower T it becomes the most efficient of
all mechanisms. However, such temperatures are too low
to be important for the cooling theories: the neutrino lu-
minosity of NS becomes then smaller than the photon lu-
minosity from the stellar surface, and the neutrino emis-
sion looses its significance as a source of the NS cooling.
Figure 2 shows how the neutrino emission displayed in
Figure 1 is modified by the nucleon superfluidity. In our
A.D. Kaminker et al: Neutrino pair emission
9
Fig. 1. Neutrino energy loss rates from various processes, ver-
sus temperature in a non–superfluid npe matter of density
5.6 × 1014 g cm−3. Murca - total contribution of two branches
of the modified Urca process, Eq. (47); NN - total contribu-
tion of three nucleon bremsstrahlung processes, Eq. (48); ep
and ee - bremsstrahlung processes, Eq. (39), due to the ep
and ee scattering, respectively; solid lines - synchrotron radi-
ation of electrons for three values of the magnetic field.
example, the protons and neutrons become superfluid at
Tcp = Tcn = 2 × 109 K. These are typical, moderate
critical temperatures for the nucleon superfluidity (e.g.,
Takatsuka & Tamagaki 1993). The factors which describe
superfluid suppression of the modified Urca processes (47)
and N N collisions (48) are taken from Yakovlev & Lev-
enfish (1995). At T <∼ 109 K the nucleon superfluidity is
seen to strongly reduce all the neutrino production pro-
cesses which involve nucleons. If the magnetic field were
absent, the ee scattering would be the main neutrino gen-
eration mechanism at T <∼ 5 × 108 K. However, the ef
scattering in the presence of ¯B >∼ 1013 G also becomes
important, and it can even dominate. After the superfluid-
ity onset (T < Tcp), the magnetic field splits into fluxoids
and the synchrotron radiation transforms into the radi-
Fig. 2. Same as in Fig. 1 but in the presence of proton and
neutron superfluidities with Tcp = Tcn = 2×109 K. At T < Tcp
the initially uniform magnetic field splits into fluxoids, and the
synchrotron emission transforms into the emission due to the
ef scattering. The nucleon superfluidity suppresses all mecha-
nisms but the ee and ef scattering.
ation due to the ef scattering. This enhances Qflux ow-
ing to the field amplification within the fluxoids (Sect. 4).
The enhancement is more pronounced at lower ¯B ∼ 1012
G at which the field amplification is stronger. This is a
rare case in which the neutrino emissivity increases with
decreasing T . For the conditions displayed in Fig. 2, the
proton plasma temperature Tp = 5.47 × 1010 K is much
higher than the superfluid critical temperature Tcp. There-
fore, the synchrotron regime (Sect. 4.2) in the ef scatter-
ing operates only at T <∼ Tcp, during the phase of the
fluxoid formation. Very soon after T falls below Tcp, the
synchrotron regime transforms into the bremsstrahlung
regime (Sect. 4.1) which operates further with decreasing
T .
We have analysed a number of cases, varying the pro-
ton and neutron critical temperatures Tcp and Tcn. Our
principal conclusion is that the standard neutrino pro-
10
A.D. Kaminker et al: Neutrino pair emission
duction mechanisms (47) and (48) dominate in the tem-
perature domain of practical interest if either Tcp and/or
Tcn are not too high (not higher than about 109 K). If,
however, both critical temperatures are high, the situa-
tion is similar to that shown in Fig. 2. The superfluidity
suppresses all the traditional neutrino generation mecha-
nisms, and the main neutrino production at T <∼ 5 × 108
K comes either from the ef or the ee scattering.
In principle, neutrinos can be generated also in the pp
collisions of normal protons in the non-superfluid cores
of the fluxoids (Sect. 2) as well as in the en collisions.
However, the non-superfluid cores are very thin, the emis-
sion volume is minor, and the first process is inefficient.
The neutrino bremsstrahlung due to en scattering is also
inefficient since it occurs through electromagnetic interac-
tion involving neutron magnetic moment (e.g., Baym et
al. 1969) and since it is suppressed by the neutron super-
fluidity.
While it is commonly accepted that protons form a
type II superconductor, a possibility that they actually
form a type I superconductor cannot be excluded. This
uncertainty results from the lack of a precise knowledge
of the nucleon-nucleon interaction in dense nuclear mat-
ter, and from the approximations and deficiencies of the
many-body theory of dense nucleon matter. In the case
of a type I proton superconductor, which corresponds to
ξ0 > √2λ0 (de Gennes 1966), cooling below Tcp is ex-
pected to be accompanied by a transition of the magne-
tized interior to an "intermediate state" (de Gennes 1966,
Baym et al. 1969). The "intermediate state" would con-
sist of alternating regions of normal matter containing
magnetic flux, and superconducting regions exhibiting a
complete expulsion of magnetic flux. The specific spatial
structure of the "intermediate state" would result from
the condition of the minimum of the thermodynamic po-
tential at a fixed macroscopic magnetic flux (de Gennes
1966). Qualitatively, we expect that the presence of an
"intermediate state" in the superconducting proton core
will imply an enhancement of its electron synchrotron ¯νν
emissivity, compared to the case of a normal proton core,
because of B2 > ( ¯B)2. However, for ¯B <∼ 1014 G this en-
hancement will be much smaller than that characteristic
of the type II superconductor, in which magnetic field is
confined to fluxoids.
7. Conclusion
We have considered (Sects. 3 and 4) the neutrino-pair ra-
diation (4) due to scattering of relativistic, degenerate
electrons off threads of quantum magnetic flux - fluxoids
- in the npe matter within the superfluid NS cores. We
have shown that this mechanism is a natural generaliza-
tion of the synchrotron emission of neutrinos by electrons
in a non-superconducting matter (with a locally uniform
magnetic field) to the case in which the protons become su-
perfluid and the magnetic field splits into the fluxoids. Ac-
cording to our results, this mechanism can operate either
in the synchrotron (Sect. 4.2) or in the bremsstrahlung
(Sect. 4.1) regime. We have obtained a simple fitting ex-
pression (36) which reproduces the emissivity (19) for any
parameters of practical interest.
Furthermore, we have estimated (Sect. 5) the neu-
trino emissivities in two additional neutrino production
mechanisms (39), the ep and ee bremsstrahlung processes.
In a non–superfluid matter, these mechanisms are much
weaker than the standard neutrino emission mechanisms,
such as the modified Urca processes (47) and the nucleon–
nucleon bremsstrahlung (48). If, however, both superfluid
critical temperatures, Tcp and Tcn, are higher than 109
K, then the superfluidity strongly suppresses all the tra-
ditional (standard) neutrino energy losses. In such a case,
the main neutrino generation in a NS core at T <∼ 5 × 108
K occurs either via the electron-fluxoid or via the ee scat-
tering (Sect. 6). Therefore our results can be of particular
importance for simulating the cooling of NSs with highly
superfluid cores.
Acknowledgements. The authors are grateful to Kseniya Lev-
enfish who participated at the initial stage of this work. One of
the authors (ADK) acknowledges excellent working conditions
and hospitality of N. Copernicus Astronomical Center in War-
saw. This work was supported in part by the RBRF (grant
No. 96-02-16870a), INTAS (grant No. 94-3834), DFG-RBRF
(grant No. 96-02-00177G), and the KBN grant 2P 304 014 07.
References
Baym, G., Pethick, C., Pines, D. 1969, Nature, 224, 673
Berestetskii, V.B., Lifshitz, E.M., Pitaevskii, L.P. 1982, Quan-
tum Electrodynamics, Pergamon, Oxford
Bhattacharya, D., Srinivasan, G. 1995, in: W.H. Lewin, J. van
Paradijs, E.P.J. van den Heuvel (eds) X-Ray Binaries. Cam-
bridge UP, Cambridge, p. 495
Festa, G.G., Ruderman, M.A. 1969, Phys. Rev., 180, 1227
Friman, B.L., Maxwell, O.V. 1979, ApJ, 232, 541
de Gennes, P.G. 1966, Superconductivity of Metals and Alloys,
Benjamin, New York
Gnedin, O.Y., Yakovlev, D.G. 1995, Nucl. Phys., A582, 697
Haensel, P., Kaminker, A.D., Yakovlev, D.G. 1996, A&A, 314,
328
Jones, P.B. 1987, MNRS, 228, 513
Kaminker, A.D., Levenfish, K.P., Yakovlev, D.G. 1991, Astron.
Lett., 17, 1090
Kaminker, A.D., Levenfish, K.P., Yakovlev, D.G., Amster-
damski, P., Haensel, P. 1992, Phys. Rev., D46, 3256
Lattimer, J.M., Pethick, C.J., Prakash, M., Haensel, P. 1991,
Phys. Rev. Lett., 66, 2701
Lifshitz, E.M., Pitaevskii, L.P. 1980, Statistical Physics, part
2, Pergamon, Oxford
Lorenz, C.P., Ravenhall, D.G., Pethick, C.J. 1993, Phys. Rev.
Lett., 70, 379
Muslimov, A.G., Tsygan, A.I. 1985, Astrophys. Space Sci., 115,
43
Page, D., Applegate, J.H. 1992, ApJ (Letters), 394, L17
Pethick, C.J. 1992, Rev. Mod. Phys., 64, 1133
A.D. Kaminker et al: Neutrino pair emission
11
Prakash, M., Ainsworth, T.L., Lattimer, J.M. 1988, Phys. Rev.
Lett., 61, 2518
Sauls, J.A. 1989, in: H. Ogelman and E.P.J. van den Heuvel
(eds) Timing Neutron Stars. Dordrecht: Kluwer Academic
Publishers, p. 457
Schaab, C., Weber, F., Weigel, M.K., Glendenning, N.K. 1996,
Nucl. Phys., A605, 531
Soyeur, M., Brown, G.E. 1979, Nucl. Phys., A324, 464
Srinivasan, G., Bhattacharya, D., Muslimov, A.G., Tsygan,
A.I. 1990, Current Science, 59, 31
Takatsuka, T., Tamagaki, R. 1993, Progr. Theor. Phys. Suppl.,
112, 27
Tilley, D.R., Tilley, J. 1990, Superfluidity and Superconductiv-
ity, Graduate student series in physics. Bristol: IOP Pub-
lishing
Yakovlev, D.G., Levenfish, K.P. 1995, A&A, 297, 717
Yakovlev D.G., Urpin V.A. 1980, SvA 24, 303
This article was processed by the author using Springer-Verlag
LaTEX A&A style file L-AA version 3.
|
astro-ph/9710317 | 1 | 9710 | 1997-10-28T12:50:14 | The Inner Galaxy resolved at IJK using DENIS data | [
"astro-ph"
] | We present the analysis of three colour optical/near-infrared images, in IJK, taken for the DENIS project. The region considered covers 17.4 square deg and lies within |l|<5 deg, |b|<1.5 deg. The adopted methods for deriving photometry and astrometry in these crowded images, together with an analysis of the deficiencies nevertheless remaining, are presented. The numbers of objects extracted in I,J and K are 748000, 851000 and 659000 respectively, to magnitude limits of 17,15 and 13. 80% completeness levels typically fall at magnitudes 16, 13 and 10 respectively, fainter by about 2 magnitudes than the usual DENIS limits due to the crowded nature of these fields. A simple model to describe the disk contribution to the number counts is constructed, and parameters for the dust layer derived. We find that a formal fit of parameters for the dust plane, from these data in limited directions, gives a scalelength and scaleheight of 3.4+-1.0 kpc and 40+-5 pc respectively, and a solar position 14.0+-2.5 pc below the plane. This latter value is likely to be affected by localised dust asymmetries. We convolve a detailed model of the systematic and random errors in the photometry with a simple model of the Galactic disk and dust distribution, to simulate expected colour-magnitude diagrams. These are in good agreement with the observed diagrams, allowing us to isolate those stars from the inner disk and bulge. After correcting for local dust-induced asymmetries, we find evidence for longitude-dependent asymmetries in the distant J and K sources, consistent with the general predictions of some Galactic bar models. We consider complementary L-band observations in a second paper. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 000–000 (0000)
Printed 13 November 2017
(MN LATEX style file v1.4)
The Inner Galaxy resolved at IJK using DENIS data
M. Unavane,1 Gerard Gilmore,1,4 N. Epchtein,3 G. Simon,2 D. Tiph`ene,3 B. de Batz,2
1 Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK
2 Observatoire de Paris-Meudon, DASGAL, CNRS/URA335, 5 place Jules Janssen, F-92195 Meudon Cedex, France
3 Observatoire de Paris-Meudon, DESPA, CNRS/URA264, 5 place Jules Janssen, F-92195 Meudon Cedex, France
4 Institut d'Astrophysique de Paris, 98bis Boulevard Arago, F-75014 Paris, France
13 November 2017
ABSTRACT
We present the analysis of three colour optical/near-infrared images, in IJK, taken
for the DENIS project. The region considered covers 17.4 deg2 and lies within l < 5◦,
b < 1.5◦. The adopted methods for deriving photometry and astrometry in these
crowded images, together with an analysis of the deficiencies nevertheless remaining,
are presented. The numbers of objects extracted in I,J and K are 748 000, 851 000
and 659 000 respectively, to magnitude limits of 17,15 and 13. 80% completeness levels
typically fall at magnitudes 16, 13 and 10 respectively, fainter by about 2 magnitudes
than the usual DENIS limits due to the crowded nature of these fields. A simple model
to describe the disk contribution to the number counts is constructed, and parameters
for the dust layer derived. We find that a formal fit of parameters for the dust plane,
from these data in limited directions, gives a scalelength and scaleheight of 3.4±1.0 kpc
and 40±5 pc respectively, and a solar position 14.0±2.5 pc below the plane. This latter
value is likely to be affected by localised dust asymmetries. We convolve a detailed
model of the systematic and random errors in the photometry with a simple model
of the Galactic disk and dust distribution, to simulate expected colour-magnitude
diagrams. These are in good agreement with the observed diagrams, allowing us to
isolate those stars from the inner disk and bulge. After correcting for local dust-induced
asymmetries, we find evidence for longitude-dependent asymmetries in the distant J
and K sources, consistent with the general predictions of some Galactic bar models.
We consider complementary L-band observations in a second paper.
Key words: Galaxy: stellar content – ISM: dust, extinction – Galaxy: structure –
Stars: statistics – Stars: infrared – Galaxy : model – Galaxy : bar – extraterrestrial
intelligence
7
9
9
1
t
c
O
8
2
1
v
7
1
3
0
1
7
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
INTRODUCTION
The central kiloparsec of the Galaxy is dominated by an ex-
tremely dense stellar cluster of unknown origin and history,
and poorly known properties. It is unknown if this cluster is
a remnant of the core about which the Galaxy grew, is the
product of a later merger, is a product of a long-lived bar
in the disk feeding gas into continuing star formation in the
central galaxy, or has some other history.
Its relationship, if any, to the larger Galactic bulge, halo,
and disk and to the smaller Galactic non-thermal nucleus is
entirely unknown. This cluster, whose density approaches
106 solar masses per cubic pc, or 107 times that of the So-
lar neighbourhood, is the most extreme dynamical system
available for detailed study.
The central degree or so is also an extended X-ray
source, with temperature some 108 K, and gas pressures
1000 times those of the normal ISM. Moreover, the clus-
ter changes its luminosity density profile by 2 in the power
law index in some unobserved region between the central
few arcsec and the optically observable region some degrees
away. How and where? And are more complex spatial distri-
butions possible? For example, in M31, the nearest similar
spiral, the central region shows two luminosity maxima, nei-
ther of which corresponds to the centre of the larger scale
gravitational potential, or is understood. (van der Marel et
al. 1997)
In practice, because of the high extinction, it is neces-
sary to work in the infrared.
Many high spatial resolution near IR wavelength stud-
ies of the central arcminute or so are available (cf Gen-
zel, Hollenbach & Townes 1994) for a detailed review).
However, remarkably little data is available concerning the
larger scale structure. Balloon and satellite surveys (e.g.
c(cid:13) 0000 RAS
2 M.Unavane et al.
COBE/DIRBE) are all of very low spatial resolution. The
only large higher resolution survey, covering about 2◦ × 0.5◦
around the galactic centre,
is by Glass, Catchpole, and
Whitelock (1987). It was performed in the J, H and K bands
up to a limiting magnitude of K=12. While almost the en-
tire J map was dominated by heavy interstellar extinction,
those at H and K show progressively more detail of the inner
region. They show clear changes in spatial structure for dif-
ferent populations, suggesting that analysis of low resolution
data will necessarily be problematic
For the galactic centre this means that M and late K gi-
ants can be reached in K but not at the shorter wavelengths
as the extinction will be too strong (up to 5 magnitudes
in J [Catchpole, Whitelock & Glass 1990] leading to an ex-
pected apparent J magnitude of ≈ 16.5 mag; [Wainscoat et
al. 1992]). We also expect that essentially all I and most J
objects seen in the plane will be disk objects.
Two large scale, high resolution surveys in near infra-
red bands have begun recently. 2MASS (2-micron all sky
survey) (Skrutskie et al. 1997) aims to survey the whole
celestial sphere in J,H and K, with 2 arcsecond pixels, from
two specially built identical telescopes in each hemisphere.
The second project is a European joint venture, called
DENIS (Deep Near Infra Red Southern Sky Survey) which
aims to map the whole of the southern sky in I,J and K with
3 arcsecond pixels. (Epchtein et al 1997). Technical details
are given in Copet et al. (1997).
It is the data from this project which are relevant to the
central regions of the galaxy that we consider in detail here,
considering specifically the ability of DENIS-like survey data
to study the inner Galaxy.
2 DENIS – DATA REDUCTION
2.1 Using the images
DENIS (DEep Near
InfraRed Southern Sky Survey)
(Epchtein 1997) will be a complete deep near infrared survey
of the southern sky, with the objective to provide full cover-
age in 2 near infrared bands (J at 1.25µm and K at 2.2µm)
and one optical band (I at 0.8µm), using a ground-based
telescope and digital array detectors. Spare time at the end
of DENIS observation nights during summer/autumn 1996
were used to take images in rasters (see Figure 2). Note
that the shaded circle in this figure indicates the size of the
DIRBE beam for comparison.
The standard pipeline processing of images ensures that
for each image, sky subtraction is performed, the back-
ground is made flat, so that the images are ready for ex-
traction of sources. (Borsenberger, 1997). The early pipeline
extraction procedures proved unsatisfactory for these very
crowded fields.
2.2 Source Extraction
Source extraction and aperture photometry were performed
with the SExtractor software (Bertin & Arnouts 1996).
Sources were extracted to 2σ above the background noise
level, and and aperture of 3.5 arcseconds radius was used.
The calibration and flat fielding carried out in the prepro-
cessing of the images is assumed to be sound. It is important
to note that if a background map were to be constructed for
each image as part of the source extraction procedure, it
would remove one of the effects being studied (namely, the
small scale variation of the extinction).
The most crowded fields have ∼ 8000 sources in K, cor-
responding to a mean separation of ∼ 10 arcsec, some 3 times
larger than the radius of the aperture. The zero point was
taken from standard star frames taken immediately before
and after the observation of each raster.
2.3 Source matching between IJK
Each frame suffers its own distortions – software was devel-
oped to take triplets of corresponding I,J and K catalogues
for a single image field, and map all onto a single consistent
coordinate system. The chosen coordinate system, to make
mapping straightforward, was chosen to be the coordinate
system of the J frame. The choice is astrophysically moti-
vated. The difference in the nature of the brightest sources
in the presence of high extinction varies so much between I
and K that matching these two would be unreliable if au-
tomated. The technique adopted uses the iterative fitting,
with rejection of outliers, of a 2nd order, two-dimensional
polynomial to several hundred of the brightest sources for
the transformation between images. Higher-order transfor-
mations are found to be no better.
2.4 Absolute astrometry
The usefulness of these data cannot be fully realised unless
cross matching with other databases is possible. To this end,
absolute astrometry is required. After transformation of the
positions of the I and K images to the J reference frame, we
need to find a transformation between this J frame, and an
absolute (α, δ) coordinate system. Examining available cat-
alogues, we find that as of the time of writing, the catalogue
with the greatest number of astrometric objects is the Hub-
ble Guide Star Catalogue (GSC). This catalogue contains
nearly 19 million objects, designed to satisfy the operational
needs of the Hubble Space telescope. The objects are thus
distributed almost evenly across the sky.
A DENIS image is about 12 arcminutes on a side, which
means that, on average, the number of GSC sources found in
an image is 18. However, the coverage is not entirely homo-
geneous. Near the Galactic plane, the regions of particular
interest in this case, there are far fewer sources. For exam-
ple, the 1◦ × 1◦ around ℓ = 0◦, b = 0◦ contains only 307
GSC sources, which corresponds, on average, to 12 per DE-
NIS image. Poisson statistics tell us that there is then a 5%
chance of ≤ 6 objects being found in an image. In fact, the
patchiness of foreground extinction in these regions means
that even Poisson statistics are an unreliable guide as to the
number of objects we may find per frame.
Furthermore, at least second order fits are likely to be
necessary to derive positions based on astrometric points,
simply judging by the fact that a 2nd order transformation
was needed between I and J images (taken by the same
instrument, but using different optical paths, and detectors.)
An average of 12 points per frame is not good enough to
c(cid:13) 0000 RAS, MNRAS 000, 000–000
fit the 12 coefficients required for such a fit – on average,
the solution will be numerically unstable by being forced to
provide 12 coefficients from just about 12 points, and about
half the time (at least), there would not be enough points
at all.
2.5 Digitised Sky Survey
The Digitised Sky Survey (DSS) is available on a set of CD-
ROMS, and also online. In order to produce a list of as-
trometric points for use in performing absolute astrometry
with DENIS, a viable option is to load relevant regions of the
Digitised Sky Survey, source extract the images, and fit the
GSC stars in that image to provide an astrometric database
with a far greater density of objects.
For the region of interest indicated in fig 2, a grid of
1◦ × 1◦ DSS images was overlaid to cover it. Each image
was source extracted using SExtractor, the acceptance cri-
terion being 2σ above background. The aperture 'magni-
tudes' obtained are unimportant, and do not relate linearly
to the true magnitudes, since these images derive from pho-
tographic plates. However, they are useful in limiting the
number of sources used for matching to the GSC, to a few
hundred.
The GSC typically has 300 stars deg−2 in these regions.
As in the case of the IJK images, the same software was used
to make the transformations between (x, y) in the image and
(α, δ) as determined by the GSC stars, projected to the local
tangent plane.
The increase in density of astrometric reference points
results in a density about 50 times greater than when using
the GSC alone, leading to, on average, 500 points per DE-
NIS image. We call this our auxiliary astrometric catalogue
(AAC).
2.6 Using AAC
The I band images, taken at ∼0.8µm, correspond closely in
wavelength to the scans used for the DSS, and matching
sources between them is a relatively straightforward exer-
cise. Using the same technique as before, a 2nd order, two
dimensional polynomial fit is made between the I images
and the AAC extracted from the DSS. The catalogue of I
images used is already transformed to J-image coordinates,
so that the resulting transformation is directly from J →
AAC. Finally this J → AAC transform can be applied to all
the image source coordinates to give 3 separate catalogues
in I,J and K.
3 FINAL CATALOGUE
For each of the I,J and K bands, and for each raster as
shown in figure 2, a final one-band catalogue is made, and
sources duplicated in the overlaps of frames are removed.
The overlapping images are used to give estimates of random
astrometric and photometric scatter. (See later section on
random errors)
For each raster, these large I,J and K catalogues are
matched up. Seven classes of source are distinguished – those
present in IJK, those present in JK only, those in IJ only,
those in IK only (rare!), those only in I, those only in J and
c(cid:13) 0000 RAS, MNRAS 000, 000–000
DENIS and the Inner Galaxy
3
those only in K. The matching is carried out by assigning
initially as the same, objects closer than 3 arcseconds to one
another in different bands.
The output catalogue contains Right Ascension and
Declination for each of the 3 bands separately, a symbol in-
dicating which of the 7 classes the object belongs to, and I
and/or J and/or K magnitudes. The reason for keeping all
3 positions is explained in the following sections.
3.1 Astrometric and photometric precision of
final catalogues
3.1.1 Random errors
The separations between matched objects in a complete
raster (in this case C02), in each colour catalogue separately,
is shown in figure 3.1.1. Some 70%, 80% and 75% respec-
tively, in the I,J and K catalogues, have the same objects
within 1 arcsecond of one another (no mean feat for 3 arcsec-
ond pixels!). The random photometric scatter can be seen
in the left of figure 3.1.1, again by comparison of duplicated
sources from image overlaps. The figure indicates the differ-
ence between magnitudes m1 and m2, measured from differ-
ent images. A fit to the core values every 1
2 magnitude gives
values for the standard deviation of the difference m1 − m2,
and since σ2(m1 − m2) = σ2(m1) + σ2(m2), m1 and m2
coming from the same underlying population, the standard
random error in one measurement of the magnitude is given
in each case by the σ = σ(m1 − m2)/√2. This distribution
is indicated in the right panel of figure 3.1.1.
3.2 Systematic errors
The major systematic errors in magnitude measurement will
be the result of crowding in the fields. The pixelsize used in
the DENIS cameras is 3 arcseconds inevitably making it
difficult to properly resolve sources in the crowded galactic
centre regions.
We make an estimate of the size of this systematic pho-
tometric uncertainty by adding artificial stars to DENIS im-
ages. The images are then processed just as for the unaltered
images, and the extracted magnitudes are compared to the
input magnitudes.
These simulated sources are analgous to individual real
sources – they can be considered as single additional sources
in the presence of very many nearby sources. The statistics
of the difference between the observed and generated mag-
nitudes will be representative of the same statistics for real
sources. (e.g. see Sodemann & Thomson 1997)
A crowded image (at roughly ℓ =0◦ and b=1◦) was used
as the base onto which artificial stars were added. Average
values for the parameters of full-width at half-maximum,
ellipticity, and orientation of the extracted sources in that
image were found, and these were used to generate random
sources. The sources were distributed at random over the
image in steps of 1 magnitude from 11 to 17 in I, from 9 to
15 in J, and from 7 to 13 in K. Each image had 225 stars
added (this is a small fraction when the number of sources
per image is typically 4–5000). The resulting 7 images in
4 M.Unavane et al.
each band were treated as described above for the untreated
images.
The derived catalogues were searched for the artificial
stars to a distance of up to 3 pixels from their input posi-
tions. If found, their magnitudes and positions were noted.
The magnitudes were characterised by a mean offset and a
scatter about that mean offset.
The distribution of magnitude offsets is shown in figure
3.2. At the brightest magnitudes, there is little systematic
shift in the magnitude due to the source extraction proce-
dure. As higher magnitudes are reached, the measured mag-
nitude (calculated using an aperture of 7 pixels diameter as
indicated above) is systematically brighter than the actual
magnitude. This can be understood in the context of the
severe crowding as the result of the flux of nearby sources
entering the aperture.
3.3 Completeness
Using the same simulations as above, the numbers of sources
recovered within a given tolerance of the nominal position
can be assessed and used as a measure of the complete-
ness in these crowded fields. Figure 3.3 shows the fraction of
simulated sources recovered from a typical crowded DENIS
image, for various tolerances from the nominal position (1",
2" and 3").
Deviation (arcsec)
80% Completeness
I
J
K
15.8
17.0
>17
12.6
13.6
13.9
9.9
11.3
12.1
1
2
3
Note that these values are inferior to the expected lim-
its for the DENIS survey in general, since we are confusion
limited in very dense fields.
3.4 Multiband completeness
For ease of data treatment, the colour magnitude diagrams
are made by considering only sources matched between pairs
of images less than a fixed distance apart. In the next sec-
tion, for the analyses, we use 1 arcsecond. This value has
no meaning when considering the statistics only of I, J or
K data, since we are not interested in the absolute precision
with which the source position has been found, but only
with the numbers of such sources at given magnitudes.
However, a tolerance of 1 arcsecond (or some other
value) plays an important role in determining the distri-
butions in IJ, JK and IJK statistics.
Presented below are the results of a Monte-Carlo tech-
nique for determining what fraction of IJ,JK and IJK
matched images will be found if an upper limit is placed
on their separation. The same artificial star experiments as
above are used. We use the distribution of displacement from
nominal position with magnitude (as indicated in a discrete
manner in figure 3.3 for three values of displacement).
Taking the case of J and K as an example, we use the
probability distribution for J and K for the displacement
from the nominal position and build up a 2-dimensional grid
for the fraction retrieved according to the stringent criterion
that the J and K sources lie within a given distance of one
another.
Points are generated according to the measured proba-
bility distribution function, and the acceptance fraction is
deduced using the numbers of J and K points which lie
within 1 arcsecond of each other. The contour plot in fig-
ure 3.4 shows the result of this calculation, plotted on the
same J−K,K plane as for the other plots. Shown in the fig-
ure are contours showing the probability of retrieval of a
J−K pair, if the relative displacement tolerance is set at 1
arcsec. As expected, for objects bright in both bands (ie.
low K and low colour), the retrieval probability is high, and
this probability falls markedly as the K magnitude and/or
colour is increased.
This diagram, smoothed by the fitting of a low-order
two-dimensional polynomial is used below as a convolving
mask applied to model colour magnitude diagrams to enable
a quantitative comparison with the observed diagrams.
In exactly the same manner, an analogous analysis is
carried out for the case of the I−J and J colour magnitude
diagram. The same features are seen (bright blue objects are
almost all retrieved, while dimmer, redder objects are not
so well retrieved).
Finally, in the case of the two colour I−J and J−K
diagram, contributions to a given part of the diagram come
from objects of different magnitudes. At the brightest K
magnitudes, as expected, the regions of low J−K value are
essentially complete, and as we go to fainter K magnitudes,
the completeness gets progressively lower. In a similar way,
a three dimensional (K,I−J and J−K being the variables)
polynomial is used as a mask.
3.5 Comments about completeness levels assigned
The test image used to derive these completeness results was
an image at latitude ∼1◦, where the source density in I,J and
K is the highest of any position in the fields observed. In this
sense, the completeness levels and estimates of photometric
shifts are worst-case estimates. In general, completeness will
be better, and photometric shifts less.
4 LOOKING FOR STRUCTURE
4.1 Overview
The data reduction described in the previous sections results
in a wealth of point source data in the I,J and K bands.
The region observed forms an irregular polygon contained
within l <5◦ and b <1.5◦. The latitude coverage is great-
est nearest to zero longitude, just as the longitude coverage
is greatest near zero latitude. Figure 2 shows the coverage.
The data is severely confused due to the large (3 arcsec-
ond) pixels used coupled with the huge numbers of sources
at low latitudes. The effects of crowding on photometry and
completeness are described above.
In this section, we first present colour-magnitude dia-
grams for different parts of the region covered, and attempt
to gain a qualitative understanding. Many of the unusual
c(cid:13) 0000 RAS, MNRAS 000, 000–000
morphological features in the CM diagrams can be under-
stood in terms of crowding and completeness problems. The
major effect seen in these diagrams is clearly that of a promi-
nent, thin dust lane running through the fields centred near
b=0.
A simple model is constructed to understand the colour
magnitude diagrams resulting from observations of sources
in a stellar disk in the presence of a strong dust layer.
With the aid of latitude-colour diagrams, we establish,
using this model, parameters to describe the dust layer, and
compare synthetic colour magnitude diagrams thus derived
to the observed ones. It becomes clear that disk stars alone
can describe many of the features in the appearance of the
CM diagrams.
This model is then used to deduce the distribution in
distance of the sources seen, in a statistical way, from the
distribution in magnitude that results from the observations.
Magnitude cuts are made in the luminosity functions corre-
sponding to two regions, one dominated by near-disk ob-
jects, and the other by far-disk/bulge objects.
As a test of various Galactic bar models, the cuts above
are used to remove the effect of nearby disk asymmetries,
and to penetrate the far disk and central bulge regions. In J
and K, asymmetries in the same sense as predicted by bar
models are seen at the 3σ level.
4.2 Looking at the data
The dataset derived in the previous section is in three pass-
bands (DENIS I,J and K centred at 0.8µm, 1.25µm and
2.2µm). The number of images processed is 613 in each pass-
band, each covering an area of 770x770 arcseconds. The area
covered is about 17.44 deg2, which means that double cov-
erage occurs for 10.60 deg2 (or about 60%). Excluding the
poor quality data very near the edges, this becomes roughly
50% overlap. It is this which has allowed a good characteri-
sation of random photometric uncertainties.
4.3 Numbers
The total number of sources extracted is some 1.500×106
in I, 1.707×106 in J, and 1.324×106 in K. After removal of
duplicate observations the numbers become 0.748×106 in I,
0.851×106 in J, and 0.659×106 in K.
4.4 Colour-Magnitude and two-colour Diagrams
The following diagrams, (figures 4.4 – 4.4 and later) , indi-
cate, in the bottom left hand corner, a two-colour diagram
with (J−K) on the abcissa and (I−J) on the ordinate. The
top left panel in each case shows the colour magnitude dia-
gram of J−K against K, with the (J−K) scale of the lower
diagram being preserved. Finally, the bottom right panel
shows the (I−J) against J colour-magnitude diagram rotated
anticlockwise by 90◦ to match the (I−J) scale in the two-
colour diagram. This is indicated diagramatically in the ex-
planatory figure 4.4. Note that each diagram is constructed
separately, so that the number of points in the two-colour
diagram will be fewest of all (requiring coincident I,J and K
sources).
c(cid:13) 0000 RAS, MNRAS 000, 000–000
DENIS and the Inner Galaxy
5
The first three sets of diagrams are for 0.2◦ latitude
cuts between −0.1◦ and 0.1◦, 0.4◦and 0.5◦ and 1.0◦ and
1.2◦ (Figures 4.4, 4.4, 4.4). On each diagram is also marked
a reddening line, calculated as in a later section.
Also shown, in figure 4.4, is a cut for a region around
ℓ =3◦ and b=0◦.
Before commenting on these digrams, it is worthwhile to
present colour-magnitude and colour-colour diagrams in the
bands I,J and K for disk type III and type V objects. These
will enable a comparison to be made. Figure 4.4 shows these
diagrams. The data for the absolute K magnitudes are taken
from a study by Garwood & Jones (1987), who produced a
local disk luminosity function. The transformations to I and
J are made according to data tabulated by Zombeck (1990).
Notice that despite the wide range in absolute magnitudes,
both the I−J and J−K colours have a very small intrinsic
spread ( <
The magnitude cuts in I,J and K respectively are at
magnitudes 11,9 and 8 for the bright end (there are satura-
tion problems brighter than this) and at 17,15 and 13 at the
faint end.
∼ 1).
The most marked feature in figures 4.4,4.4 and 4.4 is the
straight line seen in the two-colour diagrams, which follows
the reddening line. This demonstrates more clearly than any-
thing else how important a role interstellar extinction plays
in the interpretation of these diagrams. From figure 4.4, it
is clear that intrinsic stellar properties will not contribute
more than a magnitude of colour shift, so these diagrams
immediately indicate that extinction in the line of sight of
at least AV =10–15 is to be expected. These figures also show
some scatter to the right of the reddening line. This can be
accounted for by crowding effects as will be demonstrated
in a later section. Also, in the two-colour panel in figure 4.4,
the highest latitude field, there is a clear bend at the red-
dest part of the line, so that objects are redder in (I−J), or
equivalently, bluer in (J−K) than expected. Again, this is
an effect which can be understood for this dataset in terms
of crowding problems, and is discussed later.
The other panels in these figures are less easy to inter-
pret. Some morphological features are clear. In the (J),(I−J)
panels, there are clear striations running almost parallel to
the reddening line. The very dominant blue faint part of
the diagram, clear at the lowest latitudes, is much dimin-
ished at latitude ∼ 1◦, while the fainter redder part be-
comes dominant. This can be plausibly understood in terms
of a young, blue, main-sequence population whose presence
will be stronger at the lowest latitudes in the disk. Features
which are 'reddened off the page' become progressively more
apparent at higher latitudes when the optical path through
dust is lessened.
Similar features are evident in the (K),(J−K) diagrams,
with the clear concentration centred at a colour of J−K=4 in
figure 4.4 being shifted to J−K∼3.5 and J−K∼2.2 in figures
4.4 and 4.4 respectively. Again, this is plausibly due to a line
of sight which sees less of the absorbing material in the disk.
Finally, figure 4.4 is included as an example of the in-
homogeneous distribution of absorbing material in the line
of sight. In all three panels, a clear break is seen in one
6 M.Unavane et al.
of the striations. In the (J),(I−J) panel, the striation start-
ing at (J,I−J)=(13,3.6) and continuing to (11,2.0), shows
a clear break in the direction of the reddening line, corre-
sponding to a wall of extinction of AV ∼ 3. The same break
is seen in the (K),(J−K) diagram in the striation starting
at (J−K,K)=(2.0,12) and continuing to (1.2,10). This too is
reproduced in the next sections by means of a model.
5 SIMPLE DISK MODEL
The approach we adopt in order to understand these colour-
magnitude data is to construct a model of the Galactic
disk from which we can generate sources, and make sta-
tistical comparisons to the observed data. Details of this
model, and justification for its simplicity, are given in the ap-
pendix. Only a summary is given here. We employ a model
which consists of a disk, exponential in both the vertical
z-coordinate, and the radial coordinate. The dust is also
represented as an exponential disk with its own scaleheight
and radial scalelength.
For the stellar disk, we use recently derived values
for the scalelength and scaleheight of 2.7 kpc and 0.20 kpc.
(Freudenreich 1996, Kent, Dame & Fazio 1991) The small
region of interest (l <5◦, b <1.5◦) is not very sensitive
to changes in these disk parameters, but rather more to the
dust layer parameters. The dust layer, and its appearance
to us, is characterised by four parameters – the scaleheight
of the dust (zd), the radial scalelength (rd), the height of
the sun above the plane of this dust (z0), and the strength
of extinction in the midplane of the disk (AV in mag/kpc).
These are derived from a fit to the data.
The model output is convolved with the completeness
levels and photometric scatter derived above and is subse-
quently compared to the observations. A sketch of the model
geometry is given in figure 5
5.1 Luminosity Function
The brightest sources visible at near-IR wavelengths, as in-
dicated in figure 4.4, are the later type giant stars.
5.1.2 Main Sequence – Type V
Main sequence stars have intrinsically low near IR lumi-
nosities, mainly due to their small size. What some main
sequence stars lack in size, they make up for in surface lumi-
nosity (i.e. high temperature) so that the brightest of these
stars (O and B) despite having their peak in emission far
from the near IR wavelengths we are interested in, never-
theless show bright magnitudes at these wavelengths.
We adopt the K-band luminosity function described by
Garwood & Jones (1987). Again, we fit a low-order poly-
nomial to fit the colours (J−K) and (I−K) as given by
(Zombeck 1990). A third order polynomial is sufficient to
prevent errors of greater than 0.1 mag. The polynomials are
also shown in table 5.1.
It is assumed in this model that the luminosity function
φ(mK) and the geometrical parameters ρ(x, y, z) are inde-
pendent. This is an oversimplification since it is well known
that the scale height of stars in the disk depends upon the
stellar type (Schmidt 1963). Late type V stars (hence typ-
ically older) are to be found with large scaleheights while
early type (and hence younger) stars have distributions with
smaller scaleheights. The mechanism is clearly diffusive, as it
is thought that stars form in the disk of the galaxy, and over
the course of several revolutions diffuse to distributions with
larger scaleheights. Giant stars, on the other hand, show an
essentially constant scale height with type.
However, the majority of sources seen at the near IR
wavelengths are the giant stars and the older type V stars,
which can all be taken to have a (relatively) large scaleheight
(some 2–300 pc).
As a test of this, below we show the luminosity func-
tions at various heights above the plane, normalised to a
scale height of 200 pc, generated by using the above lumi-
nosity functions, and scaleheight parameters for the different
stellar types. The biggest differences occur between absolute
magnitudes 0 and 3 (depending on waveband) where there
are few sources that feature in the simulations presented.
The three lines shown show the luminosity function at 3 dif-
ferent heights in the plane – z=0 pc, z=200 pc and z=400 pc
(zero, one and tw scaleheights of the oldest population).
We adopt a luminosity function for zero height (i.e. scale
heights the same for the different stellar types) for the rea-
sons given above.
5.1.1 Giants – Type III
5.2 Extinction coefficients for the DENIS filters
As a source for a luminosity function for the type III (gi-
ant) stars, we refer to Garwood & Jones (1987) , who ob-
servationally determined a local luminosity function in the
K-band. Colours for local objects are taken from Zombeck
(1990), who collates infra-red field star colours from a va-
riety of sources. Colours for each of the types of source are
given, from which a polynomial fit is made to determine an
analytical transformation from the absolute magnitude in K,
mK, to the absolute magnitudes in I and J. (mI and mJ ). A
second order polynomial was found to be sufficient in these
case, and the greatest discrepancy with the tabular values
was less than 0.1 mag. The polynomials used are tabulated
in table 5.1.
Since large values of extinctions along the line of sight play
a major role in determining the appearance of the colour
magnitude diagrams in the plane of the galaxy, it is impor-
tant to establish precise values for the extinction coefficients.
We use the tabulation by Mathis (1990) for the relative ex-
tinction in magnitudes as a function of wavelength. Between
1.25 µm and 3.4 µm, we parametrise his tabulation as
A(λ)
A(J ) = 1.484 − 5.60109x + 8.395624x2 − 4.5947083x3
where x = log10(λ/µm). The error in this least-squares
3rd order fit is not more than 4%.
Referring to Copet et al. (1997), we use the instru-
ment+sky response profiles in each of the I,J and K bands,
and convolve them with the Mathis data to give the follow-
ing values for A(X)/A(1.25µm):
c(cid:13) 0000 RAS, MNRAS 000, 000–000
Band (X) A(X)/A(1.25µm) A(X)/A(V )
I
J
K
1.968
0.994
0.396
0.554
0.280
0.112
For comparison, the Mathis (1990) values of A(λ)/AV
for wavelengths 0.90µm, 1.25µm and 2.2µm, corresponding
to Johnson's I,J and K, are 0.479, 0.282 and 0.108.
5.2.1 Application to model
The chosen functional form for the dust layer is that of a
disk exponential in both radial and vertical components. The
amount of the dust is represented in terms of its absorption
as:
X(r, z) =∝ e−z/zd−r/rd
where X is a measurement of AV in mag/kpc, and r and
z are galactocentric cylindrical coordinates. The function is
normalized by setting dX/dρ = X0 , where ρ is distance
along the line of sight and X0 is the local extinction per
unit distance. The functional form for the dust is integrated
along the given line of sight to yield a function AV (ρ) which
represents the amount of extinction in the line of sight up
to a distance ρ from the sun.
5.3 Summary of model
In summary, the model consists of the following steps. De-
tails of the Monte-Carlo method used to generate stars are
given in an appendix.
(i) For given lines of sight, a probability distribution func-
tion of distance is generated. An extinction/distance curve
is also generated by integrating the function representative
of dust along the line of sight.
(ii) By combining functions in the three wavebands with
the distance distribution, a limiting parameter w0 is found
to speed up the Monte Carlo process (see Appendix).
(iii) Observable distance/stellar class pairs are generated,
and K magnitudes are calculated from the luminosity func-
tion.
(iv) Analytical representations in terms of the K magni-
tude are used to give the corresponding I and J magnitudes
(v) To each absolute magnitude, the modified distance
modulus is added ( 5lg(r/kpc)+10+AX, where X=I,J or K).
(vi) To each magnitude, is further added a random scat-
ter in magnitude as determined by the artificial star exper-
iments above.
(vii) A systematic magnitude offset to simulate crowd-
ing/extraction effects is added.
(viii) The source is rejected if its magnitude falls outside
specified limits.
(ix) Steps 3–8 are repeated until the specified numbers of
sources have been generated.
6 MODEL OUTPUT
Figures 6, 6 and 6 show the results obtained from running
the model for three different latitudes b=0◦, b=0.5◦, and
b=1.1◦, which may be compared with figures 4.4, 4.4 and
c(cid:13) 0000 RAS, MNRAS 000, 000–000
DENIS and the Inner Galaxy
7
4.4 respectively. Figure 6 shows the result of a simulation
where a cloud of extincting material with AV =3 has been
placed between 1.5 and 2.5 kpc from the sun in the line of
sight. The raw output from the model has been treated with
the systematic shifts and random scatter in photometry due
to crowding, and the colour-magnitude and colour-colour
diagrams have been convolved with the completeness masks
derived at the end of the first section.
We stress again that these models are constructed for a
disk only, and it is expected that whatever structure remains
in the CM diagrams after the removal of this disk structure
is attributable to the bulge. We show in figure 6 a quanti-
tative comparison of the model for latitude b = 0.5◦and the
observations (figures 6 and 4.4). The figure shows the ratio
of sources observed to those predicted in the model as a con-
tour map over the (J−K)-(I−J), (J−K)-(K) and (I−J)-(J)
planes. The dotted contours indicate small variations likely
due to patchy extinction, for which account cannot be made
in this model. The three dotted contours represent ratios of
0.5, 1.0 and 2.0. The solid contours represent number ratios
from 4 up to 12. It is clear that in the (J)-(I−J) diagram,
there is very little difference between the model and the ob-
servations. As expected for a diagram limited by the I band,
distance penetration is low and only the disk is seen. How-
ever, for the (J−K)-(K) diagram, there is clearly an excess
population of sources peaking at a colour of J−K∼3.5. This
may be identified with a reddened giant branch in the bulge,
which is in accord with the expectation that the power of
the J and K bands to penetrate dust will allow the bulge to
be seen. There remains a hint of this effect in the reddest
sources of the two-colour diagram (J−K)-(I−J), but as for
the (J)-(I−J) diagram, this is limited by the limited distance
penetration of the I band, and will be well represented by
a disk only. The ratio of bulge to disk sources seen here at
b = 0.5◦of <
∼ 12 is in good agreement with the result ( <
15) we derive in a later section from published bulge/disk
ratios.
∼
The derivation of the parameters for the exponential
dust-disk is detailed in the next section.
7 DERIVING THE DUST PARAMETERS
7.1 Elementary characterisation of dust layer
It is clear by looking even at integrated light images of the
Galactic Central regions (e.g. Madsen et al., 1986) that there
is a prominent dark band in the plane of the galaxy. This is
due to the dust which pervades the disk, and causes light to
be attenuated, especially at short wavelengths. The feature
is diminished at longer wavelengths.
If we consider the dust to be a uniform plane of finite
thickness 2d, extending in the direction of the plane of the
galaxy, with the sun centrally placed, we can easily calcu-
late that the path length in the dust layer is given, in terms
of the galactic latitude, b, by d cosec b. This optical path
retains the same functional dependence on b if the uniform
8 M.Unavane et al.
layer is replaced by a dust distribution with an exponen-
tial dependence (For a vertical scaleheight of zh, the result
becomes just zh cosec b).
Figure 7.1 shows the mean colours for cuts of height
0.2◦ covering the full range of the dataset from 5◦ to −5◦.
The colour shown is found by taking the mean value of all
sources in the relevant colour-magnitude diagrams.
The variation with mean colour with latitude is mainly
symptomatic, at these low latitudes, of extinction in the
plane.
The major features in figure 7.1 are the clear rise in
mean J−K colour towards the plane, and the equally clear
fall in mean I−J colour towards the plane. These can both
be understood in terms of different degrees of reddening and
different populations sampled.
For the J−K diagram, the penetration into the dust
is limited by the J waveband completeness. The dominant
sources, seen to large distances, are type III giants, and the
closer the approach to the plane, the greater the amount of
dust in the line of sight, leading to a peaking in the mean
colour towards the plane.
For the I−J diagram, the same effect will clearly be
taking place, but the limitation in this figure is the effect
of the dust convolved with completeness in the I waveband.
Towards the plane, the distance observed is low, and the
sources are dominated by nearby, main-sequence, blue stars.
These have an intrinsic colour from I−J=−0.2 to 0.5. As we
look away from the plane, due to the decreased dust in the
line of sight, we see to greater distances, and sample the
giant stars visible to a much larger distance. The figure is
effectively the same as the J−K figure (as suggested by the
tails), with a large wedge removed from the middle due to
the limited distance penetration at the wavelength of the I
band.
It is also clear in the figure that there is an asymmetry
about b=0 for the mean J−K colour of extracted sources.
The peak, in fact, appears at b∼0.15◦. This can be at-
tributed to a non-zero height for the sun above the plane
of the local dust.
We can thus, using the I−J diagram, place a very ap-
proximate limit on the height of the extinction layer by not-
ing that the abrupt change occurs at ∼ ±1◦, and that the
brightest main-sequence stars may are visible to distances
of 2–3 kpc (Neckel & Klare 1980). One degree at this dis-
tance corresponds to about 40–50 pc, and can be seen as an
indication of the scaleheight of the dust.
Similarly, a simple estimate of the displacement of the
sun can be made by noting that in J and K bands, the
typical colour of objects seen to any significant distance in
the (dusty) plane is between 0.6 and 1.0 (type III objects –
see figure 4.4).
Indicated on the left in figure 7.1 are cosec law fits of
the tails of the distribution of extinction, using an intrin-
sic colour of J−K=1.0. The fits correspond to scaleheights
differing by about 40%. i.e.
zd − z0
zd + z0
= 1.4
where zd is the scale height of the dust, and z0 is the
displacement of the sun above the plane. The solution we
obtain is that z0/zd ∼ −0.17. The fit is very insensitive to
the intrinsic colour adopted. (e.g. changing (J−K)0 from 1.0
to 0.6 changes z0/zd from −0.17 to −0.15).
Using the scale height of about 40–50 pc estimated
above, we deduce that the sun lies about 7 pc below the
local galactic dust plane.
And finally, noting that the reddest J−K colour of
sources near b=0 is some 2.0–2.4 magnitudes redder than the
intrinsic colour (corresponding to an AV of 18–22), we can
estimate, assuming DENIS sees some giants at the Galac-
tic centre at about 8 kpc distance, that AV is on average
roughly 2.5 mag/kpc.
These estimates for the scale height, distance from
plane, and AV /kpc value can be refined in a model-
dependent way.
8 MODEL DEPENDENT DERIVATION
The dust layer, in this simple model, is characterised by
three parameters : the scaleheight of the assumed exponen-
tial profile (zd), the radial scalelength (rd) and the local
rate of extinction (AV in mag/kpc). In addition, the dis-
placement of the sun from this midplane must be included
(z0).
The rough method described above provided initial es-
timates for the values of these three parameters – zd,Av in
mag/kpc and z0. For the fourth parameter rd, we use a value
of 2.7 kpc as for the stellar disk, as a starting guess.
The model described above was set up with the adopted
parameters for the stellar disk and dust scalelength, and the
three parameters to be determined by a fit to the data (i.e.
rd,zd,Av in mag/kpc and z0).
A 5×5×5×5 grid of synthetic colour-latitude diagrams
was made by generating stars according to the model de-
scribed above, using all the possible combinations of the
following values of the three parameters:
• rd=2,4,6,8,10 kpc
• zd=20,30,40,50,60 pc
• z0=−25,−20,−15,−10,−5 pc
• AV = 0.5,1.0,1.5,2.0,2.5 mag/kpc
The resulting point source data were convolved with the
two-dimensional completeness functions derived in section 3
for the I−J and J−K data, and were subsequently used to
generate colour-latitude diagrams by linear interpolation be-
tween points in this data tesseract. The statistic minimised
is the sum of the squares of the difference between observed
and model colours summed over both I−J and J−K colours.
The ranges in latitude in each case were limited by practical
factors.
In J−K, the ranges for comparison were limited to the
tails of the distribution beyond J−K=1, so as to avoid the
strongest bulge contamination. The tails of the J−K distri-
bution are expected to sample to large distances, thereby
averaging out, to some extent, local inhomogeneities in the
dust distribution.
The same cannot be true for I−J - this will be severely
affected by local inhomogeneities due to the limited distance
sampled by the I waveband. Contrary to the situation in
c(cid:13) 0000 RAS, MNRAS 000, 000–000
the J−K figure, we expect the lowest latitudes to be lit-
tle, if at all, contaminated by bulge sources, and a fit there
is appropriate. We do, however, exclude the highest lati-
tudes (b>1) due to the known presence of very nearby (few
100 pc) dusty star forming complexes at positive latitudes
which may severely bias the result (e.g. ρ Oph)
The 18 points thus chosen for the fit are the data points
at latitudes b = −1.2,−1.0,1.0,1.2,1.4,1.6,1.8◦ in J−K, and
all points between −1.2◦ and 1.0◦ inclusive in I−J. All were
given equal weight in a least squares fit. The statistic X
used was simply the sum of squares of differences between
the model and the observations:
i.e.
X =
N
Xi=1
(ci − c(bi; a1, a2...aM ))2
where ci denote the observed mean colour values and
c(bi; a1, a2...aM ) denote the model derived mean colours. bi
is the latitude associated with that colour point, and a1...aM
are the M parameters associated with the model. In the
present case, M=4, and a1,a2,a3 and a4 are local AV in
mag/kpc,zd,rd and z0.
The number of data points, N, is 18 since the summation
is over the limited set of 18 points in the I−J colours and
the J−K colours from −1.2◦ to 1.8◦ inclusive in 0.2◦ steps.
This statistic can be used to estimate the standard de-
viation associated with each data point by means of the
following formula (Press et al., 1992):
σ2 = PN
i=1(ci − c(bi; a1, a2...aM ))2
N − M
Minimising the statistic X with respect to variations
in AV , zd,rd and z0, we find a value X=0.484 occurring at
values AV /kpc = 1.40, zd = 40 pc, rd = 3.4 kpc and z0 =
−14.0 pc. An estimate for σ is thus 0.13, which agrees well
with inspection of the plots. A value can thus be assigned
to χ2 by using the standard formulation:
χ2 =
N
Xi=1
(cid:18) ci − c(bi; a1, a2...aM )
σ
(cid:19)2
Subsequently, a 1σ estimate of the uncertainties in the
parameters derived can be obtained by looking for the vari-
ations in those parameters which give rise to an increase in
χ2 of ∆χ2 = 1.00 (Press et al., 1992). This corresponds to
a value ∆X of 0.0346.
The result obtained in this way is:
• AV /kpc = 1.40 ± 0.11
• rd = 3.4 ± 1.0 kpc
• zd = 40 ± 5 pc
• z0 = −14.0 ± 2.5 pc
The model and observations are compared in figure 8.
The fit to the wings of the J−K distribution is very good,
which is as expected since the J−K sample is expected to
sample a large path length of the disk, and any small scale in-
homogeneities, such as the conspicuous +ve latitude nearby
dust regions, are averaged out. The fit to the I−J data is less
convincing as local structure in the interstellar dust strongly
biasses the mean colours.
c(cid:13) 0000 RAS, MNRAS 000, 000–000
DENIS and the Inner Galaxy
9
Note that the above method does not use a truly in-
dependent value for σ. This is something which cannot be
readily defined for the dataset. The uncertainties are not
measurement error, but are due to the random nature of
the extinction in the line of sight. As a consequence, the
uncertainties in the results above reflect the uncertainties of
this best fit within these limitations. They do not in any way
indicate to what extent the functional forms are justified.
These dust distribution parameters are used in the
model of the galaxy described above.
8.1 Comparison to other results
A whole sky fit to the COBE/DIRBE data has been made by
Freudenreich (1996). He excluded the difficult central region
(∼ 40◦ × 30◦) as well as other parts of the plane, and fitted a
28 parameter model, constraining simultaneously the scales
and intensity of the dust layer, and the stellar disk.
Converting his dust layer parameters to the units used
here, he gives:
• AV /kpc = 1.53±0.01
• rd = 3.85 ± 0.10 kpc
• zd = 46±1 pc
• z0 = 15.55±0.23 pc
Freudenreich uses a functional form sech2(z/2zd), as op-
posed to the form exp(−z/zd) used here. These two forms
are equivalent for z ≫ zd (within a numerical factor of 4),
but importantly, near z = 0 in the plane, these functions
differ significantly. At height zero, they differ by a factor of
4, and at heights of zd and 2zd, still by factors of 1.8 and
1.3. The fits made by Freudenreich exclude regions within a
few degrees of the plane in most places, and he chooses the
sech2 functional form because it has some basis in theory as
the density law of an isothermal self-gravitating disk (van
der Kruit & Searle 1981). We find here roughly the same
value of the parameter for scaleheight, but we assume an
exponential form for the dust layer.
The values obtained for the intensity of local extinction,
and the scale length of the dust layer are in good agreement.
Integrating this model in a line of sight towards the centre
results in a value for total AV of 56. The wide uncertainty
in rd results in large uncertainties on this value between 40
and 110. Many observations towards the galactic centre (e.g.
Becklin & Neugebauer 1968; Catchpole, Whitelock & Glass
1990) seem to agree on a value of extinction in the line of
sight towards the very centre of the galaxy as AV ∼ 40, and
our results are consistent with this.
The value we find for the distance of the sun from the
local dust plane suggests that it lies below the plane of this
dust (i.e. towards the South Galactic Pole). Indeed, look-
ing at the surface photometry map produced using DIRBE
data at 1.25µm (Weiland et al. 1994 – their figure 1), we
see that the part of the bulge at negative latitudes appears
brighter than that at positive latitude. Taken at face value,
this would suggest a greater path length of dust towards the
northern part of the bulge, and hence a position below the
dust plane. However, as is clear from the same map after
correction for extinction (Weiland et al. 1994), the most ap-
propriate position for the sun is above the plane of the disk.
This is further corroborated by many other studies using
widely different methods. For example, Cohen (1995) finds
10 M.Unavane et al.
a distance of 15 pc above the plane by comparing north and
south Galactic pole star counts. Figure 3b of Freudenreich
et al. (1994) shows that the DIRBE 240µm emission lies
below the Galactic plane, and though the relation between
240µm flux and dust is not well calibrated, this nevertheless
suggests a displacement above the plane for the sun. Binney,
Gerhard & Spergel⋆ (1997) also find a value of 14 pc above
the plane after modelling the inner Galaxy. Several more
studies (e.g. Conti & Vacca 1990; Toller 1990) all agree on
a value of ∼10–20 pc above the plane. Clearly, all work to
date is in agreement that the sun lies above both the Galac-
tic stellar and dust planes.
Our result is not sensitive to the stellar plane, but rather
to the dust plane – the stellar plane acts only as a lumi-
nosity source of near-constant colour. The simple model we
construct imposes a global symmetry, and is fitted based on
data from a very limited set of directions, unlike most of
the other references cited above. We did not expect to de-
rive globally reliable parameters which describe the three-
dimensional complexity of dust distribution present in the
disk/bulge, and indeed the position we derive for the sun
is a demonstration of this fact. There is clearly asymmetry
in the local extinction (for example, there is substantially
greater extinction present north of the Galactic plane, much
associated with the nearby ρ Oph star-forming region, than
to the south). Presumably it is this which the analysis of the
DENIS data is sensitive.
9 PECULIARITIES IN THE
COLOUR-MAGNITUDE DIAGRAMS
Some of the peculiarities described above can be understood
in a model dependent way.
9.1 Asymmetric Scatter to the right of the two
colour line
Scatter can be seen to the right of the two-colour line as in
figures 4.4,4.4,4.4. If real, these would represent objects with
very high intrinsic J−K colour, compared to I−J colour.
Figure 4.4 indicates that normal stars do not appear like
this. It is possible that some of these are dust-enshrouded
stars, which shine brightly at longer wavelengths, but are
very much more obscured at shorter wavelengths. But be-
fore turning to these astrophysical explanations, we shall
consider the data reduction.
Figure 9.1 shows the two-colour diagram for a typical
crowded field, when the matching radius, within which all
three images (I,J and K) must fall in order to be accepted,
is varied between 0.3" and 3". It is clear that as the radius
is decreased, spurious points scattered to the right of the
line become fewer. If the matching radius for finding com-
mon sources between the 3 images is left too large (e.g. 3"),
⋆ Note that the preprint version (Binney, Gerhard & Spergel
1996) and the published version of the article by Binney, Ger-
hard & Spergel (1997) differ in the position ascribed to the sun.
The preprint consistently states a position of 14 pc below the disk
plane for the sun, while the published version states a value of
14 pc above the plane.
there are many sources apparently very far from the line. Re-
ducing the matching radius to 1" loses some 10% of sources
of deviation zero, but loses over 80% of those with deviation
2 magnitudes.
9.2 Bent two-colour line
In looking at the highest latitude colour-colour diagrams,
(such as figure 4.4), we notice that I−J colours are too red
compared with J−K colours at the same reddening. The
two-colour line is effectively bent upwards.
AGB stars are expected to be found in this region of the
colour-colour diagram. (Groenewegen, 1997) But another ar-
tifact of data processing can explain at least part of these
bends, by attributing them systematic errors in photometry
caused by crowded fields. The effect is well reproduced in
the model when the systematic offsets in magnitude are in-
cluded. Figure 6 shows a clear upward bend in the reddest
parts of the two colour diagram. This can be understood in
terms of the systematic effects in magnitudes. In particu-
lar, the J-band magnitudes show a very large offset for the
faintest magnitudes retained, larger than those of the cor-
responding I and K magnitudes. The offset, for the reasons
described above, causes the magnitudes to be too bright, or
numerically, too low. This means that J−K is too low, and
I−J is too high. This is precisely the effect seen in both the
simulation and the observations.
9.3 Broken striations
Figure 4.4 shows a set of diagrams where striations in the
colour-magnitude diagrams are fractured in the direction of
the reddening line. This too is well understood in terms of
the model by allowing a 'wall' of extinction in the line of
sight, as demonstrated by figure 6.
10 DISTANCE DISTRIBUTIONS
The value of the simulated datasets, which are a good match
to the observations, is that they can be interrogated for
distance information, which will help to deduce the three
dimensional distribution of stars in the inner galaxy. The
information obtained in this way is summarised for the sim-
ulation at b = 0.5◦in figure 10. This particular latitude has
been chosen to enable the dust to be used to our advantage
to separate source in the near disk and in the inner disk
and bulge. At higher latitudes, the extinction due to dust is
much less and penetration is good in all three bands. Near
and far sources are not spread out by the reddening in the
line of sight to a great extent. On the contrary, at very low
latitudes approaching zero, the extinction is so severe that,
given the magnitude limits of the system, it is not possible to
see past the nearby disk population into the bulge regions,
except possibly at K band.
From figure 10, it is clear that in the I band, due to
the inpenetrability of the local dust at short wavelengths,
even the faintest magnitudes fail to penetrate very far from
the sun. The situation changes at J band, where while pen-
etration is low ( <
∼ 8 kpc) at the lowest magnitudes (J=9.0–
11.0), there is some penetration of the bulge regions at the
c(cid:13) 0000 RAS, MNRAS 000, 000–000
fainter magnitudes (J=13.0–15.0). At the K band, the pen-
etration is even greater. Even the very brightest magnitudes
may penetrate to the bulge region, but at the fainter mag-
nitudes, the source counts are dominated by stars near the
bulge. Remember that this simulation includes only a disk
component, so when the model suggests penetration in as
far as the bulge, the counts at that magnitude are likely to
be dominated by bulge objects. This is shown in the con-
tour plot above (figure 6), where the ratios of bulge to disk
objects will be at <
∼ 12.
11 DISK/BULGE ASYMMETRIES
11.1 Bar models
There has been recent interest in the possibility of a kilo-
parsec scale bar at the centre of our galaxy. Various lines of
evidence suggest the presence of a bar, and though they dis-
agree on the exact parameters which best describe the form
of the bar, they agree that the major axis is oriented towards
the first quadrant. Methods employed include gas dynamics
(Binney et al., 1991; Blitz & Spergel 1991), and modelling
of integrated light distributions from the COBE/DIRBE ex-
periment (Dwek et al., 1995; Binney et al., 1997). The asym-
metries expected in projection are such that number counts
at equally positive and negative longitudes should in general
be greater at positive longitudes for large longitudes, and at
negative longitudes at small longitudes.
Integration along the line of sight in such models for
equal positive and negative longitude pairs results in differ-
ent number counts for the two lines of sight. In figure 11.1,
the number count asymmetries predicted by various models
are shown. The effect of luminosity function will be small
at these longitudes. The difference in distance modulus to
the main concentration of the bar will be <
∼ 1.0 magnitude,
since the separation in angle is at most 10◦.
In practice, the amplitudes shown here will only be
realised when a tracer population which samples only the
bulge is used. In the present case, there is much disk con-
tamination, and we expect, in general, a lower signal to be
seen. An estimation of the extent of this dilution is given
by integrating a recent model fit to the bulge/bar and the
disk of the galaxy by Binney et al. (1997). For the purposes
of this discussion, the contrast between disk and bulge in
the inner disk is important. Using the model there, we find
that for sources inward of 3 kpc from the centre, observed
from the sun, and for longitudes of <
∼ 5◦, the number count
contrasts vary strongly only with latitude (because of the
thinness of the disk).
b(deg)
0.2
0.08
0.4
0.06
0.6
0.05
1.0
0.03
These ratios are small, and in the light of other uncer-
fdisk/total
0.0
0.12
0.8
0.04
tainties present in this analysis, may be neglected.
In order to optimise detection of the inner-disk/bulge
asymmetries in our data, we can use the information fur-
nished by figure 10. We then test this for left-right asymme-
tries in the disk/bulge at various distances. We must first
assume the similarity of disk luminosity functions at equal
positive and negative longitudes for a give latitude of obser-
vation. This is not a contentious assertion, as the difference
c(cid:13) 0000 RAS, MNRAS 000, 000–000
DENIS and the Inner Galaxy
11
in the lines of sight differs, in all the following cases, by less
than 10◦, so systematic age or metallicity differences are not
expected.
The effect still remaining in the data which prevents
immediate comparison between number counts at equal and
opposite longitudes is caused by patchiness in the extinction,
as there is no reason to expect this to be systematic. To
minimise this effect, we use the model to identify magnitude
ranges dominated by disk objects, and identify and correct
for any associated (foreground) asymmetry.
That is, cuts in magnitude are chosen which are dom-
inated by disk objects. A fit is made at these magnitudes
between equal negative and positive longitude pairs. Any
additional asymmetries remaining at fainter magnitudes will
(in the case of the J band and K band) contain some signal
of asymmetries in the inner disk or in the bulge. The I band,
according to the model distances, should serve as a control,
since it is not expected to penetrate very far, and the num-
ber counts seen should, after this correction for differences
in extinction, show equal values at positive and negative
longitudes.
12 RESULTS OF THIS EXPERIMENT
The magnitude limits chosen for the cuts are as follows:
Band
Fit region Test region
"Disk"
"Bulge"
I
J
K
11.0–14.5
9.0–11.0
7.5–9.0
14.5–17.0
11.0–13.0
9.0–10.5
Figure 12 shows the results for b = 0 when cuts are
made as described above to match the local disk luminosity
function. The contrasts shown are for the fainter magnitudes
as indicated above.
12.1 Deviation from unity
Do the contrasts deviate, in the mean, from unity? Taking
the mean values of the contrasts in the three cases and find-
ing the deviation of the mean we find the following:
Band Mean, µ
I
J
K
0.961
1.105
1.118
σµ
0.026
0.037
0.039
(µ-1)/σµ
−1.5
2.8
3.1
The suggestion is that the I band deviates insignifi-
cantly (no more than 1.5σ) from unity. This is as expected,
since the I band counts, though patchy, once matched for a
pair of directions according to the local distribution of the
brightest sources, show no further differences at faint mag-
nitudes. The distance model above suggests that the pene-
tration is not nearly deep enough into the disk to allow the
innermost parts of the disk, or the bulge, to be sampled.
The J band shows a more significant deviation from
unity in the ratio of counts (∼ 2.8σ). The distance model
in this case suggests that penetration is deeper, and may
12 M.Unavane et al.
reach the inner disk and bulge. Indeed, the contrasts seen
suggest that counts at negative longitudes are greater than
those at corresponding positive longitudes. Similarly, in the
case of the K-band number counts, we see a similar asym-
metry. In this case, there is a marked dip in the contrast at
a longitude of about 2 degrees. This is most plausibly due
to a mixture of structural and extinction effects close to the
centre of the galaxy. In the analysis here, we try to remove
asymmetries in the nearby disk caused by dust to show up
asymmetries in the inner disk or bulge. It is clear from the
12CO(1J →0 J) intensity plot shown in figure 12.1 (Dame
et al. 1987) that there are distributions of material towards
the centre of the Galaxy with intensities differing by several
orders of magnitude over a very few degrees. In this light it
is not surprising that the contrasts seen in the K number
counts would not show any clear bar-like signature even if
one existed. One approach to combatting this problem is to
obtain multicolour information for sources in these regions
and deredden each on a point by point basis. This technique
is used in paper 2 to treat K and L band data.
The results here are at best inconclusive. The contrast
seen in the K band is always, in the mean, greater than unity,
suggesting that the asymmetric inner disk dust effects traced
in figure 12.1 are a perturbation on a net greater negative
longitude count compared with positive longitude counts.
A similar effect is not seen clearly in the J-band, which
suggests that the realm of this central asymmetry, be it
structural or due to dust, is not reached at this shorter wave-
length.
13 CONCLUSIONS
We have derived some techniques to extract photometric
and astrometric information from DENIS images in crowded
fields and to characterise its deficiencies. We have used this
dataset, covering a part of the region within l <5◦, b <1◦,
to construct a model of the Galactic disk, and to fit param-
eters for a model of the dust layer. Using this model, we find
that the large numbers of very red sources in the observed
colour-magnitude diagrams, not reproduced by the model,
can be understood as bulge giant branch stars. We assume
the symmetry of the structure and luminosity function in
the disk in directions of equal and opposite longitudes and
fit number counts for a bright cut in magnitude (correspond-
ing, according to the model, to near disk sources), and look
for asymmetries in the fainter number counts (correspond-
ing to inner disk and bulge objects). We find that in the I
band, there is no asymmetry at fainter counts, consistent
with the model-based expectation that the I band does not
penetrate very far into the disk. In J and K bands, there
is ∼3σ evidence for a ratio of negative longitude to positive
longitude number count which is greater than one. This is
consistent with the expectation from bar models.
This large scale statistical approach allows, to some ex-
tent, the possibility of 'averaging out' localised anomalies in
extinction which are common in the plane, and towards the
centre of the galaxy. However, as can be seen especially in
the K-band contrasts, number counts can be seriously af-
fected by the distribution of dust regions near the centre
of the Galaxy itself. In the present set of observations, it is
not clear that the J-band data penetrates the densest dust
regions reliably. One possibility is to use longer wavelengths
where the extinction coefficient AX /AV is lower still, so that
dust is a less severe problem. This method is explored in pa-
per 2, where we combine DENIS K data and UKIRT nbL
data (3.6µm, AnbL = 0.047 AV ).
14 ACKNOWLEDGEMENTS
The DENIS project is supported by the SCIENCE and the
Human Capital and Mobility plans of the European Com-
mission under grants CT920791 and CT940627, the Euro-
pean Southern Observatory, in France by the Institut Na-
tional des Sciences de l'Univers, the Education Ministry and
the Centre National de la Recherche Scientifique, in Ger-
many by the State of Baden–Wurttemberg, in Spain by the
DGICYT, in Italy by the Consiglio Nazionale delle Ricerche,
in Austria by the Science Fund (P8700-PHY, P10036-PHY)
and Federal Ministry of Science, Transport and the Arts,
in Brazil by the Fundation for the development of Scientific
Research of the State of Sao Paulo (FAPESP).
MU would like to thank GS at the Observatoire de Paris
for his hospitality during visits there, as well as Francine
Tanguy, Jean Borsenberger and Lionel Provost for their help
with access to the DENIS archive. MU acknowledges the fi-
nancial support of the Particle Physics and Astronomy Re-
search Council.
REFERENCES
Becklin E.E., Neugebauer G., 1968, ApJ, 151, 145
Bertin E., Arnouts S., 1996, A&AS, 117, 393
Binney J., Gerhard O.E., Stark A.A., Bally J., Uchida K.I., 1991,
MNRAS, 252, 210
Binney J., Gerhard O., Spergel D., 1996, MNRAS, preprint astro-
ph/9609066
Borsenberger J., 1997, in The Impact of Large Scale Near-IR sur-
veys, eds. F.Garz´on et al., (Kluwer), p181
Blitz L., Spergel D.N., 1991, ApJ, 379, 631
Catchpole R.M., Whitelock P.A., Glass I.S., 1990, MNRAS, 247,
479
Cohen M., 1995, ApJ, 444, 874
Conti P.S., Vacca W.D., 1990, AJ, 100, 431
Copet E. et al., 1997, A+AS, submitted
Dame T.M., et al., 1987, ApJ, 322, 706
de Jong R.S., 1996, A&AS, 118, 557
Deul E., 1997, in preparation
Dwek E. et al., 1995, ApJ, 445, 716
Epchtein N., 1997, in The Impact of Large Scale Near-IR surveys,
eds. F.Garz´on et al., (Kluwer), p15
Freudenreich H.T. et al., 1994, ApJ, 429, L69
Freudenreich H.T., 1996, ApJ, 468, 663
Garwood R., Jones T.J., 1987, PASP, 99, 453
Genzel R., Hollenbach D., Townes C.H., 1994, RPP, 57,417
Georgelin Y.M. & Georgelin Y.P., 1976, A&A, 49, 57
Glass I.S., Catchpole R.M., Whitelock P.A., 1987, MNRAS, 227,
373
Groenewegen M.A.T., 1997, in The Impact of Large Scale Near-
IR surveys, eds. F.Garz´on et al., (Kluwer), p165
Hammersley P.L., Garz´on F., Mahoney T., Calbet X., 1995, MN-
RAS, 273, 206
Kent S.M., Dame T.M., Fazio G., 1991, ApJ, 378, 131
Madsen C., ESO, Laustsen S., 1986, ESO Messenger, 46, 12
c(cid:13) 0000 RAS, MNRAS 000, 000–000
DENIS and the Inner Galaxy
13
Mathis J.S., 1990, ARA&A, 28, 37
Neckel T., Klare G., 1980, A&AS, 42, 251
Press W.H., Teukolsky S.A., Vetterling W.T., Flannery B.P.,
1992, Numerical Recipes, (CUP)
Schmidt M, 1963, ApJ, 137, 758
Skrutskie M.F. et al., 1997, in The Impact of Large Scale Near-IR
surveys, eds. F.Garz´on et al., (Kluwer), p.25
15.1 Distance
First of all, the function p(r) is converted to a function of
distance modulus, Dm. Noting that
p(r)dr = p(Dm)dDm
Sodemann M., Thomsen B., 1997, A+A preprint, astro-
and that
ph/9704282
Toller G.N., 1990, in IAU Symp. 139, Galactic and Extragalactic
Background Radiation, ed S.Bowyer & C. Leinart (Dordrecht:
Kluwer),21
van der Kruit P.C., Searle L., 1981, A&A, 95, 105
van der Marel R.P., de Zeeuw P.T., Rix H.W., Quinlan G.D.,
1997, Nature, 385, 610
Wainscoat R.J., Cohen M., Volk K.,Walker H.J, Schwartz D.E.,
1992, ApJS, 83, 111
Zombeck M.V., 1990, Handbook of Space Astronomy and Astro-
physics, 2nd Edition, CUP
Weiland et al., 1994, ApJL, 425, L81
15 APPENDIX – THE GALAXY MODEL
CALCULATIONS
A Monte-Carlo method for generating stars in a Galactic
model is mentioned above. Details of the method employed
for optimising the Monte-Carlo process are given below.
For a given line of sight, a probability distribution is
generated, p(r), corresponding to the number of sources
seen, per unit distance, due to the geometrical components
included in the model. Clearly, because of the central con-
centration of the galaxy, this will mean that the peak in
this distribution will lie at approximately r0 from the sun
towards the Galactic Centre (GC), where r0 is the sun-GC
distance.
Also for this line of sight, based on the dust model, a
function AV (r) can be generated, which gives the extinction
to any distance in that line of sight.
The luminosity function (LF) used has logarithmically
more faint sources than bright sources.
Thus, if, independently, a random position is chosen ac-
cording to p(r) and a random stellar type is chosen according
to the LF, the resulting pair will most likely be a faint, main
sequence star near the Galactic Centre – which will not be
visible.
The Monte Carlo approach used to generate random
points in this two dimensional space of distance and lumi-
nosity function will in general be very wasteful. We derive
a method for limiting the space in which random points are
thrown, and which also ensures that the very low probability
near and faint objects are accurately included in the number
counts.
The functions embodied in the model generate, for a
given line of sight, the PDF (the probability per unit dis-
tance that a point will be found there) and AV (r) curves
(integrated flux diminution up to that distance). Examples
are shown for a line of sight in the direction ℓ =4.0◦ and
b=0.2◦ in figure 15.
c(cid:13) 0000 RAS, MNRAS 000, 000–000
Dm = 5 lg(r/kpc) + 10
we obtain
dr
p(Dm) =
dDm
p(r) ∝ rp(r)
This function is then numerically integrated to give a
function
F (Dm) = R Dm
−∞ p(D′
Q
m)dD′
m
where the normalisation Q is given, in terms of the up-
per limit D0
m in Dm chosen for the simulation, by:
m
Q = Z D0
−∞
p(D′
m)dD′
m
Figure 15.1 shows both p(Dm) and F (Dm) for the ex-
ample shown in figure 15.
This function F (Dm) is subsequently numerically in-
verted, so that the input of a random deviate uniformly
distributed between 0 and 1 will result in an output Dm
distributed according to the PDF for the line of sight.
Figure 15.1 shows the result of this inversion, with the
ordinate ln x, where x is a uniformly distributed deviate be-
tween 0 and 1. Furthermore, corresponding to each distance
modulus, Dm, we can assign a value AV from the lookup dia-
gram shown in the upper panel of figure 15. Each waveband,
I,J and K, corresponds to a different value of extinction,
AI ,AJ ,AK . These three modified distance modulus curves
are shown in the figure.
15.2 Luminosity Function
The luminosity functions may be treated in a similar way to
the distance modulus distribution. Taking the K-band lumi-
nosity function as an example, we can, as before, generate a
cumulative distribution, invert it, and use this as a lookup
table to convert a deviate, y, uniformly distributed between
0 and 1 into a corresponding stellar type.
15.3 Combination
To limit the numbers of sources which are tried only to those
brighter than some given magnitude limit, let us denote by
x and y two independent uniform random deviates between
0 and 1. From x, we derive a modified distance modulus
according to the distance modulus and reddening combined
lookup table, as in figure 15.1. From y, we deduce a stellar
type in the same way. The overwhelming likelihood is that a
faint source at large distance is generated, which will clearly
fall out of the magnitude range of interest.
14 M.Unavane et al.
In figure 15.3, there are a multitude of lines. Concen-
trating on the long-dashed line, we see that combining the
x figure for modified distance modulus for K-band, and the
y figure for K-band luminosity function, limits in observed
K magnitude are expressed by lines roughly parallel to lines
of constant w = ln x + ln y, or constant xy. Similarly, mag-
nitude limits can be set for J and I, leading to the limiting
lines shown (dotted and dashed lines).
15.4 Generation
The aim is to take two independent uniform deviates be-
tween 0 and 1, u and v, and convert them to deviates x and
y which uniformly cover the space (x = 0 → 1, y = 0 → 1)
excluding the region with w < w0.
Graphically, we seek to fill only the left region in figure
15.4, where the bounding function is xy = ew0 .
Now for uniform deviates x and y, the product z = xy
will be distributed as
p(z)dz = Z 1
z
dxZ 1
z
x
dy −Z 1
z+dz
or
p(z) = − ln z
dxZ 1
z+dz
x
dy
To well sample the lowest probabilities, we convert this
to a function in terms of w = ln z, to give p(w) ∝ −wew.
We then integrate and normalize this function to find a cu-
mulative form, with a lower limit w0 and upper limit w1
(w0,w1 ≤ 0). The result is:
F (w) =
ew0 (1 − w0) − ew(1 − w)
ew0 (1 − w0) − ew1 (1 − w1)
As before, when this is inverted, a uniform deviate, u
between 0 and 1 can be supplied, and the value w corre-
sponding to F (w) = u gives a value for w = ln x + ln y
such that the points (x, y) are uniformly distributed in the
allowed region.
By symmetry, x and y are distributed in the same way
for any given value of xy. We can thus find values x and
y by using the second uniform deviate v. Constructing d =
w(2v − 1), we obtain a uniform deviate between −w and w.
This corresponds to the difference ln x − ln y. This leads to
the final result, that if a deviate w and a secondary deviate
d are generated as described above, then
ln x =
d + w
2
and
ln y =
d − w
2
where points (x, y) are uniformly distributed in the al-
lowed region.
The amount of the x − y plane excluded by limiting
sources to the left of the line xy = z0 is given by z0(1−ln z0),
or ew0 (1 − w0) where w0 = ln z0.
In the example illustrated, a limit of w0=−10 allows
all observable sources to be generated. This corresponds to
sampling only 1/2000th of the xy plane. In general, we find
that w0 lies between −7 and −10, allowing savings in proces-
sor time by a factor of between 140 and 2000. The algorithm
described was translated into the Super Mongo (SM) pro-
gramming language before use.
15.5 Comments about Spiral Arms
It is not clear what the contrast between arm and interarm
regions is likely to be at near IR wavelengths. The dominant
emission is from giant stars, which will plausibly have dif-
fused away from the sites of young star formation associated
with spiral arms. Furthermore, dust associated with the spi-
ral arms, which plays a role in increasing the contrast at
short wavelengths, has a lesser effect at longer wavelengths.
Hence the contrast is likely to be less beyond 1µm than
below it. Studies of external disk galaxies (de Jong, 1996)
suggest that the contrast may be <
∼ 3, which is in agree-
ment with recent DIRBE based model-dependent analysis
of our galaxy (Binney, Gerhard & Spergel 1996). It is not
even certain that well-defined spiral arms are observable at
all wavelengths. Independent observations using bright O
stars to probe the disk (Neckel & Klare 1980) and HI re-
gions (Georgelin & Georgelin 1976) yield different spiral arm
structure.
However, the limited directions to which the model is
to be applied in this case – the inner 10◦ of the galaxy –
means that lines of sight will in general pass perpendicularly
to the directions of any spiral arms which may be in the
disk. For the purposes of investigating longitude dependent
asymmetries, lines of sight perpendicular to the arms and
separated by no more than 10◦ will not have a very different
amount of spiral arm signature present.
For these reasons, and anticipating the method used to
remove near disk signature from number counts, we do not
include spiral structure in the model. For reference, a sam-
ple set of colour-magnitude diagrams is shown when spiral
arms are included. Spiral arms of gaussian width 0.5 kpc are
placed in the line of sight at distances 3.4, 5.1 and 6.9 kpc
from the Galactic centre in accordance with the model of
Georgelin & Georgelin (1976), and are given an enhance-
ment of a factor of 3 with respect to the underlying disk
density. (figure 15.5). No modification is made to the dust
model. There is only a slight hint of difference between this
model and the comparison model without spiral arms (figure
6), and it is not useful at this level, to include this subtlety
in our model.
c(cid:13) 0000 RAS, MNRAS 000, 000–000
This figure "figure3.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure8.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure9.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure10.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure11.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure15.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure16.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure17.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure18.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure22.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure23.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
This figure "figure33.gif" is available in "gif"(cid:10) format from:
http://arxiv.org/ps/astro-ph/9710317v1
|
astro-ph/0205521 | 1 | 0205 | 2002-05-30T01:35:33 | First spectroscopic evidence for carbon stars outside the Local Group: properties of a massive star cluster in NGC 7252 | [
"astro-ph"
] | We present near-IR spectroscopy of the massive intermediate age star cluster W3 in the merger remnant galaxy NGC 7252, obtained with the NTT telescope. This cluster has an age when the integrated near-IR properties of a stellar population are dominated by the cool and luminous AGB stars. We compare the data with instantaneous curst model predictions from new evolutionary synthesis models that include: (1) the computation of the evolution through the TP-AGB for low- and intermediate-massive stars, with the initial mass and metallicity dependent formation of carbon stars; (2) spectroscopic data from a new stellar library in which differences betwenn static giants, vriable O-rich TPAGB stars and carbon stars are accounted for. The comparison of the data to the models clearly shows that carbon stars are present: for the first time, carbon star spectral features are thus detected directly outside the Local Group (abriged) | astro-ph | astro-ph | Abstract. We present near-infrared [1 − 2.3 µm] spec-
troscopy of the massive intermediate age star cluster W3
in the merger remnant and proto-elliptical galaxy NGC
7252, obtained with the NTT telescope. This cluster has
an age when the integrated near-infrared properties of a
stellar population are dominated by the cool and luminous
Asymptotic Giant Branch (AGB).
We compare the data with instantaneous burst model
predictions from new evolutionary synthesis models that
include: (i) the computation of the evolution through
the thermally pulsing AGB (TP-AGB) for low- and
intermediate-massive stars, with the initial mass and
metallicity dependent formation of carbon stars; (ii) spec-
troscopic data from a new stellar library in which differ-
ences between static red giants, variable oxygen rich TP-
AGB stars and carbon stars are accounted for. The new
evolutionary model predicts that the contribution of car-
bon rich stars to the luminosities in the near-IR passbands
is a strong function of metallicity.
The comparison of the data to the models clearly shows
that carbon stars are present: for the first time, carbon
rich star spectral features are thus detected directly out-
side the Local Group galaxies. Good fits to the available
optical/near-IR photometry and the near-IR spectrum of
NGC 7252-W3 are found for an age of 300-400 Myr and
AV ≃ 0.6 − 0.8. The models show that these parameters
depend weakly on the model metallicity in the range of
Z/Z⊙ = 0.4 − 1, with higher likelihood for solar metallic-
ity models.
At solar metallicity, a mixture of carbon rich and oxygen
rich stars is predicted. The strength of the near-IR molec-
ular bands that originated from oxygen rich AGB stars
can be used to constrain the absolute Tef f scale of these
objects, i.e. a relation between colour and Tef f . We found
that, in the framework of our set of evolutionary tracks,
the data are more consistent with the temperature scale
calibrated on Long Period Variables than on giant stars.
At a given colour, variable AGB stars have a lower Tef f
than static (or quasi-static) M giants.
Key words: stars: AGB -- galaxies: star clusters -- galax-
ies: stellar content -- infrared: galaxies
2
0
0
2
y
a
M
0
3
1
v
1
2
5
5
0
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
missing; you have not inserted them
ASTRONOMY
AND
ASTROPHYSICS
First Spectroscopic Evidence for Carbon Stars Outside the
Local Group: Properties of a Massive Star Cluster in NGC
7252⋆.
M. Mouhcine1,⋆⋆, A. Lan¸con1, C. Leitherer2, D. Silva3, M.A.T. Groenewegen3,4
1 Observatoire Astronomique de Strasbourg, 11 rue de l'Universit´e, F-67000 Strasbourg, France
2 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
3 European Southern Observatory, Karl-Schwarzschild-Str.2, D-85748 Garching, Germany
4 Instituut voor Sterrenkunde, PACS-ICC, Heverlee, Belgium
Received ; accepted ..
1. Introduction
For the past decades, great efforts have been devoted to
the determination of the star formation history of com-
plex stellar systems, such as galaxies, with the ultimate
aim of understanding the history of star formation on cos-
mological timescales. In such composite systems, we look
at a mixture of stellar populations with large spreads in
age and metallicity (Moller et al. 1997). The picture is
further blurred by the presence of dust. An extended ob-
servational wavelength coverage is a major advantage in
attempts to break the degeneracies between the effects of
age, metallicity and extinction on galaxy light. In this con-
text, near-infrared (near-IR) observations are justified by
the relatively low sensitivity of near-IR light to dust ex-
tinction (stellar subpopulations that are heavily obscured
contribute more to the near-IR light than they do at
shorter wavelengths), and because the near-IR light orig-
inates specifically in subpopulations of old or metal-rich
stars (Frogel et al. 1978, Rieke & Lebofsky 1979, Frogel
1985, Silva 1996, Lan¸con et al. 1996, Goldader et al. 1997,
Fritze-von Alvensleben 1999). In practice, the interpreta-
tion of the near-IR stellar energy distribution frequently
remains limited to the global confirmation of the presence
of "old" stars or, in starburst environments, to the sug-
gestion of a ∼ 107 yr young population of red supergiants
(Oliva et al. 1995, Lan¸con & Rocca-Volmerange 1996). In
particular, the specification of the actual age of the "old"
stars seems to have stayed out of reach. This is mainly due
to large uncertainties in the physics of cool stars, used as
input to the population synthesis calculations (Charlot et
al. 1996, Girardi 1996, Origlia et al. 1999, Lan¸con et al.
1999).
⋆ Based on observations obtained at the NTT 3.5m of the
European Southern Observatory, Chile
⋆⋆ present address: Departement of Physics & Astronomy,
UCLA, Math-Sciences Building 8979, Los Angeles, CA 90095-
1562
Important progress in extragalactic near-IR astronomy
would be achieved if contributions from intermediate age
(<2 Gyr, but > 100 Myr) and old (>2 Gyr) populations
could be safely separated in the integrated light of galax-
ies. At intermediate ages, asymptotic giant branch stars
(AGB stars) are responsible for most of the K-band flux.
Integrated broad band colours are determined by these
stars (Frogel et al. 1990, Bruzual & Charlot 1993, Girardi
& Bertelli 1998, Mouhcine & Lan¸con 2002a). But broad
band colours remain ambiguous, as they are similarly af-
fected by extinction or by an additional population of red
supergiants, AGB stars or metal-rich giants. Lan¸con et al.
(1999; see also the early review of Lan¸con 1999) have sug-
gested the use of the broad spectral signatures of upper
AGB stars for an age separation. The basic idea is that
most upper AGB stars are Long Period Variables (LPVs),
and that this variability produces more extended atmo-
spheres, thus leading to significantly stronger molecular
bands. When oxygen-rich AGB stars become carbon-rich,
they display specific molecular signatures that can again
be identified. For the first time, stellar spectral libraries
that included oxygen-rich LPVs and carbon stars had been
used as input to an evolutionary spectral synthesis code.
However, the qualitative predictions are strongly depen-
dent on as yet only partly constrained AGB evolution pa-
rameters (Lan¸con et al. 1999, Mouhcine & Lan¸con 2002).
For a given set of evolutionary tracks and a given library
of stellar spectra, a very influential parameter is the tem-
perature scale used to relate the two.
To improve the reliability of the synthesis models, we
need to check the accuracy of their physical ingredients
in simple cases. Star clusters provide an invaluable op-
portunity to do so. Because of their proximity and of the
large range of properties they display, Magellanic Cloud
(LMC, SMC) clusters were used extensively. However,
LMC/SMC clusters contain only a handful of luminous
red stars. Their integrated properties are strongly affected
by the stochastic character of this small subpopulation
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
3
(Santos & Frogel 1997, Ferraro et al. 1995). Lan¸con &
Mouhcine (2000) have shown that coeval populations with
more than 105 M⊙ of stars are needed if one wishes random
deviations from the mean near-IR properties predicted by
population synthesis models to remain smaller than the
effects of the model parameters of astrophysical interest.
Repeated observations of 105 M⊙ clusters or even larger
masses are required in order to cancel effects of the vari-
ability of the brightest stars as well. Such massive clusters
are rare in nearby resolved galaxies (Grebel 2000).
Over the last few years, the Hubble Space Telescope
has revealed the presence of numerous young, compact and
very bright knots in merging galaxies, e.g. NGC 4038/4039
(Whitmore & Schweizer 1995, Whitmore et al 1999), NGC
3597 (Holtzman et al. 1996), NGC 7252 (Miller et al 1997)
and NGC 1275 (Carlson et al 1998). Their colours, lumi-
nosities, sizes and spatial extent agree with them being
proto-globular clusters. These results suggest that galaxy
interaction and mergers are favorable environments for
the formation of globular clusters. The "Atoms for Peace
galaxy" NGC 7252 (Arp 226) is one of the most exten-
sively studied merging remnants. Star clusters of excep-
tional luminosity were discovered in NGC 7252 (Schweizer
1982, Whitmore et al. 1993). The analysis of UBV and
I photometry leads to estimated ages of several 100 Myr
and extraordinary masses of ∼ 107 M⊙ for the most lu-
minous ones (Miller et al 1997). The NGC 7252 massive
clusters offer us the possibility to calibrate the stellar in-
put of our near-IR modelling of intermediate age SSPs
without worrying about stochastic fluctuations.
In Section 2 we describe the observations and the data
reductions. Section 3 presents an overview of our new
modelling of near-IR spectra of SSPs. In Section 4, we
highlight the evidence for a significant population of car-
bon stars in he near-IR spectrum of NGC 7252 -- W3 (here-
after W3). In Section 5, we combine UBVIKs photometry
and the near-IR spectrum to derive new constraints on the
age of the cluster and the extinction on the line of sight.
Constraints on the cluster metallicity and on the temper-
ature scale of AGB spectra are discussed in Section 6.
A comparison with previous models, in which the evolu-
tionary tracks for the TP-AGB did not account for "hot
bottom burning", is given in Section 7. Finally, Section 8
summarizes the main conclusions.
2. Cluster selection, observations and data
reductions
2.1. Selection
The candidate clusters were selected from the lists of
Whitmore et al. (1993) and Miller et al. (1997). Only clus-
ters with estimated ages between 0.1 and 1 Gyr were con-
sidered good candidates for our purpose. At these ages,
no red supergiant stars are present any more, which elim-
inates one potential source of contamination. Metallici-
ties between Z=0.3 Z⊙ and Z=Z⊙ (based on the chem-
ical evolution models of Fritze-von Alvensleben & Ger-
hard 1994) in fact suggest a much narrower age range of
600 to 800 Myr for these clusters. Schweitzer & Seitzer
(1998) re-analysed the photometry and combined it with
optical spectroscopy. They favoured solar abundances and
slightly younger ages (300 -- 600 Myr for cluster W3). Fi-
nally, Maraston et al. (2001) added K band photometry
to the data and used updated models to confirm a metal-
licity range of 0.5 to 1 Z⊙, and to restrict the range of ages
to 300 -- 500 Myr (for cluster W3, they favoured Z=0.5 Z⊙
and an age of 250 -- 300 Myr).
The observable intermediate age clusters are intrinsi-
cally bright, with MV <∼ −14 and MK <∼ −16.5. Assuming
standard stellar initial mass functions (Salpeter 1955, or
Scalo 1986), this corresponds to stellar masses of the order
of 107 M⊙ or more: all these clusters are massive enough to
host a representative population of luminous AGB stars.
2.2. Imaging
J and Ks (K short; λ/∆λ = 2.162/0.275 µm) images of
NGC 7252 were taken with the IR imager-spectrometer
SOFI on the European Southern Observatory New Tech-
nology Telescope (La Silla, Chile) between August 14 and
August 19, 1999. Two of the five half nights allocated to
the program were adequate, and were used to acquire the
imaging and the spectroscopic data of this paper.
′
′
× 5
′′
SOFI has a field-of-view of ∼ 5
and a pixel size of
0.292
. All images were acquired by integrating repeatedly
for equal amounts of time on the galaxy and on nearby
sky fields. Both galaxy and sky frames were dithered with
offsets ≥ 10 ′′, so that a median of the frames would effec-
tively remove field stars, the detector defects and cosmic
rays. This technique also has the advantage of minimiz-
ing residual flat-fielding errors. The individual object in-
tegration times were 200 s in Ks (20 × 10 s) and 80 s in J
(20 × 4 s). The object/background integration cycle was
repeated three times for the Ks band and 8 times for the
J band. The resulting total on-object exposure times are
1000 s in Ks and 560 s in J. Standard stars from Persson
et al. (1998) were used for the photometric calibration.
′
The "Atoms for Peace" galaxy, NGC 7252, is located at
α = 22h20m44s.8, δ = −24◦40
42" (J2000) (Miller et
al 1997) with a recession velocity relative to the local
group of 4828 km s−1 (Schweizer 1982). We adopt a dis-
tance modulus (m − M )o = 34.08 and a Milky Way fore-
ground extinction of AV = 0.04 (de Vaucouleurs et al
1990; Ho = 75 km s−1 Mpc−1).
The images were processed with IRAF1. A dark frame
was first subtracted from every object and sky image of a
given data set. The dark frame was constructed from a me-
1 The Image Reduction and Analysis Facility by the Na-
tional Optical Astronomy Observatories which are operated by
AURA, Inc., under cooperative agreement with the National
Science Foundation, USA.
4
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
stellar populations of various ages in the near-IR are ex-
tremely broad (Lan¸con et al. 1999, Mouhcine & Lan¸con
2000). They reach over the telluric absorption bands that
mark the edges of the J, H and K atmospheric windows.
The vast majority of near-IR spectrometers only allow the
observation of one window at a time. Varying slit losses
from one observation to the next then make it extremely
difficult to recover the relative flux levels of the spectral
segments and to measure and use the spectral features.
For the first spectroscopic observations (blue setting),
the 1′′ slit was placed across two clusters, W3 and W30
(position angle 117.4). This slit position also covers the
galaxy emission close to the nucleus. As weather made it
clear that not enough signal would be collected for W30,
the subsequent observations were made at position angle
104.6, with the slit running through W3 and the galaxy
nucleus. The SOFI slit is 290′′ long. Nodding along the slit
for first order background subtraction was thus possible.
Our observations consist of series of 40 minute nodding
sequences, amounting to 9600 s exposure in the blue set-
ting and 10800 s exposure in the red settings. Standard
star spectra were taken before and after each sequence on
NGC 7252.
The spectroscopic data for cluster W3 were reduced in
the following steps. Pairs of frames obtained at two po-
sitions along the slit were combined by subtraction and
flat-fielded, using the adequate lamp-on -- lamp-off frame.
The resulting frames carry a positive and a negative ver-
sion of the spectrum. The numerous cosmic rays could only
partially be removed with automatic procedures; careful
eye inspection was used to remove cosmic rays very close
to the cluster and galaxy light. Then the spectral images
were calibrated in wavelength using a xenon lamp spec-
trum as a reference. To extract a one dimensional spec-
trum for the cluster, the cluster emission was traced along
the two-dimensional images. The adopted extraction aper-
ture width varied between 5 and 6 pixels, depending on
the seeing. The combined galaxy background and residual
sky background at each wavelength was estimated via a
cubic spline fit to background windows on either side of
the trace center. Following the background subtraction,
each individual spectrum was flux calibrated using an av-
erage of standard star spectra taken before and after the
scientific exposure. The division by the standard star spec-
trum removed telluric absorption features. Kurucz spectra
(1993) were used as models for the intrinsic energy distri-
bution of the standards, and thus to recover the energy
distribution of the cluster. Individually calibrated cluster
spectra from different images were then combined into one
final spectrum. The square of the signal-to-noise ratio was
used to weight the individual spectra. Finally, the blue and
red spectra were merged.
Fig. 1. Masked version of NGC 7252 J band image. Clus-
ters are marked following Miller et al 1997.
dian of dark exposures taken at the beginning or at the end
of the night. The dark currents were very stable. Flatfields
in the two bands were made by median combining the
standard star images and subtracting the corresponding
dark frame. These sky flatfields were found to be superior
to the dome flatfields. The flatfield image was divided into
each dark subtracted object and sky image. A reference
background image for each band was then constructed by
median-combining the sky images. This image was sub-
tracted from all galaxy images of a given series of obser-
vations. The individual sky-subtracted galaxy images were
then carefully shifted and median combined using a sigma-
clipping algorithm to form a single final image of NGC
7252. We modelled the galaxy with a 31 × 31 pixels me-
dian filter and subtracted the light profile of the galaxy, in
order to obtain a flat, low-noise background. The masked
J-band image of the central region of the galaxy is shown
in Fig. 1, where the star clusters are marked using the
identification numbers of Whitmore et al. (1993).
2.3. Spectroscopy
The low resolution grisms of SOFI make it possible to
cover the spectral range between 0.95 and 2.52 µm in only
two settings. The blue setting ranges from 0.95 and 1.64
µm , the red setting from 1.53 to 2.52 µm. The extended
wavelength coverage of each setting and the overlap be-
tween the two are decisive advantages of SOFI over other
available medium resolution infrared spectrometers on 8-
10m class telescopes. Indeed, the features that distinguish
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
5
3. Models for the near-IR emission of stellar
populations
Studies of the evolution of the near-IR spectral energy
distribution of intermediate age stellar populations need
to take into account stars of the Early and of the Ther-
mally Pulsing AGB, since both contribute significantly to
the near-IR light (Frogel et al. 1990, Charlot & Bruzual
1991, Lan¸con 1999). The sets of evolutionary tracks or
isochrones made available by stellar theory groups for use
in population synthesis models do not in general extend
beyond the Early AGB. An extension through the TP-
AGB is required. Unfortunately the mechanisms driving
the TP-AGB evolution are complex and still poorly un-
derstood (e.g. Iben & Renzini 1983, Habing 1996, Olof-
sson 1999). Uncertainties in the physics of mass loss, of
the third dredge-up process, and of hydrogen burning at
the bottom of the convective envelope, together with the
lack of reliable models for the molecular spectral features
of the coolest stars, make the inclusion of TP-AGB stars
a difficult task. The integrated properties of the stellar
populations in the near-IR will depend on the modelling
of the TP-AGB phase (see Girardi & Bertelli 1998 for a
clear illustration).
The analysis of the clusters of NGC 7252 is based on
the new spectral synthesis models presented by Mouhcine
& Lan¸con (2002a, hereafter ML2002), in which the TP-
AGB contribution to the integrated light is taken into
account. The synthetic stellar evolution code for the
TP-AGB incorporates the above-mentioned physical pro-
cesses, through parameters that have been adjusted to
reproduce observational constraints from the Magellanic
Clouds and the solar neighbourhood. It predicts the evo-
lution of stars of various masses in the HR diagram, as
well as the evolution of the surface chemistry and of the
stellar mass (thus the formation of carbon stars and of
dust-obscured mid-infrared sources). Moreover, it uses a
stellar library that includes an appropriate optical + near-
IR spectrum for each of these evolutionary stages. The
reader is referred to ML2002 for a detailed description.
Let us however summarize the main properties here.
Up to the E-AGB, the stars are assumed to follow the
evolutionary tracks of the Padova group (e.g. Bressan et
al. 1993 for [Z=0.008, Y=0.25], Fagotto et al. 1994 for
[Z=0.02, Y=0.28]). The extensions through the TP-AGB
include known physical processes that affect the evolu-
tion and play a dominant role in determining TP-AGB
lifetimes, the extent of nuclear processing and the chemi-
cal surface abundances: the third dredge-up, the envelope
burning, the mass loss and its final superwind phase. From
the end of the E-AGB on, the total mass, core mass, ef-
fective temperature, bolometric luminosity and carbon to
oxygen ratio in the envelope are evolved according to semi-
analytical prescriptions (see Wagenhuber & Groenewegen
1998). The basic relations of the model are the following:
1. a detailed core mass-luminosity relation;
2. the core mass-interpulse relation;
3. the rate of evolution of the core mass;
4. an algorithm to evaluate the effective temperature;
5. a test of the occurrence of CNO burning at the base
of the convective envelope;
6. a prescription to evaluate the mass loss rate as function
of the stellar parameters;
7. the third dredge-up and its efficiency;
8. an assumed composition of the third dredge-up mate-
rial.
The calculations used here are performed using a mixing
length parameter α = 2, and the mass loss prescription of
Blocker & Schonberner (1993) with a mass loss efficiency
of η = 0.1 (Groenewegen & de Jong 1994). The end of
TP-AGB phase is determined either by the total ejection
of the stellar envelope, or by the core mass growing up to
the Chandrasekhar mass limit.
The properties of the TP-AGB stars are sensitive to
the initial metal content. This in particular affects the
formation of carbon stars. The LM2001 models are able
to reproduce the observed relation between the metallicity
and the relative number of carbon stars (Pritchet et al.
1987), in the sense that lower initial metallicities result in
higher proportions of carbon stars among TP-AGB stars,
carbon stars being the dominant spectral type for [Fe/H]
<∼ -1 (Blanco et al. 1978, Cook et al. 1986). This behaviour
reflects the fact that the fraction of the total TP-AGB
lifetime a model star spends as a carbon rich object is a
function of metallicity. Stars at Z=0.05 with initial masses
of 2-3 M⊙ (i.e. the privileged mass range for the formation
of carbon stars; Mouhcine & Lan¸con 2001 b) spend <∼ 10 %
of their total TP-AGB lifetime as carbon stars, while the
fraction reaches ∼ 80 % for the metallicity of the LMC or
lower.
LW2001 use a code derived from P´egase (Fioc &
Rocca-Volmerange 1997) for the synthesis of integrated
spectra. The spectra of Lan¸con & Wood (2000), averaged
as described by Lan¸con & Mouhcine (2001), are used for
luminous stars with spectral types later than K5, and the
colour-corrected spectral library of Lejeune et al. (1998)
is used elsewhere along the stellar tracks. Cool and lumi-
nous TP-AGB stars are also variable stars. This variability
strongly affects the near-IR energy distribution of those
stars. They show deep water absorption bands when they
are oxygen rich, and strong C2 and CN absorption bands
when they are carbon rich. Hence the near-IR spectra of
intermediate age stellar populations will change as the pre-
dominant stars change as function of age and metallicity.
The adopted relation between the effective temperature of
TP-AGB stars and their spectrum is a second parameter
that affects the shape of the near-IR SSP spectra.
The evolutionary sequences computed for this paper
assume that stars formed in an instantaneous starburst
(star-formation timescale of 1 Myr or less), with a Salpeter
6
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
Fig. 2. The effect of carbon stars on the predicted spectrum of an intermediate age cluster (Z=0.008, age = 500 Myr).
The flux density is given in arbitrary Fλ units, but both spectra correspond to the same total stellar mass. Thick line:
standard model; thin line: result obtained with O-rich stellar spectra only.
(1955) power-law initial mass function extending from 0.1
to 120 M⊙.
To illustrate the effect of carbon stars on the inte-
grated cluster light, we have computed additional model
sequences in which exclusively the spectra of O-rich stars
were used, regardless of the actual chemical composition
given by the evolutionary tracks. The differences between
the two sequences are most clearly visualized at subsolar
metallicity (because carbon stars are more important) and
at ages between 0.5 and 1 Gyr (because the near-IR con-
tribution of the TP-AGB stars are largest at these times,
in the ML2002 models). An example is provided in Fig. 2.
The figure shows that the only broad band flux mea-
surement sensitive to the use of C-rich spectra is the J
band, due to CN and C2 opacities (see Loidl et al. 2001
for band identifications in carbon star spectra, and Lan¸con
& Wood 2000 for the O-rich features). The effect on the
integrated H and K fluxes is negligible. Using low resolu-
tion spectroscopy implicitely accounts for the JHK energy
distribution, as long as this whole spectral range is cov-
ered continuously. The main discriminating spectral fea-
tures can be measured at low spectral resolution (δλ/λ of
a few 100) or with high signal-to-noise narrow band pho-
tometry (Lan¸con et al. 1999). When carbon star spectra
are included in the models, the VO band at 1.05 µm van-
ishes in favour of the CN bandhead at 1.09 µm; the H2O
bandhead at 1.33 µm is replaced with a CN bandhead at
1.35 µm; the absence of the wings of the H2O bands on
both sides of the H window give the H spectrum a flat ap-
pearance, with an abrupt cut-off at 1.77 µm due to C2; and
for the same reason the K band spectrum is also straight.
In the rest of this paper, our aim is to answer the
following questions:
• Can models for the near-IR spectrophotometric evo-
lution of star clusters reproduce the spectrum of W3, with
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
7
an age, metallicity, extinction and mass consistent with
constraints from optical spectra ?
• What constraints are obtained on the chemical na-
ture of the AGB stars present in W3 ?
• Conversely, what constraints does the AGB popula-
tion of W3 set on the models for the formation of carbon
stars ?
• What temperature scales for AGB star spectra are
consistent with the data ?
• How much of this information could have been de-
termined from the near-IR SOFI spectrum only, i.e. in the
absence of an priori knowledge of the age from an optical
spectrum ?.
4. The Near-IR Spectrum of Cluster W3: the
Detection of Carbon Stars
It is well known from counts in the Magellanic Cloud clus-
ters that carbon stars represent a large fraction of the
AGB stars of intermediate age clusters, at least at sub-
solar metallicity (e.g. Aaronson & Mould 1980, Persson et
al. 1983, Frogel et al. 1990, Westerlund et al. 1993, Ferraro
et al. 1995). Population synthesis models predict the pre-
dominance of carbon stars over M type stars in the near-
IR, under favourable age and metallicity circumstances. In
agreement with star counts, they predict that carbon stars
may provide up to 50 % of the near-IR light (LM2001).
Nevertheless, carbon stars have never been searched for
in galaxies with unresolved stellar content. By including
spectra of carbon stars in evolutionary spectral synthesis
models, Lan¸con et al. (1999) opened a new window on
distant carbon star populations. In this Section, we use
NGC 7252 -- W3 to validate that approach.
Are carbon rich stars present in cluster W3, as ex-
pected from the intermediate age cluster models? In Fig. 3,
the SOFI spectrum of W3 is compared with models for a
300 Myr old stellar population at solar metallicity. In the
right frame, carbon star spectra are used when the sur-
face abundance of carbon exceeds that of oxygen; in the
left frame, oxygen rich spectra are used exclusively. The
choices of age, metallicity and extinction are justified in
Section 5 and 6.
At 300 Myr, the contribution of the TP-AGB stars to
the near-IR light is smaller than at the 500 Myr of Fig. 2.
Thus the effect of carbon stars is also less pronounced.
In addition, some key molecular bandhead could not be
recovered because of the strong and variable telluric ab-
sorption at La Silla. Nevertheless, in all the remaining
spectral regions of interest (identified in Fig. 2) the agree-
ment is systematically better when carbon star spectra are
included. The strongest evidence for carbon stars comes
from the absence of the VO band at 1.05 µm, the presence
of the CN bandhead at 1.08 µm, and the small curvature
of the energy distribution inside the H window. For each
of these, the deviation between the smoothed data and
the exclusively O-rich models are at the 3 σ level. Taken
together, the signatures provide strong evidence for the
presence of carbon stars. The shape of the K window en-
ergy distribution also favours this conclusion, but here the
difference between models is not statistically significant in
view of the observational noise.
The detection of carbon stars in W3 demonstrates that
TP-AGB studies can be extended to distant, unresolved
stellar populations. Clearly, better signal-to-noise ratios
should be seeked. Broadband colours are not sufficient
for studies of the nature of the TP-AGB stars, in par-
ticular when the amount of reddening is uncertain. The
shape of the low resolution near-IR spectrum between 1
and 2.5 µm carries most of the relevant information. It is
essential to recover this global shape. Spectrographs with
a broad wavelength coverage and a wide overlap between
individual spectral segments are the most appropriate in-
struments. Multiple-filter narrow-band photometry, for in-
stance with the future refurbished camera NICMOS on
board HST, is an alternative approach if the absolute flux
levels can be calibrated with a high accuracy (Lan¸con et
al. 1999).
The effects of carbon stars on the spectrum of W3 also
make it clear that near-IR evolutionary synthesis models
must incorporate the spectral contribution of these objects
if they are to be used in the analysis of future high signal-
to-noise near-IR data.
5. Age and extinction from the near-IR spectrum
and the UBVIKs colours
Figure 3 not only highlights the role of carbon stars in
W3. It shows that the new spectral synthesis models of
Mouhcine & Lan¸con (2002a) are indeed able to reproduce
the integrated spectrum of the cluster, for an intermedi-
ate age and at a metallicity consistent with the estimates
based on optical spectra (Schweizer & Seitzer 1998, Maras-
ton et al. 2001).
In this and the following sSctions, we will examine
what information on age, extinction and metallicity can
be recovered in the absence of optical spectroscopy, using
only the near-IR spectrum, only the UBVIKs photometry,
or these two sets of data combined.
Here, we first focus on age and extinction. As the red-
dening of NGC 7252 and, more specifically, W3 is poorly
known, we consider visual extinction as a parameter that
needs to be better constrained with the new data.
The analysis of the optical spectra of the brightest
star clusters of NGC 7252 by Schweizer & Seitzer (1998)
and Maraston et al. (2001) has yielded metallicity esti-
mates. Schweizer & Seitzer (1998) favoured solar metal-
licity. Maraston et al. prefer the metallicity of the LMC
for W3, but solar abundances for other luminous clusters
with similar ages. Although it is not strictly impossible
that coeval clusters formed from material with different
abundances in a merger, we estimate that the apparent
8
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
Fig. 3. Evidence for carbon stars in the spectrum of W3. Right panel shows comparison between the observed W3
spectrum (dotted-line) and the synthesis model (continuous line )where carbon stars were allowed to form, while the
left panel shows a comparison to a synthesis model where all TP-AGB stars were forced to remain oxygen rich during
the whole phase (see Section 4). The data has been smoothed to a resolution of ∼ 50 A.
spread in metallicity is more likely to represent uncertain-
ties. Therefore, this paper will consider both solar and
LMC metallicities. Here, we mainly work at Z⊙, and we
come back to this choice in Section 6.
5.1. The broadband UBVIKs colours
To assess the relative quality of various model adjustments
to the data, we have computed goodness-of-fit maps.
Each map displays the variations of the reduced, weighted
quadratic difference between model and data (noted ∆2
hereafter), for model ages ranging from 107 to 1.5× 1010 yr
and for model extinctions ranging from AV = 0 to AV = 2.
Figure 4 shows the typical aspect of such maps, based on
the broad band colours (U-B), (B-V), (V-I) and (I-Ks)
(Miller et al. 1997, Maraston et al. 2001).
The figure displays a rather regular pattern. A strong
degeneracy between age and extinction is apparent: as
time passes, the models become intrinsically redder, thus
requiring less extinction. The shape of blue young cluster
SEDs cannot be made to match the observed UBVIKs dis-
tribution with any amount of obscuration. At solar metal-
licity, old model clusters are intrinsically too red and must
also be excluded. In agreement with previous studies, only
intermediate ages are found to be acceptable.
A quantitative interpretation of the maps requires a
closer look at the way ∆2 values are computed, and at the
10+10
5
2
10+9
5
2
10+8
5
2
10+7
0.0
0.54
10.56
1.5
10.55
0.5
1.0
1.5
2.0
Fig. 4. Goodness-of-fit map obtained in the age-extinction
plane with only the broadband (U-B), (B-V), (V-I) and (I-
Ks) colours. Plotted are AV (magnitude) and age (yr) on
the X and Y-axis respectively. The models assume Z=0.02.
The extinction law of Cardelli et al. (1989) with R=3.1 is
used. The contour delineates the region of good fits (see
text). The minimum ∆2 value and a few other illustrative
values are indicated.
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
9
sources of uncertainties. We have:
∆2 =
1
4
4
X
i=1
(OCi − M Ci)2
σ2
i
where OCi is an observed colour, M Ci is the correspond-
ing model colour and σi is the estimated r.m.s. error on
colour i. σi combines uncertainties in the data and in the
models. For the observational uncertainties, we adopted
σ = 0.06 mag for (U-B) and 0.04 mag for the three other
colours, as suggested by Miller et al. (1997) and Maraston
et al. (2001).
Miller et al. (1997) have used the transformations of
Holtzman et al. (1995) to transform HST/WFPC2 colours
into the "standard" ground-based UBVRI system (of Lan-
dolt 1983). The UBVI filter passbands of Bessell (1990)
that we use here are closely matched to that standard
system, suggesting passband mismatch errors smaller than
0.02 mag. When measuring colours on a model spectrum,
we adopt zero colours for a model of Vega. We have
verified that the systematic errors due to this conven-
tion (Holtzman et al. 1995) and to the slight differences
around the Balmer lines between our Vega model and
the one used in standard HST data reductions (IRAF
stsdas.hst_calib.synphot package) are <∼ 0.015 mag.
The Ks measurements of Maraston et al. (2001) are tied
to the photometric system of Persson et al. (1998), who
defined the Ks zero point by imposing Ks = K for the
type A0 standards of Elias et al. (1982). Systematic er-
rors on (I-Ks) are likely to be at least of the same order as
those on optical colours (but are more difficult to estimate
quantitatively, as empirical photometric systems focus on
either the near-IR or the optical, not the two combined).
Using 9 tabulations of extinction laws for the Milky Way
or the Magellanic Clouds from the literature, we estimate
r.m.s. dispersions of 0.005 to 0.025 magnitudes at AV =1,
depending on the colour considered (or more if we include
extragalactic extinction laws such as given by Calzetti et al
1994, although it is not known if they apply to individual
lines of sight in a galaxy). Errors due to stellar evolution
prescriptions and spectral libraries are difficult to assess
precisely. Based on comparisons between the colours pre-
dicted by various research groups (e.g. Charlot 1996, Yi
et al. 2001, Lejeune & Buser 1999, Salasnich et al. 2000),
one must account for ∼ 0.03 mag r.m.s. errors due to the
uncertainties in population synthesis model inputs. The
effects of choices in the modelling of the TP-AGB are not
counted as errors since they represent the parameters our
study must test (the TP-AGB parameters do not affect
(U-B) and (B-V) significantly). In view of all the factors
of this paragraph, we have accounted for the uncertain-
ties in the models globally with a factor of 1/2 on ∆2
(grossly corresponding to what is expected if the uncer-
tainties on each model colour equals the uncertainty on
the observed colour) and an extinction dependent correc-
tion (corresponding to 0.02 mag at AV =1).
Fig. 5. Comparison between the 4 observed colours (rep-
resented as 2 σ error bars connected with a solid line) and
model colours (open circles). Top: Effect of AV on the
broad band colours, at an age of 300 Myr. Bottom: Ef-
fect of age on the broad band colours, for AV = 0.6. The
extinction law of Cardelli et al. (1989) with R=Av/E(B-
V)=3.1 is used. The circle sizes are roughly representative
of 2 σ errors on the models (see text).
The interpretation of the resulting values in terms of
agreement probabilities between the data and a model, us-
ing Pearson's reduced χ2 probability distribution, would
imply the crude assumptions that the errors have a normal
distribution and are independent of one another. In partic-
ular, the uncertainties in the model colours are clearly not
independent (any stellar component contributes to more
than one colour, and extinction affects all colours coher-
ently). Here, we consider that a model provides a good fit
to the data if ∆2 < 2 on the scale of our maps. If our
χ2 estimator indeed followed Pearson's probability distri-
bution with 4 degrees of freedom, the uncertainties would
have a 15 % probability of producing larger values. Thus
our limit is rather conservative. Exemples of comparisons
between observations and models are given in Fig. 5.
5.2. The near-IR spectrum alone
Figure 6 shows the goodness-of-fit map considering only
the near-IR spectrum as constraints.
We have used the r.m.s. deviation σλ around a
smoothed continuum as an estimate of the uncertainties in
the data. The squared difference between the model and
the data is weighted with 1/σ2
λ and the sum of these dif-
ferences is minimized by applying the adequate scaling to
the model. ∆2 is obtained by dividing the resulting sum
by the number of degrees of freedom, (N − 1), where N
is the number of pixels used along the spectrum. We note
that we do not correct for the size of the spectral PSF,
10
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
which introduces correlations between errors on 2-3 ad-
jacent pixels. Observational uncertainties on the slope of
the spectrum affect estimates of AV . A 5 % uncertainty
on the ratio of the fluxes at 1 µm and at 2 µm translates
into an uncertainty of 0.15 magnitudes in AV .
10+10
5
2
10+9
5
2
10+8
5
2
10+7
0.0
1.89
3.05
1.37
1.46
3.05
1.46
1.37
1.46
1.26
0.5
1.0
1.5
2.0
Fig. 6. Goodness-of-fit map obtained in the age-extinction
plane with the SOFI spectrum of W3 only (axes as in
Fig. 4). The models assume Z=0.02. The extinction law
of Cardelli, Clayton & Mathis (1989) with R=3.1 is used.
The minimum ∆2 value and some other illustrative values
are given. The contour indicates regions of acceptable fits.
Again, the numerical values of ∆2 cannot be directly
transformed into acceptance probabilities. The various
sources of uncertainties that we have cited and the eye
inspection of a variety of model adjustments leads us to
set the limit of acceptable ∆2 values at 1.5. The contour
in Fig. 6 identifies the corresponding region of acceptance.
The model adjustment shown in Fig. 3 (b) corresponds to
[t=300; AV =0.7]. It provides a good fit to the data with
a ∆2 of 1.39 on Fig. 6.
A curved line of relatively low ∆2 values runs through
the map, reflecting the degeneracy between extinction and
stellar evolution in producing a red enough broad band
energy distributions between 1 and 2.5 µm. Three main
segments can be identified along this curve. (i) At ages
between 107 and 108 yr, red supergiants are the main
sources of near-IR light. At about 20 Myr, they provide
intrinsic near-IR colours that are similar to those of W3
even without much extinction. Later on, stars on extended
blue loops of the evolutionary tracks make the integrated
colours bluer, shifting the ∆2-valley towards higher ex-
tinctions. (ii) In the intermediate age regime, the intrinsic
near-IR colours become redder with time, as the relative
contribution of cool AGB stars increases. The need for
extinction is progressively reduced. The poor fit between
500 Myr and ∼ 1 Gyr is due to the strength of molecular
features in the model spectra at these ages. In the observed
spectrum, these features are present, but with a relatively
small contrast. (iii) At old ages (> 109 yr), the near-IR
model spectra are intrinsically as red as or redder than
the data and any extinction would worsen the agreement.
The minimum ∆2 values are obtained at relatively
young intermediate ages (150-200 Myr) with relatively
high values of the extinction (AV = 1 − 1.4, i.e. values
consistent with those obtained for the H II region S101 of
NGC 7252 by Schweizer & Seitzer 1998). The exact loca-
tion of the minimum depends on the wavelength depen-
dence of the weighting adopted in the ∆2 computation
and on the extinction law. For instance, varying the ra-
tio R=AV /E(B-V ) from 2.7 to 5 (as observed along var-
ious lines of sight in the Milky Way; Cardelli et al. 1989)
shifts the minimum ∆2 from [t=200; AV =1.4] to [t=150;
AV =1.0], without modifying the actual value of the min-
imum ∆2 significantly.
5.3. The broadband colours and the near-IR spectrum
combined
The comparison of Figs. 6 and 4 shows that, in the case
of cluster W3, the degeneracy between age and extinction
can be broken with the simultaneous use of the UBVIKs
colours and the 1 -- 2.3 µm spectrum. Synthetic near-IR
spectra at ages of 500 -- 700 Myr, with low extinction opti-
cal depths, provide satisfactory UBVIKs colours but the
molecular bands they display are too pronounced to be
acceptable. As a matter of fact, the observed molecular
bands are rather shallow. Models younger than 250 Myr
are rejected because of their UBVIKs colours. The over-
lap of the contoured regions of the two figures identifies
the models that comply with all the constraints. Our pref-
ered solar metallicity model corresonds to [t=300 Myr;
AV = 0.8]. Taking into account the various sources of
errors mentioned above (extinction laws, weighting of
the spectral elements, photometric filter passbands and
zero points), acceptable models lie between [t=200 Myr;
AV = 1.0] and [t=400 Myr; AV = 0.5].
Fig. 7 shows a goodness-of-fit map constructed as a
linear combination of the ones previously discussed. The
weights are chosen such as to give the set of 4 broad-band
colours and the large set of near-IR flux densities equal im-
portance in the combined ∆2. Maps of this type will allow
us to compare model adjustments when varying internal
parameters such as the metallicity or the temperature of
the input LPV spectra.
6. Constraints on metallicity and on the
temperature scales of LPV spectra
Previous studies of the clusters of NGC 7252 suggested
metallicities between Z⊙ and ZLMC (see Section 2.1). Our
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
11
10+10
5
2
10+9
5
2
10+8
5
2
10+7
0.0
1.53 1.47
1.78
0.5
1.0
1.5
2.0
Fig. 7. ∆2 map obtained in the age-extinction plane con-
sidering simultaneously the broad-band photometry and
the near-IR spectrum of W3 cluster (Z=0.02, extinction
law as in Fig. 4 and 6).
models favour a solar metallicity for cluster W3, although
LMC metallicity cannot be safely rejected.
The
combined goodness-of-fit map,
an LMC-
metallicity equivalent of Fig. 7, has a very similar
appearance, but the minimum value is 1.79 instead of
1.47 at Z⊙ (cf Fig. 8, top). The location of the minimum
is unchanged: [t=300 Myr; AV =0.8].
The comparison of the middle plot of Fig. 8 with Fig. 4
suggests that at a given age, the intrinsic colours of a stel-
lar population are globally bluer when the stellar metal-
licities (and therfore opacities) are reduced and a higher
value of AV compensates for this effect.
At [t=300 Myr; AV =0.8], both the near-IR and the
UBVIKs ∆2 values are slightly higher than those obtained
at solar metallicity. In the near-IR, the difference is not
significant. The difference in the UBVIKs map is more
important and according to our discussion in Sec. 5.1 ∆ >
2 is not an acceptable fit. However, a reasonable change in
the adopted extinction law produces acceptable UBVIKs
colours. For instance, ∆2 is reduced from 3.37 to 1.6 when
we use the extinction law of Cardelli et al. (1989) with
R=Av/E(B-V)=5. All creteria for a good fit are met at
[t=300 Myr; AV =0.8.
At solar metallicity, a significant fraction of the TP-
AGB stars are oxygen rich. Assuming a solar metallicity,
we may thus attempt to constrain the temperature scale
of the M type LPV spectra: the predictions of near-IR
spectrophotometric properties of intermediate-age stellar
populations (mainly narrow-band molecular indices) show
a rather strong dependence upon this calibration (Lan¸con
et al. 1999, Mouhcine & Lan¸con 2002). If one affects high
10+10
5
2
10+9
5
2
10+8
5
2
10+7
0.0
10+10
5
2
10+9
1.38
5
2
10+8
5
2
10+7
0.0
10+10
1.49
1.3
1.5
5
2
10+9
5
2
10+8
5
2
1.79
0.5
1.0
1.5
2.0
1.86
3.37
0.5
1.0
1.5
2.0
1.51
1.3
10+7
0.0
0.5
1.0
1.49
2
1.5
1.5
2.0
Fig. 8. Goodness-of-fit maps for models at Z=0.008. The
layout and the ∆2 computations are as in Figs. 7, 4 and 6.
Top: broad band colours and SOFI spectrum combined;
middle: colours only; bottom: SOFI spectrum only.
12
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
10+10
5
2
10+9
5
2
10+8
5
2
10+7
0.0
10+10
5
2
10+9
5
2
10+8
5
2
10+7
0.0
1.75
1.37 1.37
1.3
1.38
1.37
0.5
1.0
1.5
2.0
1.73
0.5
1.0
1.5
2.0
Fig. 9. Goodness-of-fit map obtained in the age-extinction
plane for solar metallicity models using the Bessell et al.
(1989a) effective temperature scale for TP-AGB oxygen
rich stars. Top: only the near-IR spectrum of the W3 clus-
ter is considered. Bottom: both the broad-band photome-
try and the near-IR spectrum are considered.
effective temperatures to stellar spectra with deep molec-
ular features in the input stellar library of the models,
one produces integrated spectra that also have pronounced
molecular features. As discussed by Lan¸con & Mouhcine
(2002), conclusions on such a calibration will be model de-
pendent, in the sense that they are strongly linked to the
way the effective temperatures of the evolutionary tracks
have been computed.
Two different effective temperature scales for LPV
spectra may be considered. The first one is that LPV stars
may have the same effective temperature scale as static gi-
ant stars; the second one is that the effective temperature
scale of LPVs may differ from that of the static giants.
In principle one may expect that, for the same effective
temperature, a variable giant have redder colours than
its parent static giant star, because pulsation makes the
atmosphere of the star more extended, and consequently
redder (Bessell et al. 1989b).
We have constructed two grids of models assuming two
different effective temperature scales for the LPV stars.
The first one is taken from Bessell et al. (1989a) who con-
sider only static M stars (though already with extended
atmospheres). The authors suggest that the proposed scale
is appropriate for AGB stars. The second effective temper-
ature scale is one proposed by Feast (1996), based on an-
gular diameter measurements for late-type stars combin-
ing Miras and non-Mira M-type stars. The second tem-
perature scale is much steeper than the first one at the
low temperatures relevent for TP-AGB stars; smaller de-
creases in temperature are needed to explain the same
spectral evolution to redder colour and deeper molecular
bands. For the evolution of integrated properties of stellar
populations, this means that models calculated with the
temperature scale of Bessell et al. (1989a) will produce in-
tegrated stellar population spectra with more pronounced
near-IR molecular features than those calculated assuming
the scale of Feast (1996).
The goodness-of-fit map for models that assume the
static giant star scale from Bessell et al. (1989a), com-
bining the constraints from the broad-band photome-
try and the near-IR spectrum, is shown in Fig. 9. All
previous figures assumed the effective temperature scale
of Feast (1996). The adopted temperature scale has al-
most no effect on the evolution of broad-band photome-
try; the goodness-of-fit map obtained comparing broad-
band colours to models constructed assuming Bessell et
al. (1989a) scale is similar to Fig. 4, obtained using the
scale of Feast. The map shows that adopting the temper-
ature scale of Bessell et al. for the O-rich spectra degrades
the fits. This temprature scale assigns a higher effective
temperature to a given spectrum than the scale of Feast.
Thus, the use of spectra with strong molecular bands is ex-
tended to higher temperature regions in the HR diagram,
and molecular features are also stronger in the integrated
spectrum of the population. As a result, the VO band
(1.05 µm) is present at ages as young as 300 Myr and the
wings of the H2O bands curve the H and K-band energy
distributions (in a way similar to what is shown in Fig. 3.
7. The influence of the TP-AGB evolutionary
tracks
As discussed in Sec. 3, the treatment of the TP-AGB evo-
lutionary phase in population synthesis models is a ma-
jor challenge. The predicted evolution of intermediate-age
stellar populations (mainly for ages in the range of 0.1 to
∼ 1.5 Gyr) from different sets of models show large dif-
ferences. In this Section we will compare our results with
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
13
those obtained with other models that use a different mod-
elling of the TP-AGB evolutionary phase.
We have compared the observed spectra of W3 to the
model sets of Lan¸con et al. (1999). These models were
calculated using the prescriptions of Groenewegen & de
Jong (1993) for the TP-AGB phase. Note that the effect of
the envelope burning on the luminosity of TP-AGB stars
was negelected in these models. The predicted evolution of
broad-band colours by Lan¸con et al. (1999) models is simi-
lar to the prediction made by the recent version of Bruzual
& Charlot models as cited in Maraston et al. (2001), at
least regarding (V-K) and (B-V). For the stellar popula-
tions in the age range of 0.1 − 1 Gyr, the age interval into
which falls W3, the models of Lan¸con et al. (1999) pre-
dict that the reddest optical/near-IR or near-IR/near-IR
colours are reached at ages of ∼ 200 Myr. The colours
become bluer afterwards, up to 1 Gyr. This discrepancy
relative to our present models, where the reddest colours
are obtained for stellar populations at ages ∼ 0.8 − 1 Gyr,
is due to the fact that the TP-AGB lifetime of massive
TP-AGB stars, and hence their contribution to integrated
light, is overestimated when the overluminosity due to en-
velope burning is neglected. Note that the new models
reproduce the evolution of the contribution of the AGB
stars to the bolometric luminosity as a function of age, as
observed in LMC clusters (Frogel et al. 1990).
We find that Z=0.008 models from Lan¸con et al. (1999)
only provide marginal matches to broad-band photometry
with any extinction or age (∆2 ≃ 4). At solar metallicity,
the UBVIKs can be matched well, but not without ex-
tinction. The "best fit" found with this model set is at
[t=300 -- 400; AV =0.6] again. Indeed, the ML2002 models
and those of Lan¸con et al. (1999) produce very similar
spectra at an age of 300 Myr.
Maraston et al. (2001) find young photometric ages even
with the assumption of negligible extinction. Comparisons
of their models and ours show clear discrepancies. Their
models predict redder colours than ours at the same age.
At ages between 0.2 Gyr and 1 Gyr, (B-V) predicted by
Maraston et al. (2001) is redder than ours by 0.15 magni-
tudes or more, their (V-K) is bluer than ours by 0.1 mag.
at 0.2 Gyr, and redder by 0.1 mag. at 1 Gyr. At the ap-
proximate age of W3, a change of 0.15 mag in (B-V) corre-
sponds to a change of 50 − 60 % in age. The relative young
photometric ages derived by Maraston et al. (2001) can be
explained partly by the differences in the (B-V) evolution.
The inclusion of the TP-AGB stars in both models is quite
different, and may be another source of the differences be-
tween their and our results for the cluster photometric age.
At fixed age, their models predict a higher contribution of
the AGB stars to the K-band light than ours. Indeed, at
the approximate age of W3, our models predict that AGB
stars contribute ∼ 45% to the integrated K-band light,
while the models used by Maraston et al. (2001) predict
that these stars contribute of ∼ 55%.
As noted in the previous Sections, our best fit has
an age of 300 Myr and AV =0.7-0.8. If we had used
only UBVIKs photometry as constraints, we would have
favoured older ages (600-800 Myr) and lower extinction
values (AV =0 -- 0.3). This would have been consistent
with the early photometric age estimates of Miller et
al. (1997), who assumed low dust optical depths. Op-
tical spectroscopy leads to younger ages (300-600 Myr:
Schweizer & Seitzer 1998; 250-300 Myr, Maraston et al.
2001). Our ages agree with previous spectroscopic ages,
which are independent of extinction.
8. Summary and conclusions
Several of the star clusters in NGC 7252 are sufficiently
massive to overcome the problem of the large stochastic
fluctuations due to the small number of bright AGB stars
in intermediate-age star clusters in nearby galaxies. This
gives the unique opportunity to test the accuracy of the-
oretical modelling of intermediate-age populations in the
near-IR. With this aim, we have obtained near-IR spec-
troscopic and photometric observations of star clusters in
the merger remnant NGC 7252.
We used new models for simple stellar populations,
in which the contribution of AGB stars to the integrated
light is treated in a consistent way. We were able to match
the relevant observational constraints on the intermediate-
age AGB stellar population.
UBVIKs photometry imposes the global range of ages
but with a degeneracy between age and extinction. The
absence of independent extinction value determinations
for the target has a significant effect on model rejection
criteria (at AV > 0.5). Higher metallicities move solutions
to older ages and/or higher extinction values.
At intermediate ages, the strength of the molecular
bands is a discriminant. Fitting the near-IR spectrum of
the W3 cluster shows that carbon stars are needed to ob-
tain an acceptable fit of the data. Considering only oxy-
gen rich stars as representative of the whole AGB stellar
populations leads to synthetic spectra with spectral fea-
tures that are not compatible with the observations. This
is the first clear indication of the presence of carbon stars
outside Local Group galaxies. Our models show that the
available UBVIKs photometry is not sufficient to detect
carbon stars in the star cluster system of NGC 7252. Using
the strength of the molecular bands as discriminator, we
derived constraints on the W3 age: strong TP-AGB fea-
tures appear after ∼ 300 Myr, weaker ones before. W3 is
intermediate. We mention that the low S/N of our data is
a severe limitation of the diagnostic power of the near-IR
spectral features. Spectro-photometric data are consistent
with a metallicity 0.4−1 Z⊙ for W3, with higher likelihood
for solar metallicity.
Contrary to previous studies of star cluster system of
NGC 7252 that assumed negligible extinction, we found
14
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
that good fits of the data require high extinction inde-
pendently of metallicity (i.e., AV = 0.8). Using the near-
IR constraints has allowed us to break the age-extinction
degeneracy, to show that the assumption of low extinc-
tions was incorrect, and to find ages consistent with opti-
cal spectroscopic (extinction- independent) ages.
The comparison of the observed near-IR spectrum of
cluster W3 to grids of models calculated using our evolu-
tionary tracks with various TP-AGB effective temperature
scales, shows that the data are better reproduced when an
effective temperature scale calibrated on LPV stars is used
(i.e. steeper than the effective temperature scale of static
giant stars).
Acknowledgements. Thanks to Dr P. Francis to provide us the
dome flats and the bad pixels mask used for the data reduction.
References
Aaronson, M., & Mould, J., 1980, ApJ, 240, 804
Bessell, M.S., Brett, J.M., Scholz, M., & Wood, P.R., 1989a,
A&AS, 77, 1
Bessell, M.S., Brett, J.M., Scholz, M., & Wood, P.R., 1989b,
A&A, 213, 209
Bessell, M.S., 1990, PASP, 102, 1181
Blanco, B.M., Blanco, V.M., & McCarthy, M.F., 1978, Nature,
271, 638
Blocker, T., & Schonberner, D. 1993, in IAU Symp. 155, Plan-
etary Nebulae, ed. R. Weinberger & A. Acker (Dordrecht:
Kluwer Acad. Publ.), 479
Bressan, A., Fagotto, F., Bertelli, G., & Chiosi, C., 1993,
A&AS, 100, 647
Bruzual, G. & Charlot, S., 1993, ApJ, 405, 538
Calzetti, D., Kinney, A.L., & Storchi-Bergmann, T., 1994, ApJ,
429, 582
Cardelli, J.A., Clayton, G.C., & Mathis, J.S., 1989, ApJ, 345,
245
Carlson, M.N., Holtzman, J.A., Watson, A.M., et al., 1998, AJ,
115, 1778
Charlot, S. 1996, in From Stars to Galaxies: the Impact of
Stellar Physics on Galaxy Evolution, ed. C. Leitherer, U.
Fritze-von Alvensleben, J.P Huchra, ASP Conf. Ser., 98,
275
Charlot, S. & Bruzual, G., 1991, ApJ, 367, 126
Charlot, S., Worthey, G., & Bressan, A. 1996, ApJ, 457, 625
Cook, K.H., Aaronson, M., & Norris, J., 1986, ApJ, 305, 634
Elias, J. H., Frogel, J. A., Matthews, K., & Neugebauer, G.
1982, AJ, 87, 1029
Feast, M.W., 1996, MNRAS, 278, 11
Ferraro, F., Fusi Pecci, F., Testa, V., et al., 1995, MNRAS,
272, 391
Fagotto, F., Bressan, A., Bertelli, G., & Chiosi, C. 1994, A&AS,
105, 29
Fioc, M., & Rocca-Volmerange, B., 1997, A&A, 326, 950
Fritze-von Alvensleben, U., & Gerhard, O., 1994, A&A, 285,
775
Fritze-von Alvensleben, U. 1999, in Spectrophotometric Dating
of Stars and Galaxies, ed. Hubeny I., Heap S.R., Cornett
R.H., ASP Conf. Ser., 192, 273
Frogel, J.A., Persson, S.E., Aaronson, M., & Matthews, K.,
1978, ApJ, 220, 75
Frogel, J.A., 1985, ApJ, 298, 528
Frogel, J.A., Mould, J.R., & Blanco, V.M., 1990, ApJ, 352, 96
Girardi, L. 1996, Mem. Soc. Astr. It., 67, 839
Girardi, L. & Bertelli, G., MNRAS, 1998, 300, 533
Groenewegen, M.A.T. & de Jong, T., 1993, A&A, 267, 410
Groenewegen, M.A.T. & de Jong, T., 1994, A&A, 283, 463
Goldader, J.D., Joseph, R.D., Doyon, R., & Sanders, D.B.,
1997, ApJ, 474, 104
Grebel, E.K. 2000, in Massive Stellar Clusters, ed. A. Lan¸con
& C. Boily, ASP Conf. Ser., 211, 262
Habing, H.J. 1996, A&AR, 7, 97
Holtzman, J.A., Burrows, C.J., & Casertano, S. 1995, PASP,
107, 1065
Holtzman, J.A., Watson, A.M., & Mould, J.R., et al. 1996, AJ,
112, 416
Iben, I. & Renzini A., 1983, ARA&A, 21, 27
Lan¸con, A. 1999,
in IAU Symp. 191 on Asymptotic Giant
Branch Stars, ed. T. Le Bertre, A. L`ebre, C. Waelkens (San
Francisco: ASP), 579
Lan¸con A., & Mouhcine M., 2000, in Massive Stellar Clusters,
ed. Lan¸con A., Boily C.M., ASP Conf. Ser. 211, 34
Lan¸con, A. & Mouhcine, M., 2002, A&A, in press
Lan¸con, A. & Mouhcine, M. 2001, in Starbursts Near and Far,
ed. L. Tacconi & D. Lutz (Berlin:Springer), in press
Lan¸con, A. & Rocca-Volmerange, B., 1996, New Astron., 1,
215
Lan¸con, A., Rocca-Volmerange, B., & Thuan, T.X. 1996,
A&AS, 115, 253
Lan¸con, A. & Mouhcine, M., 2000, in Massive Stellar Clusters,
ed. Lan¸con A. & Boily C.M., ASP Conf. Ser., 211, 34
Lan¸con, A. & Wood, P.R., 2000, A&AS, 146, 217
Lan¸con, A., Mouhcine, M., Fioc, M., & Silva, D., A&A, 1999,
344, L21
Landolt, A.U., 1983, AJ 88, 439
Lejeune, T., Cuisinier, F., & Buser, R., 1998, A&A, 130, 65
Lejeune, T., & Buser, R., 1999, in Spectrophotometric Dating
of Stars and Galaxies, ed. I. Hubeny, S. R. Heap, R. H.
Cornett, ASP Conf. Ser., 192, 211
Loidl, R., Lan¸con, A., & Jørgensen, U., 2001, A&A, 371, 1065
Lutz, D., 1991, A&A, 245, 31
Maraston, C., Kissler-Patig, M., Brodie, J.P., Barmby, P., &
Huchra, J.P., 2001, A&A, 370, 176
Miller, B.W., Whitmore, B.C., Schweizer, F., & Fall, M., 1997,
AJ, 114, 2381
Moller, C.S., Fritze-v. Alvensleben, U., & Fricke, K.J. 1997,
A&A, 317, 676
Mouhcine, M., & Lan¸con, A., 1999 b, In Spectrophotometric
Dating of Stars and Galaxies, ed. I. Hubeny, S. R. Heap,
R. H. Cornett, ASP Conf. Ser. 192, 113
Mouhcine, M., & Lan¸con, A., 1999 a., In Spectrophotometric
Dating of Stars and Galaxies, ed. I. Hubeny, S. R. Heap,
R. H. Cornett, ASP Conf. Ser. 192, 303
Mouhcine, M., & Lan¸con, A., 2000., In Massive Stellar Clus-
ters, ed. Lan¸con A., Boily C., ASP Conf. Ser. 211, 144
Mouhcine, M., & Lan¸con, A., 2002 a, A&A, accepted (ML2002)
Mouhcine, M., & Lan¸con, A., 2001 b, MNRAS, submitted
Oliva, E., Origlia, L., Kotilainen, J.K., & Moorwood, A.F.M,
1995, A&A, 301, 55
Olofsson, H., 1999, in IAU Symp. 191 in Asymptotic Giant
Branch Stars, ed. T. Le Bertre, A. L`ebre, C. Waelkens (San
Francisco: ASP), 579
Mouhcine et al.: Near-IR Spectrum of Cluster in NGC 7252
15
Origlia, L., Goldader, J.D., Leitherer, C., Schaerer, D., & Oliva,
E. 1999, ApJ, 514, 960
Persson, S.E., Aaronson, M., Cohen, J.G., Frogel, J.A., &
Matthews, K., 1983, ApJ, 266, 105
Persson, S.E., Murphy, D.C., Krzeminski, W., Roth, M., &
Rieke, M.J., 1998, AJ, 116, 2475
Pritchet, C.J., Schade, D., Richer, H.B., Crabtree, D., & Yee,
H.K.C., 1987, ApJ, 323, 79
Rieke, G.H., & Lebofsky, M.J., 1979, ARA&A, 17, 477
Salasnich, B., Girardi, L., Weiss, A., & Chiosi, C. 2000, A&A,
361, 1023
Salpeter, E. E. 1955, ApJ, 121, 161
Santos, J.F.C., & Frogel, J.A., 1997, ApJ, 479, 764
Silva, D.R. 1996, in Spiral Galaxies in the Near-IR, eds. Minniti
D., Rix H.-W. (Berlin: Springer), p.3
Schweizer, F., 1982, ApJ, 252, 455
Schweizer, F., & Seitzer, P., 1998, AJ, 116, 2206
Wagenhuber, J., & Groenewegen, M.A.T., 1998, A&A, 340,
183
Westerlund, B.E., Azzopardi, M., Rebeirot, E., & Breysacher,
J., 1991, A&AS, 91, 425
Whitmore, B.C., Schweizer, F., Leitherer, C., Borne, K., &
Robert, C., 1993, AJ, 106, 1354
Whitmore, B.C., & Schweizer, F., 1995, AJ, 109, 960
Whitmore, B.C., Zhang, Q., Leitherer, C., et al., 1999, AJ, 118,
1551
Yi, S., Demarque, P., Kim, Y.-C. et al. 2001, astro-ph/0104292
|
astro-ph/0107344 | 1 | 0107 | 2001-07-18T16:19:37 | Deep HST WFPC2 Photometry of M31's Thick Disk(?) | [
"astro-ph"
] | We present deep color-magnitude diagrams (CMDs) for a field along the outer disk of M31 based on archival Hubble Space Telescope Wide Field Planetary Camera 2 observations in the F555W (~V) and F814W (~I) filters. The CMDs, which contain a total of about 50,000 stars, feature a prominent red giant branch (RGB) along with a significant population of helium burning red clump stars. In addition, they exhibit the rarely seen asymptotic giant branch clump as well as a weak `Pop II' horizontal branch. There is also the hint of a ~2 Gyr subgiant branch at the faintest levels of the CMDs. After adopting an M31 distance of (m-M)o = 24.5 and a reddening of E(B-V) = 0.08, we draw the following conclusions. 1) The I-band absolute magnitude of the helium burning red clump stars is M(RC) = -0.29 +/- 0.05, which is in accord with the value derived from Hipparcos parallaxes of solar neighborhood clump stars by Stanek & Garnavich. 2) The metallicity distribution function constructed from bright RGB stars shows a characteristic shape; however, a pure halo population consisting of metal-poor and intermediate metallicity components (as advocated in the literature) are not sufficient to account for this shape. Instead, an additional Gaussian component with <[Fe/H]> = -0.22 +/- 0.26, comprising 70% of the total number of stars, is required. 3) A comparison of our CMD with theoretical isochrones indicates that the majority of stars in our M31 field have ages that are >~1.5 Gyr. 4) These points, along with the physical location of our field in M31, suggest that we are observing the thick disk population of this galaxy. | astro-ph | astro-ph |
Deep HST WFPC2 Photometry of M31's Thick Disk(?)
Ata Sarajedini1
Department of Astronomy, University of Florida, P. O. Box 112055, Gainesville, FL 32611-2055
[email protected]
Jeffrey Van Duyne2
Astronomy Department, Wesleyan University, Middletown, CT 06459
[email protected]
ABSTRACT
We present deep color-magnitude diagrams (CMDs) for a field along the outer disk
of M31 based on archival Hubble Space Telescope Wide Field Planetary Camera 2
observations in the F555W (∼V) and F814W (∼I) filters. The CMDs, which contain
a total of about 50,000 stars, feature a prominent red giant branch (RGB) along with
a significant population of helium burning red clump stars. In addition, they exhibit
the rarely seen asymptotic giant branch clump as well as a weak 'Pop II' horizontal
branch. There is also the hint of a ∼2 Gyr subgiant branch at the faintest levels of
the CMDs. After adopting an M31 distance of (m − M )0 = 24.5 and a reddening of
E(B − V ) = 0.08, we draw the following conclusions. 1) The I-band absolute magnitude
of the helium burning red clump stars is MI (RC) = −0.29 ± 0.05, which is in accord
with the value derived from Hipparcos parallaxes of solar neighborhood clump stars by
Stanek & Garnavich. 2) The metallicity distribution function constructed from bright
RGB stars shows a characteristic shape; however, a pure halo population consisting of
metal-poor and intermediate metallicity components (as advocated in the literature)
are not sufficient to account for this shape. Instead, an additional Gaussian component
with h[F e/H]i= −0.22 ± 0.26, comprising 70% of the total number of stars, is required.
3) A comparison of our CMD with theoretical isochrones indicates that the majority
of stars in our M31 field have ages that are >∼1.5 Gyr. 4) These points, along with
the physical location of our field in M31, suggest that we are observing the thick disk
population of this galaxy.
Subject headings: galaxies: individual (M31) -- Local Group -- stars: abundances --
color-magnitude diagrams
1Guest User, Canadian Astronomy Data Centre, which is operated by the Dominion Astrophysical Observatory
for the National Research Council of Canada's Herzberg Institute of Astrophysics.
2Current Address: Department of Astronomy, Yale University, P. O. Box 208101, New Haven, CT 06520-8101
-- 2 --
1.
Introduction
The determination of star formation histories for spiral galaxies is an important ingredient in
constraining models of galaxy formation (Bullock 1999; Grebel 2000; Silk 2000). The three most
readily available subjects are, of course, our own Milky Way galaxy, M31, and M33. Being within
the Milky Way makes much of this work difficult, which is why a good understanding of the disk
and halo of M31 is so important. However, little is known about the early star formation history
of M31 due primarily to the lack of sufficiently deep photometry of its disk and halo.
There is a rich history of ground-based work dealing with the stellar populations of M31. The
reader is referred to the review by van den Bergh (1999) for a discussion of many of these results.
For the purposes of the present paper, we concentrate on previous photometric investigations of the
outer disk and inner halo. We begin by noting that ground-based photometry has only generated
color-magnitude diagrams (CMDs) of M31 field stars as deep as the horizontal branch, which is
roughly at I=24.5 (Durrell, Harris, & Pritchet 1994). Nevertheless, a great deal can be learned
about M31 by examining the colors and magnitudes of the brighter stars.
Pritchet & van den Bergh (1988) present photometry for an inner halo field located 40 arcmin
southeast of the nucleus along the minor axis. They find a mean abundance of [F e/H] ∼ −1.0,
similar to the Galactic globular NGC 6171, with a dispersion of ∼0.3 dex. The work of Davidge
(1993) on an inner halo field situated on the opposite side of the galaxy near NGC 205 confirms these
results; namely, he finds a mean [F e/H] between -- 1.3 and -- 0.7 along with a metallicity dispersion of
∼0.3 dex. Davidge (1993) also presents observations that sample the outer disk of M31 at a location
25.6 arcmin from the nucleus. As one would expect, this field presents a rather complicated stellar
population profile, suggesting a significant range in age and metallicity among the stars. Previously,
Hodge & Lee (1988; see also Hodge, Lee, & Mateo 1988 and Massey, Armandroff, & Conti 1986)
imaged six disk fields in M31 and were able to study reddening variations in their fields along with
the properties of the luminosity functions. The color-magnitude diagrams published by Richer,
Crabtree, & Pritchet (1990) and Morris et al. (1994) reveal the metal-rich nature of the M31 disk.
The latter study finds a mean metal abundance that is significantly higher than that of 47 Tuc
([F e/H] = −0.7). Recent wide field imaging observations of M31's outer disk by Cuillandre et
al. (2001) are used to investigate the correlation between the young star population, HI column
density, and the dust content.
Extending the ground-based studies with the Hubble Space Telescope (HST) has resulted in
CMDs that reach 2 -- 3 magnitudes fainter than the M31 field horizontal branch. In spite of this,
HST has been used relatively sparingly in examining the field halo and disk populations of M31
(Holland, Fahlman, & Richer 1996; Rich, Mighell, & Neill 1996). Instead, the majority of HST
photometry of M31 has been directed at the star clusters (Fusi Pecci et al. 1996; Holland et al.
1997; Jablonka et al. 2000).
One of these HST programs (GO-5420) was designed to construct CMDs of M31 star clusters
using broad-band filter observations with the Wide Field and Planetary Camera 2 (WFPC2). One
-- 3 --
object on the target list was intended to be the M31 globular cluster G272 (α2000 = 00h44m14.5s,
δ2000 = 41◦ 19' 19.8"). Apparently however, a bright nearby star was imaged instead (α2000 =
00h44m47.7s, δ2000 = 41◦ 18' 48"). As a result, the WFPC2 observations of this field were appar-
ently ignored and do not appear in any of the subsequent analysis papers which focus mainly on
the clusters (Fusi Pecci et al. 1996).
Having come across these images in the HST archive and realizing that the exposure times
were extraordinarily long (Section 2), the scientific utility of these observations became apparent.
As discussed in Sec. 2, this region of M31 samples both the disk and halo stellar populations.
The next section describes the observations and data reduction. Section 3 presents the resulting
CMDs and the properties of the field that can be determined therein, such as the magnitude of the
red clump and the metallicity distribution function. This section also contains the results of our
artificial star experiments, further analysis of the populations and their properties, and discussions
of our results. Section 4 presents our conclusions and suggestions for future work.
2. Observations and Data Reduction
Our field center is situated at a α2000= 00h44m50.6s and a δ2000= +41o19′11.1′′, which is
positioned at a ∼45o angle from the minor axis of the M31 disk, as seen in Fig. 1a. Figure 1b is an
expanded view of this area, showing the HST-WFPC2 orientation on the sky. A scale conversion
from arcminutes to kpc at our adopted M31 distance of 790 kpc (Da Costa et al. 2000; Durrell,
Harris, & Pritchet 2001) places our field at a projected distance of 5.5 kpc from the center of M31.
We obtained the processed images of the 'G272' field from the Canadian Astronomy Data
Centre3. The observations were taken on 1995 January 23 and include two frames taken in the
F555W (∼ V) filter and five taken with the F814W (∼ I) filter with total exposure times of 3800s
and 10,800s, respectively. The Wide Field 3 chip was ignored due to the extreme saturation effects
of the foreground star, as is clearly seen in Fig. 1c. Table 1 provides more detailed information
about the observations.
The photometric reduction was performed on the three remaining chips (PC1, WF2, and
WF4), in a manner similar to the procedure described in Sarajedini et al. (2000), utilizing the
DAOPHOT II / ALLSTAR / ALLFRAME (Stetson 1994) profile fitting software. The reader is
referred to that paper for details. In summary, high single-to-noise WFPC2 F555W and F814W
point spread functions (PSFs), kindly provided by Peter Stetson, were fitted to all detected profiles
on each individual image using the ALLFRAME software. The resultant instrumental magnitudes
were edited and matched to form colors. Aperture corrections were applied as described in the
next paragraph to bring the total magnitudes to a 0.5" radius. Then, a correction for the well-
3Canadian Astronomy Data Centre is operated by the Dominion Astrophysical Observatory for the National
Research Council of Canada's Herzberg Institute of Astrophysics.
-- 4 --
known charge transfer efficiency (CTE) problem was applied (Sarajedini et al. 2000) and standard
magnitudes were calculated using the equations of Silbermann et al. (1996). These describe the
photometric system established by the HST "Cepheid Distance Scale" key project and are coupled
to the PSFs we used earlier in the reduction procedure.
The only departure from the procedure adopted by Sarajedini et al. (2000) was in the deter-
mination of the aperture corrections for the PC1 chip. Because of the lack of sufficient numbers
of bright stars in the Planetary Camera, we adopt the following procedure to ensure consistency
between the photometric scales of the three CCDs. We derived and applied aperture corrections
to the WF2 and WF4 data as described by Sarajedini et al. (2000). Then, under the assumption
that the spatially adjacent regions of the three chips feature the same peak magnitude for the red
clump (see Figs. 1 and 2), we offset the photometry from each chip to match the average peak red
clump magnitude of the WF photometry. The offsets to the WF data (0.03 mag in V and 0.04 in
I) correct systematic offsets between the photometric scales of these two CCDs and the offsets to
the PC1 data (0.03 mag in V and I) account for the fact that no aperture corrections were applied.
In any case, after the application of these offsets, we estimate that the photometric scales of the
three CCDs are consistent to within ±0.015 mag.
3. Results and Discussion
3.1. Color-Magnitude Diagrams
The left panels of Figures 2a through 2c show the color-magnitude diagrams (CMDs) of the
PC1 (5249 stars), WF2 (22793 stars), and WF4 (20943 stars) fields, respectively, in the apparent
(V, V − I) plane, while the right panels show the same fields in the (I, V − I) plane. The two most
prominent features in these CMDs are the well-populated red giant branch (RGB) which exhibits a
large range in color and an obvious red clump at V∼25.5 and I∼24.3. There is also a hint of a 'Pop
II' horizontal branch located blueward of the red clump extending from V −I∼0.1 to V −I∼0.7 with
V ∼25.4 and appearing most prominently in the WF2 CMD. This feature is likely to be associated
with the relatively small metal-poor population present in our field as discussed in Sec. 3.6. The
tip of the first ascent RGB appears to be at I∼21.8. All of the CMDs also feature a main sequence
(MS) of young stars with ages between 108 and 109 years (see below). Note also the existence of
what appears to be a subgiant branch developing at V ∼26.5 (see Fig. 12 and the discussion in
Sec. 3.7). Lastly, there is a clump of stars at V ∼24.7 and I∼23.2, which we tentatively identify
with the asymptotic giant branch (AGB) clump (Gallart 1998).
To investigate radial variations in the CMDs, we divided our sample into three regions (1, 2, and
3) based on radial distance from the center of M31. Each star's right ascension and declination was
determined using the IRAF4 routine METRIC and then separated into three 1 arcmin (projected)
4IRAF is distributed by the National Optical Astronomy Observatories which are operated by the Association of
-- 5 --
wide regions. The black lines in Fig. 1c show these divisions; region 1, which is closest to the M31
nucleus, is on the right hand side of this figure.
3.2. Distance and Reddening
The reddening of our M31 field was estimated using the Burstein & Heiles (1982) maps. At the
location of our field (l = 121.60, b = −21.53), these maps indicate a reddening of E(B − V ) = 0.08,
which is consistent with the value used by Holland et al (1999). We adopt the relations E(V − I) =
1.25E(B − V ) and AI = 1.48E(V − I) from Schlegel et al. (1998) for the HST filters utilized herein.
It is important to point out at this juncture that we have deliberately neglected the effects of
differential reddening across our field and along the line of sight. Differential reddening due to the
Galactic foreground across our (small) WFPC2 field of view is likely to be negligible. However, the
effects of dust internal to M31 may have a significant effect on our results. We return to this point
in Sec. 3.6.
For the distance modulus of M31, we adopt a value of (m − M )0 = 24.5 ± 0.1 based on the
mean of those presented by Da Costa et al. (2000), who quote distances to M31 based on the field
halo RR Lyraes and giant stars as well as the M31 globular clusters. This distance modulus is on
the scale of Da Costa & Armandroff (1990). Figures 3a, b, and c depict our distance and reddening
corrected CMDs for the three radial regions defined above. We note that the appearance of the
three CMDs is qualitatively indistinguishable.
A particularly striking feature is that the tip of the RGB, at MI ∼ −3, is significantly fainter
in all three CMDs than the canonical value of MI = −4.05 ± 0.10 (Da Costa & Armandroff 1990,
hereafter DCA; Sakai et al. 2000; Bellazzini, Ferraro, & Pancino 2001). In contrast, the RGB tip
is easily identified at I∼20.6 (MI ∼ -- 4.0) in Fig. 2 of Holland et al. (1996) showing the CMD of the
field around the M31 globular cluster G302 constructed from HST observations similar to those
considered herein. The faintness of the RGB tip is consistent with the fiducials of very metal rich
star clusters such as NGC 6553 and NGC 6528 (Bica, Barbuy, & Ortolani 1991), giving us our first
hint that this field in M31 contains a very metal-rich stellar population. More discussion of this
phenomenon is provided in Sec. 3.4.
3.3. Horizontal Branch and Red Clump
As mentioned above, the most conspicuous feature of the M31 field star CMD presented herein
is the helium burning red clump (RC). Based on Fig. 3, we find peak values of MI (RC) =
−0.28 ± 0.05, −0.30 ± 0.05, and −0.29 ± 0.05 for regions 1, 2, and 3, respectively. All of these values
Universities for Research in Astronomy, Inc. under cooperative agreement with the National Science Foundation.
-- 6 --
are identical to within the errors and very close to the red clump absolute magnitude advocated
by Stanek & Garnavich (1998); they find MI (RC) = −0.23 ± 0.03 based on Hipparcos parallaxes
of solar neighborhood red clump stars.
Not so conspicuous but nevertheless present in the CMDs is a 'Pop II' horizontal branch most
easily seen in the CMD of Region 1 shown in Fig. 3a. The location of this feature is evident
in Fig. 4 where we have superimposed the HBs and RGBs of the well-known Galactic globular
clusters M68 (Walker 1994) and M5 (Johnson & Bolte 1998). The metallicities, distance moduli,
and reddenings for these clusters are taken from Table 1 of Layden & Sarajedini (1997) adopting
AV = 3.1E(B − V ).
3.4. Asymptotic Giant Branch
As already noted, the CMDs of the G272 field show evidence for the AGB clump. Analogous
to the RGB clump (Fusi Pecci et al. 1990; Sarajedini & Forrester 1995; Ferraro et al. 1999),
the AGB feature is caused by a temporary pause in the evolution of stars as they proceed up
the AGB and is related to the formation of the Helium burning shell (Caputo et al. 1989). As
shown by Alves & Sarajedini (1999), the luminosity of the AGB clump (alternatively referred to
as the AGB bump because of its appearance in a luminosity function) is primarily a function of
the mean age and metal abundance of the stellar population. To eliminate the effects of distance
uncertainties, the magnitude of the AGB bump is measured relative to the horizontal branch/red
clump [∆V (AGBC − RC)].
Building upon the work of Castellani, Chieffi, & Pulone (1991) and Ferraro (1992), Alves &
Sarajedini (1999) present ∆V (AGBC − RC) values for four Galactic globular clusters - M5, NGC
1261, NGC 2808, and 47 Tuc, the only ones in which the AGBC could be isolated at the time.
They showed that the predictions of the theoretical models are in good agreement with the ages,
metallicities, and ∆V (AGBC − RC) values of these clusters.
For our M31 field, we have already measured the luminosity of the red clump and thus adopt
MI (RC) = −0.29 ± 0.05 as the mean of the three regions shown in Fig. 3. The average color
of the red clump is (V − I)o = 1.00 ± 0.05 making the absolute V magnitude of the red clump
equal to MV (RC) = 0.71 ± 0.07. Performing the same calculation on the AGB clump stars, we
find MV (AGBC) = −0.10 ± 0.07 leading to a difference of ∆V (AGBC − RC) = −0.81 ± 0.10.
Assuming that the G272 field is not significantly older than the Galactic globulars, Fig. 6 of Alves
& Sarajedini (1999) suggests that the metal abundance of this field is likely to be greater than
∼ -- 0.4 dex. However, because the Alves & Sarajedini models do not extend to more metal-rich
regimes, it is difficult to say anything more definitive. In any case, we have once again confirmed
the metal-rich nature of the stellar population in the 'G272' field.
-- 7 --
3.5. Artificial Star Experiments
As noted above, the RGB shows a rather large color width; one of the factors that could
influence this width is the photometric error.
In order to assess its importance, we performed
a number of artificial star experiments. These consist of placing stars of known magnitude and
color on our original CCD frames and applying the same reduction techniques as those used in the
reduction of the genuine stars. We placed these stars on each frame arranged in a grid pattern
under the constraint that no two artificial stars were within 2 PSF radii of each other. The resultant
images were reduced using the same method described in Section 2. Four trials were performed
with a total of 1668 stars recovered after reducing the three CCDs. The magnitudes of the artificial
stars were taken from a locus with 22.55 < I < 24.95 along the RGB as displayed in Fig. 5a. The
locations of the stars at their recovered magnitudes and colors are shown by the points in Fig. 5a.
The difference between the input and measured colors is shown in Fig. 5b as a function of the
apparent I magnitude.
To determine the photometric error along the RGB, we fit a Gaussian curve to the color
distribution of artificial with 22.55 ≤ I ≤ 22.75 (Fig. 5c). As discussed in the next section, we have
chosen to use the brighter RGB stars to study the metallicity distribution of this field. Thus, at I
= 22.65, which corresponds to MI = −2 using our adopted distance modulus of (m − M )I = 24.65,
the artificial stars exhibit a standard deviation of σerr = 0.049 in (V -- I). This quantity is taken to
be the mean photometric error at this magnitude (i.e. I=22.65, MI = −2) and will be used in the
metal abundance analysis of the next section.
The lower panel of Fig. 6 shows our actual photometry in the distance and reddening corrected
CMD and the solid lines are the RGBs of M15 (left) and NGC 6791 (right), which encompass
the majority of RGB stars; the dotted lines enclose the magnitude range −2.1 ≤ MI ≤ −1.9
(22.55 ≤ I ≤ 22.75) over which the color histogram shown in the upper panel has been constructed.
The solid line in the upper panel shows the weighted least squares fit of a Gaussian function to
this distribution. The standard deviation of this fit is σobs = 0.182, which is much larger than the
contribution purely from the photometric error. In fact, if we subtract σerr from σobs in quadrature,
we find an intrinsic color spread of σint = 0.175 mag which is represented by the dashed curve in
the upper panel of Fig. 6. The solid and dashed curves have been scaled to have the same area.
Note that these two distributions are virtually indistinguishable suggesting that the influence of
photometric error on the metallicity distribution function is likely to be negligible.
3.6. Metallicity Distribution Function
Now that we have utilized artificial star experiments to quantify the photometric errors on
the RGB, it is possible to use the color of each RGB star to construct a metallicity distribution
function (MDF). Figures 3a through c show our distance and reddening corrected CMDs compared
with the RGB fiducials of the Galactic globular clusters M15, 47 Tuc (both from DCA), and NGC
-- 8 --
6553 (Guarnieri et al. 1998), along with the RGB of the open cluster NGC 6791 (Garnavich et al.
1994).
We prefer to use observed cluster fiducials over theoretically calculated RGB sequences because
of the uncertainties in the color-temperature calibration and our limited knowledge of the physics
of convection. We attempted to make use of the empirical RGB grid constructed by Saviane et
al. (2000), but realized that the metallicity range of the grid is not sufficient to cover the range of
stars observed in our M31 field. In any case, over the metallicity range common to both RGB sets,
there is good agreement between the Saviane et al. (2000) grid and our RGBs discussed below.
The adopted distance moduli and reddenings used to place the cluster fiducials in Fig. 3 are
given in Table 2. The latter values are taken predominantly from the work of DCA. In the case
of NGC 6553, we adopted the reddening advocated by Guarnieri et al.
(1998) based on their
application of the simultaneous reddening and metallicity method developed by Sarajedini (1994).
For NGC 6791, we used an average of the values quoted by Garnavich et al. (1994) and Chaboyer,
Green, & Liebert (1999). The distance moduli in Table 2 deserve a more detailed explanation.
This is because we utilized a slightly different technique for clusters with RR Lyrae variables as
compared to those with red clumps (RC). To begin with, we note again that our adopted M31
distance modulus is that of Da Costa et al. (2000) and is based on MV (RR) = 0.17[F e/H] + 0.82
(Lee, Demarque, & Zinn 1990) for the RR Lyraes. For the three clusters with red clumps, we must
modify the calculated MV (RR) to take account for the fact that the red clump and the RR Lyraes
have different absolute magnitudes and that this magnitude is a function of metallicity and age
(Cole 1998; Alves & Sarajedini 1999; Sarajedini 1999; Girardi & Salaris 2001). We consider each
of the red clump clusters in turn.
47 Tuc: We make use of the results published by Sarajedini, Lee, & Lee (1995, hereafter SLL)
which are based on synthetic HB models that are consistent with our distance scale. From the
work of SLL, we see that MV (RC) = 0.64 so that (m − M )V = 13.42 and thus (m − M )I = 13.37
for 47 Tuc. This is somewhat smaller than the DCA value of (m − M )V = 13.51 which was based
on arbitrarily setting the apparent RR Lyrae magnitude of 47 Tuc 0.15 mag fainter than the red
clump magnitude. The globular cluster compilation of Harris (1996) gives (m − M )V = 13.37.
NGC 6553: Since the age of NGC 6553 is similar to that of 47 Tuc (Zoccali et al. 2001), we need
only to correct the red clump luminosity of 47 Tuc for the metallicity difference between these two
clusters. This correction gives MV (RC) = 0.71 so that (m − M )V = 16.17 and (m − M )I = 15.18
for NGC 6553. Our values compare favorably with the distance modulus estimated by Guarnieri
et al.
(1998) of (m − M )V = 15.98 ± 0.15, but differs somewhat from two of the most recent
determinations; Zoccali et al. (2001) find (m − M )V = 15.70 ± 0.13 while Beaulieu et al. (2001)
derive (m − M )V = 15.4.
NGC 6791: The metal abundance of NGC 6791 is greater than the most metal-rich models presented
by SLL. As a result, the corrections for age and metallicity will be performed relative to the MV (RC)
value of 47 Tuc using the HB models of Girardi et al. (2000, see also Crowl et al. 2001). Thus,
-- 9 --
because NGC 6791 is 0.99 dex more metal-rich than 47 Tuc and 4 Gyr younger (Chaboyer et al.
1999), we calculate its red clump to be at MV (RC) = 0.76 making its apparent distance moduli
equal to (m − M )V = 13.79 and (m − M )I = 13.63. As a comparison, we note that Garnavich et
al. (1994) used a modulus of (m − M )V ∼ 13.6 in their study of NGC 6791.
The cluster RGBs shown in Fig. 3 are used to calibrate the dereddened RGB color at MI = −2
[(V − I)o,−2] as a function of [F e/H]. This magnitude level was chosen to minimize the effects of
asymptotic giant branch stars while at the same time maximizing the effects of metallicity on color.
The weighted least squares relation shown in the lower panel of Fig. 7 is given by
[F e/H] = −27.24 + 46.18(V − I)0 − 26.49(V − I)2
0 + 5.16(V − I)3
0.
(1)
The root mean square deviation of the points from the relation is 0.03 dex. The upper panel of
Fig. 7 shows the metallicity dispersion introduced by the color error of σerr = 0.049 at MI = −2.
This figure indicates that the photometric error translates to a typical error of only ∼0.2 dex in
metallicity. Furthermore, for NGC 6553, which has the most uncertain distance modulus among
the clusters in Table 2, an uncertainty of 0.2 mag in (m − M )I also leads to a metallicity error of
∼0.2 dex.
To construct the MDF, we take the dereddened color of each star within ±0.1 mag of MI = −2
and convert it to a metal abundance using Equation 1.
In addition, the photometric error is
converted to σ[F e/H]. A generalized histogram of the metallicities is then constructed by adding
up the unit Gaussians representing the abundance of each star. The resultant MDF of the 271
stars in this magnitude range is shown in the upper panel of Fig. 8 wherein the filled circles and
the solid line are the binned and generalized histograms, respectively. This comparison helps to
illustrate which features are significant and which are diluted by the photometric errors. In this
regard, we see that the MDF displays a prominent peak at [F e/H]∼ -- 0.1, a possible secondary
peak at [F e/H]∼ -- 0.7, and an extended tail to the metal-poor regime. The lower panel of Fig. 8
displays the effect on the MDF of changing the adopted reddening by ±0.02 mag in E(B -- V). We
note that the location of the peak changes by less than 0.1 dex and the overall shape of the MDF
remains largely unchanged.
We pointed out earlier that we have neglected differential reddening along the line of sight
caused by dust internal to M31. Taking out this effect will tend to reduce the range of metallicities
present in our MDF and it could systematically lower the higher metallicity measurements. Because
it is difficult to precisely account for this, the reader should keep this possibility in mind as the
results of the analysis are presented.
-- 10 --
3.7. Comparison With Other MDFs
Figure 9 shows the MDF for our M31 field (G272) located at a projected radial distance of 23.9
arcmin compared with the M31 field halo MDFs from Holland et al. (1996, G302), Durrell et al.
(1994), and Durrell et al. (2001). These are located at projected radial distances of 32 arcmin, 40
arcmin, and 90 arcmin, respectively and have been scaled to have the same area as the 'G272' field.
In addition, we have shifted the Durrell et al. (2001) MDF by -- 0.3 in metal abundance to account
for the fact that they quote [M/H] rather than [F e/H] (Durrell 2001, private communication).
There are a number of features to note in Fig. 9. First, all of the MDFs share the same
overall shape; they feature a prominent peak at the metal-rich end with an extended tail to more
metal-poor regimes. Relative to the peak, this tail appears to be most prominent in the Holland
et al. (1996) and Durrell et al. (2001) MDFs and less so in the Durrell et al. (1994) distribution.
Second, the metallicity of the 'G272' peak occurs at [F e/H] ∼ −0.1, which is significantly more
metal-rich that those of the other MDFs (see also Mould & Kristian 1986; Pritchet & van den
Bergh 1988).
Another way in which we can intercompare these MDFs is to scale them so that their metal-
poor tails match, as shown in the bottom panel of Fig. 9. Keeping in mind that the three dashed
MDFs are predominantly halo stars in M31, we find that the peak of the halo MDF is consistent
with the 'secondary peak' at [F e/H]∼ -- 0.7 in the 'G272' distribution. This suggests that the latter
population contains not just halo stars but also a significant population from another component.
We tentatively assign this population to the thick disk of M31, an assertion that is not unrealistic
judging from the physical location of our field in M31 (Fig. 1a). Furthermore, we note that, because
M31 halo populations do not exhibit a radial abundance gradient for R >∼ 5 kpc (van den Bergh
1999), it is unlikely to be the case that the dominant population in the 'G272' field is simply a
higher metallicity halo. Additionally, based on the discussion presented by Durrell et al. (2001),
there is the possibilty that the metal-rich component in our field is actually the bulge of M31.
However, given that the central regions of the bulge are around Solar metallicity (Renzini 1999;
Jablonka et al. 1999; 2000) and that the mean abundance is expected to decrease at the rate of
∼0.1 dex/kpc (Durrell et al. 2001), we should expect a peak bulge metallicity of ∼ -- 0.6 dex at the
location of our field. If present, this population would be indistinguishable from our 'secondary
peak' at [F e/H]∼ -- 0.7.
We can quantify our claim that the metal-rich peak in our MDF belongs to the M31 thick disk
by fitting multiple Gaussian distributions to the 'G272' MDF representing the three populations
that appear to be present - metal-poor and intermediate metallicity components that belong to the
M31 halo and a metal-rich component that may be the thick disk. For the first two populations,
we adopt the Gaussian parameters in Table 3 of Durrell et al. (2001). We can then fit a third
Gaussian to our MDF to solve for the metallicity parameters (peak and width) of the M31 thick
disk. The two panels of Fig. 10 illustrate the results of this exercise with the upper panel showing
the fit to the binned histogram and the lower panel displaying the fit to the generalized histogram
-- 11 --
(see Fig. 8). There is relatively good agreement between the fitted data and the fits as well
as between the fits to the binned and generalized histograms. This procedure suggests a mean
metallicity of h[F e/H]i= −0.22 ± 0.26 for the M31 thick disk comprising 70% of the stars in this
field. This is in contrast with the abundances of the metal-poor and intermediate metallicity peaks
of [F e/H] = −1.50 ± 0.45 (10%) and [F e/H] = −0.82 ± 0.20 (20%), respectively, from Table 3 of
Durrell et al. (2001) adjusted by -- 0.3 dex.
At this point,we return momentarily to the metallicities of the HB and red clump populations.
We speculate that the most metal-poor component with [F e/H] ∼ −1.5 is likely to be associated
with the above-mentioned 'Pop II' HB while the most metal-rich population with [F e/H] ∼ −0.2
is probably producing the strong red clump. However, the location of the helium burning stars
associated with the intermediate metallicity ([F e/H] ∼ −0.8) component is unclear.
In Fig. 11, we compare the 'G272' MDF with those of the Milky Way's thick disk stars (Wyse
& Gilmore 1995), field halo stars (Ryan & Norris 1991), globular clusters (Harris 1996), and M31's
globular clusters (Barmby et al. 2000). Interestingly, the 'G272' field resembles the Milky Way's
thick disk stars more than it does any of the other halo components (e.g. M31 globulars). One
interpretation of this would be that the stellar population of the 'G272' field is dominated by the
thick disk of M31, an assertion which would support our conclusion based on the comparisons in
Fig. 9 above. However, this is a spurious line of reasoning because if we follow it to its logical end,
it implies that the Durrell et al. (2001) MDF (of the M31 outer halo) does not represent the M31
halo because it does not resemble the MDF of the Milky Way halo field stars in Fig. 11b.
3.8. Age of the Field Population
Further evidence of our assertion that the 'G272' field is dominated by M31 thick disk stars
can be obtained by examining the age structure of this field. Figure 12 shows the same CMDs
as Fig. 3 with solar abundance (Z=0.019) isochrones of 108, 6.3 × 108, 109, and 1.6 × 109 years
(Girardi et al. 2000) overplotted. From these comparisons, we note that the fraction of the stellar
population younger than 109 years is very small, comparable with the appearance of the G302 field
CMD in Holland et al. (1996). In contrast, the observations of M31's disk presented by Williams
& Hodge (2001) reveal a significantly larger population of stars with ages younger than 109 years.
This difference suggests that the contribution of the M31 thin disk to the 'G272' field is minimal.
This, coupled with the fact that the dominant population in the 'G272' field is likely to be >∼1.5
Gyr old, provides further circumstantial evidence that the 'G272' field is probably dominated by
intermediate-to-old age thick disk stars.
-- 12 --
4. Summary and Conclusions
We present the deepest HST/WFPC2 photometry of a field in M31. The VI color-magnitude
diagram is based on 3,800s of exposure time in the F555W filter and 10,800s in the F814W filter.
After adopting a distance of (m − M )0 = 24.5 and a reddening of E(B − V ) = 0.08, we draw the
following conclusions.
1) The I-band absolute magnitude of the helium burning red clump stars is MI (RC) = −0.29±0.05,
which is in accord with the value derived from Hipparcos parallaxes of solar neighborhood clump
stars by Stanek & Garnavich (1998).
2) The V-band absolute magnitude of the asymptotic giant branch (AGB) clump stars is MV (AGB) =
−0.10 ± 0.07; coupled with the red clump luminosity, this value is consistent with those predicted
by the models of Alves & Sarajedini (1999) for an intermediate age metal-rich population.
3) The metallicity distribution function constructed from bright RGB stars shows a characteristic
shape with a prominent peak at [F e/H] ∼ −0.1 and an extensive tail to metal poor regimes as low
as [F e/H] ∼ −2.5.
4) A pure halo population consisting of metal-poor and intermediate metallicity components (Dur-
rell et al. 2001) is not sufficient to account for the shape of our MDF. Instead, an additional
Gaussian component with h[F e/H]i= −0.22 ± 0.26, comprising 70% of the total number of stars,
is required.
5) A comparison of our CMD with the theoretical isochrones of Girardi et al. (2000) indicates that
the majority of stars in our M31 field have ages older than ∼1.5 Gyr.
6) All of the above points along with the physical location of our field in M31 suggest that we have
observed the thick disk population of this galaxy.
We close by emphasizing the need for a robust model describing the spatial distribution of the
various M31 components (e.g. thin disk, thick disk, bulge, halo). It is difficult for us to draw more
compelling conclusions about the stellar populations of M31 without such a model.
We are grateful to Pat Durrell, Rupali Chandar, Andy Stephens, and Pascale Jablonka for
helpful comments on an early version of this manuscript. The suggestions of an anonymous referee
also helped to clarify the presentation tremendously. This project has benefited from financial
support from NSF CAREER grant No. AST-0094048.
Alves, D., & Sarajedini, A. 1999, ApJ, 511, 225
REFERENCES
-- 13 --
Barmby, P., Huchra, J. P., Brodie, J. P., Forbes, D. A., Schroder, L. L., & Grillmair, C. J. 2000,
AJ, 119, 727
Beaulieu, S., Gilmore, G., Elson, R. A. W., Johnson, R. A., Santiago, B., Sigurdsson, S., & Tanvir,
N., 2001, astro-ph/0102312
Bellazzini, M., Ferraro, F., & Pancino, E. 2001, ApJ, in press (astro-ph/0104114)
Bica, E., Barbuy, B., & Ortolani, S. 1991, ApJ, 382, L15
Burstein, D., & Heiles, C. 1982, AJ, 87, 1165
Bullock, J. 1999, PhD Thesis, University of California, Santa Cruz
Caputo, F., Castellani, V., Chieffi, A., Pulone, L., & Tornamb´e, A. 1989, ApJ, 340, 241
Castellani, V., Chieffi, A., & Pulone, L. 1991, ApJS, 76, 911
Chaboyer, B., Green, E. M., & Liebert, J. 1999, AJ, 117, 1360
Cole, A. 1998, ApJ, 500, L137
Crowl, H. H., Sarajedini, A., Piatti, A., Geisler, D., Bica, E., Clar´a, J. J., & Santos, J. F. C. 2001,
AJ, 122, 220
Cuillandre, J. -C., Lequeux, J., Allen, R. J., Mellier, Y., & Bertin. E. 2001, ApJ, in press (astro-
ph/0102350)
Da Costa, G. S., & Armandroff, T. E. 1990, AJ, 100, 162
Da Costa, G. S., Armandroff, T. E., Caldwell, N., & Seitzer, P. 2000, AJ, 119, 705
Davidge, T. J. 1993, ApJ, 409, 190
Durrell, P. R., Harris, W. E., & Pritchet, C. J. 1994, AJ, 108, 2114
Durrell, P. R., Harris, W. E., & Pritchet, C. J. 2001, astro-ph/0101436
Ferraro, F. 1992, Mem. Soc. Astron. Italiana, 63, 491
Ferraro, F. R., Messineo,M., FusiPecci,F., dePalo,M.A., Straniero,O., Chieffi,A., & Limongi,M.
1999, AJ, 118, 1738
Fusi Pecci, F., Ferraro,F.R., Crocker,D.A., Rood,R.T., & Buonanno,R. 1990, A&A, 238, 95
Fusi Pecci, F., Bellazzini, M., Cacciari, C., & Ferraro, F. R. 1995, AJ, 110, 1664
Fusi Pecci,F., Buonanno,R., Cacciari,C., Corsi,C.E., Djorgovski,S.G., Federici,L., Ferraro,F.R.,
Parmeggiani,G., & Rich,R.M. 1996, AJ, 112, 1461
-- 14 --
Gallart, C. 1998, ApJ, 495, L43
Garnavich, P. M., VandenBerg, D. A., Zurek, D. R., & Hesser, J. E. 1994, AJ, 107, 1097
Girardi, L., & Salaris, M. 2001, MNRAS, in press
Girardi, L., Bressan, A., Bertelli, G., & Chiosi, C., 2000, A&AS, 141, 371
Gould, A. 1994, ApJ, 435, 573
Grebel, E. K. 2000, in A New Era of Microlensing Astrophysics, ASP Conf. Ser., edited by J. W.
Menzies & P. D. Sackett, in press (astro-ph/0008249)
Guarnieri, M. D., Ortolani, S., Montegriffo, P., Renzini, A., Barbuy, B., Bica, E., & Moneti, A.
1998, A&A, 331, 70
Harris, W. E. 1996, http://physun.mcmaster.ca/∼harris/mwgc.dat
Hodge, P., & Lee, M. -G. 1988, ApJ, 329, 651
Hodge, P., Lee, M. -G., & Mateo, M. 1988, ApJ, 324, 172
Holland, S., Fahlman, G., & Richer, H. 1996, AJ, 112, 1035
Holland, S., Ct, P., & Hesser, J. E. 1999, A&A, 348, 418
Jablonka,P., Bridges, T., Sarajedini, A., Meylan,G., Maeder, A., & Meynet, G. 1999, ApJ, 518, 627
Jablonka,P., Courbin,F., Meylan,G., Sarajedini,A., Bridges,T.J., & Magain,P. 2000, A&A, 359, 131
Johnson, J., & Bolte, M. 1998, AJ, 115, 693
Layden, A. C., & Sarajedini, A. 1997, ApJ, 486, L107
Lee, Y. -W., Demarque, P., & Zinn, R. J. 1990, ApJ, 350, 155
Lee, Y. -W., Demarque, P., & Zinn, R. J. 1994, ApJ, 423, 248
Massey,P., Armandroff,T.E., & Conti,P.S. 1986, AJ, 92, 1303
Morris, P. W., Reid, I. N., Griffiths, W. K., & Penny, A. J. 1994, MNRAS, 271, 852
Mould, J. R., & Kristian, J. 1986, ApJ, 305, 591
Ortolani, S., Barbuy, B., & Bica, E. 1990, A&A, 236, 362
Pritchet, C. J., & van den Bergh, S. 1988, ApJ, 331, 135
Renzini, A. 1999, in When and How do Bulges Form and Evolve?, edited by C. M. Carollo, H. C.
Ferguson, & R. F. G. Wyse (Cambridge:Cambridge University Press), p. 9
-- 15 --
Rich, R. M., Mighell, K., & Neill, J. D. 1996, in Formation of the Galactic Halo...Inside and Out,
ASP Conf. Ser. Vol. 92, edited by H. Morrison & A. Sarajedini (ASP:San Francisco) p. 544
Richer, H. B., Crabtree, D. R., & Pritchet, C. J. 1990, ApJ, 355, 448
Ryan, S. G., & Norris, J. E., 1991, AJ, 101, 1865
Sakai, S. Zaritsky, D., & Kennicutt, R. C., Jr. 2000, AJ, 119, 1197
Sarajedini, A. 1994, AJ, 107, 618
Sarajedini, A. 1999, AJ, 118, 2321
Sarajedini, A., & Forrester, W. L. 1995, AJ, 109, 1112
Sarajedini, A., Lee, Y. -W., & Lee, D. -H. 1995, ApJ, 450, 712
Sarajedini, A., Geisler, D., Schommer, R., & Harding, P. 2000, AJ, 120, 2437
Saviane, I., Rosenberg,A., Piotto,G., & Aparicio,A. 2000, A&A, 355, 966
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525
Silbermann, N. A. et al. 1996, ApJ, 470, 1
Silk, J. 2000, MNRAS, in press (astro-ph/0010624)
Stanek, K. Z., & Garnavich, P. M. 1998, ApJ, 503, 131
Stetson, P. B. 1994, PASP, 106, 250
van den Bergh, S. 1999, ARA&A, 9, 273
Walker, A. 1994, AJ, 108, 555
Williams, B. F., & Hodge, P. 2001, ApJ, 548, 190
Wyse, R., & Gilmore, G., 1995, AJ, 110, 2771
Zoccali, M., Renzini, A., Ortolani, S., Bica, E., & Barbuy, B. 2001, astro-ph/0101200
This preprint was prepared with the AAS LATEX macros v5.0.
-- 16 --
Table 1. Observing Log
Dataset
Filter
Exp.time (sec)
u2gv0401 F555W
u2gv0402 F555W
u2gv0403 F814W
u2gv0404 F814W
u2gv0405 F814W
u2gv0406 F814W
u2gv0407 F814W
1500
2300
2300
2300
2300
2300
1600
Table 2. Adopted Cluster Parameters
Cluster
[F e/H]
E(V − I) V(HB)a (m − M )I
47 Tuc
−0.71 ± 0.07
NGC 1851 −1.29 ± 0.07
M2
−1.58 ± 0.06
NGC 6397 −1.91 ± 0.14
M15
−2.17 ± 0.07
NGC 6553 −0.28 ± 0.15
NGC 6791 +0.28 ± 0.18
0.05
0.03
0.03
0.23
0.13
0.99
0.16
14.06
16.05
16.05
12.90
15.86
16.88
14.55
13.37
15.43
15.48
12.18
15.29
15.18
13.63
aAll values from Da Costa & Armandroff 1990 except for NGC
6553 (Guarnieri et al. 1998) and NGC 6791 (Sarajedini 1999).
-- 17 --
Fig. 1. -- (a) The location of our HST/WFPC2 field is indicated by the square in this 1.5 x 1.5
degree digitized sky survey image of M31. North is up and east is to the left. (b) Close-up view
of the 'G272' field with the WFPC2 footprint overlaid. (c) The mosaicked WFPC2 image with
the 'compass' markings showing the direction of north (arrow) and east. The black lines show the
three radial regions into which the photometry has been divided; region 1, which is closest to the
M31 nucleus, is on the right hand side.
Fig. 2. -- (a) The left panel shows the (V,V -- I) color-magnitude diagram (CMD) of the PC1; the
right panel shows the (I,V -- I) CMD. (b) Same as (a) except that the WF2 CMDs are shown. (c)
Same as (a) except that the WF4 CMDs are shown.
Fig. 3. -- (a) The left panel shows the distance and reddening corrected [MV , (V − I)o] color-
magnitude diagram (CMD) for Region 1 assuming (m − M )0 = 24.5 and E(B − V ) = 0.08; the
right panel shows the [MI , (V − I)o] CMD. The solid lines are the red giant branches of M15, 47
Tuc, NGC 6553, and NGC 6791. (b) Same as (a) except that the Region 2 CMDs are shown. (c)
Same as (a) except that the Region 3 CMDs are shown.
Fig. 4. -- The left panel shows the distance and reddening corrected [MV , (V −I)o] color-magnitude
diagram (CMD) for Region 1; the right panel shows the [MI , (V − I)o] CMD. The solid lines are the
red giant and horizontal branches of M68 and the dashed lines are those of M5. This comparison
is designed to illustrate the presence of a Pop II horizontal branch in our M31 field.
(a) The solid sequence
Fig. 5. -- The results of the artificial star experiments are displayed.
represents the red giant branch of artificial stars placed on the images while the points show the
locations of stars after they have been measured using our photometric technique. (b) The open
circles illustrate the difference between the measured and actual V -- I colors [∆(V − I)] of the
artificial stars as a function of I magnitude.
(c) For all artificial stars in the magnitude range
22.55 ≤ I ≤ 22.75, we show the histogram of ∆(V − I) values with the fitted Gaussian distribution
as the solid line. The standard deviation of this Gaussian is 0.049 mag, which we adopt as the
photometric error in this magnitude range.
Fig. 6. -- The lower panel shows the distance and reddening corrected CMD of our M31 field along
with the red giant branches of M15 (left-most) and NGC 6791 shown as the solid lines. The two
horizontal dashed lines indicate the magnitude range over which the color histogram in the upper
panel has been constructed. The solid line in the upper panel is a Gaussian with σobs = 0.182 mag
fitted to this histogram. Given that the photometric error in this magnitude range is σerr = 0.049
mag as yielded by the artificial stars, this implies an intrinsic color dispersion of σint = 0.175 mag
represented by the dashed Gaussian in the upper panel.
Fig. 7. -- The lower panel shows our relation between the dereddened RGB color at MI = −2
[(V − I)o,−2] and metal abundance. The filled circles are the star clusters listed in Table 2 while
the dashed curve is the weighted least squares fit shown in Equation 1. In light of the photometric
error yielded by the artificial stars of σerr = 0.049 mag, the upper panel displays the metallicity
-- 18 --
uncertainty resulting purely from the photometric errors.
Fig. 8. -- Using the dereddened colors of all stars with magnitudes in the range 22.55 ≤ I ≤ 22.75
(−1.9 ≤ MI ≤ −2.1) along with Equation 1, we contruct the metallicity distribution function
shown in the upper panel by the filled symbols. The solid line is the generalized histogram of these
abundances, which takes into account the metallicity error of each star. The lower panel shows the
variation of the MDF as the adopted reddening is varied from E(B -- V) = 0.06 (dashed line) to 0.08
(solid line) to 0.10 (dotted line).
Fig. 9. -- (a) A comparison of the 'G272' field metallicity distribution function (MDF) and that
of the G302 field from Holland et al. (1996) located at a projected distance of 32 arcmin from
the center of M31. (b) Same as (a) except that the MDF of the M31 halo field from Durrell et
al. (1994) located at 40 arcmin from the nucleus is shown. (c) Same as (a) except that the MDF
of the M31 halo field located at 90 arcmin from the nucleus and studied by Durrell et al. (2001)
is shown. All of these distributions have been scaled to have the same area. (d) The MDFs have
been scaled to the match the metal-poor tail of the 'G272' MDF.
Fig. 10. -- Gaussian fits to our metallicity distribution function after adopting the Durrell et al.
(2001) parameters for the metal-poor and intermediate metallicity halo components. The upper
and lower panels show the fits to the binned (filled circles) and generalized (solid line) histograms,
respectively. The dashed lines are the individual Gaussian components while the dotted line is the
sum of all three.
Fig. 11. -- (a) A comparison of the G272 field metallicity distribution function (MDF) with that
of Milky Way thick disk stars from Wyse & Gilmore (1995). (b) Same as (a) except that the MDF
of Milky Way field halo stars from Ryan & Norris (1991) is shown. (c) Same as (a) except that the
MDF of Milky Way Globular clusters constructed from the database of Harris (1996) is shown. (d)
Same as (a) except that the MDF of M31 globular clusters from Barmby et al. (2000) is shown.
All of these distributions have been scaled to have the same area.
Fig. 12. -- (a) The left panel shows the distance and reddening corrected [MV , (V − I)o] color-
magnitude diagram (CMD) for Region 1; the right panel shows the [MI , (V − I)o] CMD. The solid
lines are the Girardi et al. (2000) solar abundance (Z=0.019) theoretical isochrones for ages of 108,
6.3 × 108, 109, and 1.6 × 109 years. (b) Same as (a) except that the Region 2 CMDs are shown. (c)
Same as (a) except that the Region 3 CMDs are shown.
This figure "fig1a.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0107344v1
This figure "fig1b.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0107344v1
This figure "fig1c.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0107344v1
|
astro-ph/9806205 | 1 | 9806 | 1998-06-15T17:45:23 | Neutrino transport: no asymmetry in equilibrium | [
"astro-ph",
"hep-ph",
"hep-th"
] | A small asymmetry in the flux of neutrinos emitted by a hot newly-born neutron star could explain the observed motions of pulsars. However, even in the presence of parity-violating processes with anisotropic scattering amplitudes, no asymmetry is generated in thermal equilibrium. We explain why this no-go theorem stymies some of the proposed explanations for the pulsar ``kick'' velocities. | astro-ph | astro-ph | CERN-TH/98-192
astro-ph/9806205
8
9
9
1
n
u
J
5
1
1
v
5
0
2
6
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Neutrino transport: no asymmetry in equilibrium
Alexander Kusenko,1∗ Gino Segr`e,2† and Alexander Vilenkin3‡
1Theory Division, CERN, CH-1211 Geneva 23, Switzerland
2Department of Physics and Astronomy, University of Pennsylvania
Philadelphia, PA 19104-6396, USA
3Institute of Cosmology, Department of Physics and Astronomy
Tufts University, Medford, MA 02155, USA
Abstract
A small asymmetry in the flux of neutrinos emitted by a hot newly-born neutron
star could explain the observed motions of pulsars. However, even in the presence
of parity-violating processes with anisotropic scattering amplitudes, no asymmetry is
generated in thermal equilibrium. We explain why this no-go theorem stymies some
of the proposed explanations for the pulsar "kick" velocities.
CERN-TH/98-192
May, 1998
∗ email address: [email protected]
†email address: [email protected]
‡email address: [email protected]
The observed rapid motions [1] of magnetized, rotating neutron stars, pulsars,
can have a natural explanation if the neutrino emission from a cooling newly-born
neutron star exhibits a small, ≈1% asymmetry. One necessary condition for such
an asymmetry -- a preferred direction -- is, clearly, satisfied. A strong magnetic field
of the neutron star breaks the spherical symmetry and can, at least in principle, be
responsible for some anisotropy in neutrino emission.
One might think that, since the production and scattering of neutrinos in dense,
hot nuclear matter is affected by the magnetic field B, some parity-violating pro-
cesses could produce an asymmetric flux of neutrinos in thermal equilibrium. This
conclusion, however, is erroneous.
In fact, no asymmetry can build up in thermal equilibrium even if the scattering
probabilities are anisotropic. This was pointed out by one of the authors in Ref. [2].
The argument is similar to that of Weinberg [3] with respect to the baryon asymmetry
of the Universe. Our purpose here is to reiterate the discussion of Ref. [2] and to
elaborate on this general argument in application to the neutrino emission by a cooling
neutron star. This is particularly timely because several recent papers [4, 5] have
reached an erroneous conclusion with respect to the size of the recoil velocity the
pulsars can receive from neutrino emission.
The Boltzman equation for neutrinos ν scattering off neutrons n (it is straightfor-
ward to include the electrons, which we omit for simplicity) can be written as [6]
q0
∂
∂t
f (ν)(q, t) = Xs,s
′
Z d3p
p0
d3p′
p′
0
d3q′
q′
0
[f (ν)(q′, t) f (n)
s
′ (p′, t) W (p′, s′, q′p, s, q)
−f (ν)(q, t) f (n)
s
(p, t) W (p, s, qp′, s′, q′)].
(1)
Here f (ν)(q) and f (n)(p) are the neutrino and neutron distribution functions, respec-
tively; s denotes a neutron spin; and W (p′, s′, q′p, s, q) is the probability of scattering
n(p, s)ν(q)i → n(p′, s′)ν(q′)i per unit time per unit phase-space volume. The states
are normalized for the invariant phase-space volume d3p/p0 that appears in the in-
tegral. Here we neglected the effects of fermion degeneracy, the inclusion of which is
straightforward [3].
1
We will assume an arbitrary form for the scattering probability W , which may, in
particular, be anisotropic. The essential property of W is unitarity,
Z d3p′
p′
0
d3q′
q′
0
Xs
′
W (p, s, qp′, s′, q′) = Xs
′
Z d3p′
p′
0
d3q′
q′
0
W (p′, s′, q′p, s, q) = 1,
(2)
which is merely a requirement that the probability is conserved: with probability one,
every initial state scatters into one of the final states, and vice versa.
The unitarity of the scattering matrix, as expressed by equation (2), is sufficient
to show that the isotropic time-independent equilibrium distributions, which depend
only on energy, satisfy the Boltzman equation (1), regardless of any asymmetries in
W . In statistical equilibrium, f (ν)(q) ∝ exp{−q0/T } and f (n)
s
(p) ∝ exp{−Es(p)},
where Es(p) = p2/2m+gnµB s·B. From energy conservation, q0+Es(p) = q′
0+Es
′(p′),
so that
f (ν)(q)f (n)
s
(p) = f (ν)(q′)f (n)
′ (p′).
s
(3)
Relation (3) is quite general and can be regarded as an expression of detailed balance
in statistical equilibrium [6].
One can now prove that a time-independent spherically-symmetric distribution
function, which satisfies the condition of statistical equilibrium (3), is a solution of
the Boltzman equation (1) for any form of W . Indeed, if f (ν) is independent of time,
the left-hand side of equation (1) vanishes. The right-hand side,
Z d3p
p0
d3p′
p′
0
d3q′
q′
0
Xs,s
′
also vanishes because
f (ν)(q) f (n)
s
(p) [W (p′, s′, q′p, s, q) − W (p, s, qp′, s′, q′)],
(4)
Z d3p′
p′
0
d3q′
q′
0
Xs
′
[W (p′, s′, q′p, s, q) − W (p, s, qp′, s′, q′)] = 0
(5)
by virtue of equation (2). So, if the neutrinos are thermally produced in equilibrium,
no asymmetry is generated by an anisotropic scattering probability.
It is clear why the analysis of Ref. [4] is in contradiction with the no-go theorem we
have proven. The expression for the scattering cross-section in equation (3) of Ref. [4]
2
explicitly violates unitarity because it depends asymmetrically on the scattering angle
of the outgoing neutrino and has no dependency on the initial momentum of the
incident neutrino. In our notation, this corresponds to W ∝ (q0 + hsi · q′), where
hsi is the average neutron polarization. Since this expression does not satisfy the
constraints of unitarity, it can lead to a neutrino asymmetry and hence a pulsar
kick. A similar mistake was made in Ref. [5], where it was also claimed that the
asymmetry in the distribution of outgoing neutrinos is proportional to the optical
depth of the neutron star, that is to the size of the region where neutrinos are in
thermal equilibrium with nuclear matter. It should be clear from our discussion, as
well as from the general principles of thermodynamics, that the size of a system in
equilibrium does not affect the distribution functions.
Of course, a hot neutron star that emerges from a supernova explosion is not fully
in equilibrium, and this causes some asymmetry in the flux of outgoing neutrinos [2].
The departure from thermal equilibrium is due to the variation of the macroscopic
parameters, such as B, temperature, and the matter density, which occur on some
length scale L. Therefore, the asymmetry resulting from the non-equilibrium behavior
is proportional to λ/L, where λ is the neutrino mean free path.
In addition, an asymmetry in neutrino emission can arise from some other inter-
actions of neutrinos at or near the neutrinosphere, as suggested in Refs. [7]. Since at
that point the neutrinos are free-streaming, they are already out of equilibrium and
the above no-go theorem doesn't apply.
To summarize, any attempt to explain the pulsar kick velocities by a "cumulative"
amplification of a small asymmetry in the course of numerous collisions [4, 5] is
doomed to failure. Multiple collisions imply that the thermally-produced neutrinos
continue to be in statistical equilibrium. Therefore, no asymmetry can build up even
if the scattering probabilities are anisotropic.
3
References
[1] See, e. g., A. G. Lyne and D. R. Lorimer, Nature 369 (1994) 127; J. M. Cordes
and D. F. Chernoff, astro-ph/9707308; B. M. S. Hansen and E. S. Phinney, astro-
ph/9708071; C. Fryer, A. Burrows, and W. Benz, Astrophys. J. 496 (1998) 333.
[2] A. Vilenkin, Astrophys. J. 451 (1995) 700.
[3] S. Weinberg, Phys. Rev. Lett 42 (1979) 850.
[4] C. J. Horowitz and G. Li, Phys. Rev. Lett. 80 (1998) 3694.
[5] D. Lai and Y.-Z. Qian, Astrophys. J. 495 (1998) L103; astro-ph/9802344; astro-
ph/9802345.
[6] E. M. Lifshitz and L. P. Pitaevskii, Physical kinetics (L. D. Landau and E. M. Lif-
shitz, Course of theoretical physics; vol. 10), Pergamon Press, New York, 1981.
[7] M. B. Voloshin, Phys. Lett. B 209 (1988) 360; A. Kusenko and G. Segr`e, Phys.
Rev. Lett. 77 (1996) 4872; Phys. Lett. B 396, (1997) 197; E. Kh. Akhme-
dov, A. Lanza, and D. W. Sciama, Phys. Rev. D 56 (1997) 6117; C. W. Kim,
J. D. Kim, and J. Song, Phys. Lett. B 419 (1998) 279; D. Grasso, FTUV-98-
13, astro-ph/9802060; D. Grasso, H. Nunokawa, J. W. F. Valle, FTUV/98-24,
IFIC/98-24, astro-ph/9803002.
4
|
astro-ph/0702215 | 1 | 0702 | 2007-02-07T22:43:52 | Solar-like oscillations in open cluster stars | [
"astro-ph"
] | Asteroseismology of stellar clusters is potentially a powerful tool. The assumption of a common age, distance, and chemical composition provides constraints on each cluster member, which significantly improves the asteroseismic output. Driven by this great potential, we carried out multi-site observations aimed at detecting solar-like oscillations in the red giant stars in the open cluster M67 (NGC 2682) (Stello et al. 2006). Here we present the first analysis of our data, which show evidence of excess power in the Fourier spectra, shifting to lower frequencies for more luminous stars, consistent with expectations from oscillations. If the observed power excesses were due to stellar oscillations, this result would show great prospects for asteroseismology in stellar clusters. | astro-ph | astro-ph |
Comm. in Asteroseismology
Vol. 150, 2007
Solar-like oscillations in open cluster stars
D. Stello,1 H. Bruntt,1 T. Arentoft,2,3 R. L. Gilliland,4 J. Nuspl,5 S.-L. Kim,6 Y. B. Kang,6
J.-R. Koo,6 J.-A. Lee,6 C.-U. Lee,6 C. Sterken,7 A. P. Jacob,1 S. Frandsen,2,3 Z. E. Dind,1
H. R. Jensen,2 R. Szab´o,5 Z. Csubry,5 L. L. Kiss,1 M. Y. Bouzid,7 T. H. Dall,8
T. R. Bedding,1 H. Kjeldsen 2,3
1 School of Physics, University of Sydney, NSW 2006, Australia
2 Institut for Fysik og Astronomi (IFA), Aarhus Universitet, 8000 Aarhus, Denmark
3 Danish AsteroSeismology Centre, Aarhus Universitet, 8000 Aarhus, Denmark
4 Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, USA
5 Konkoly Observatory, 1525 Budapest, PO Box 67, Hungary
6 Korea Astronomy and Space Science Institute, Daejeon 305-348, Korea
7 Vrije Universiteit Brussel, Pleinlaan 2, B-1050 Brussels, Belgium
8 European Southern Observatory, Casilla 19001, Santiago 19, Chile
Introduction
Asteroseismology of stellar clusters is potentially a powerful tool. The assumption of a com-
mon age, distance, and chemical composition provides constraints on each cluster member,
which significantly improves the asteroseismic output. Hence, detecting oscillations in cluster
stars in a range of evolutionary stages holds promise of providing more stringent tests of
stellar evolution theory. Driven by this great potential, we carried out multi-site observations
aimed at detecting solar-like oscillations in the red giant stars in the open cluster M67 (NGC
2682). To obtain these data we observed for 43 days with nine 0.6-m to 2.1-m class telescopes
in January and February 2004 (Stello et al. 2006). The photometric time series comprises
roughly 18000 data points. The properties of the red giant stars we present here are given in
Table 1 and their location in the colour-magnitude diagram is shown in Fig. 1 (left panel).
Figure 1: Left Panel: The colour-magnitude diagram of the open cluster M67. The red giant target stars
and their ID are indicated. Right Panel: Average power distributions for three groups of stars sorted
according to luminosity: most luminous (clump stars), intermediate, and least luminous (lower RGB).
Arrows show expected locations of solar-like oscillations (see Table 1). Only stars with a white-noise level
lower than 50 µmag have been used.
2
Solar-like oscillations in open cluster stars
Table 1: Properties of red giant target stars. Luminosities and temperatures are from photometry (Mont-
gomery et al. 1993). Estimates of oscillation amplitudes, characteristic frequencies and large separations
are from known scaling relations in the literature (Kjeldsen & Bedding 1995, Brown et al. 1991), Cross
references are to Sanders (1977) and Gilliland et al. (1991).
ID L/L⊙
Teff
K
δL/L
µmag
νmax ∆ν0
µHz
µHz
Cross-ref.
8
9
10
2
18
5
17
7
15
14
13
50.8
50.2
48.2
46.4
45.8
25.4
22.4
20.2
16.9
11.2
9.9
4750
4772
4727
4727
4772
4815
4835
4854
4873
4945
4966
287
281
275
265
256
140
122
109
91
58
51
35.8
36.8
37.0
38.4
40.3
74.8
86.0
96.9
117.3
187.3
213.2
4.3
4.4
4.4
4.6
4.7
7.6
8.4
9.2
10.6
15.2
16.8
S1010/G2
S1084/ --
S1279/G7
S1074/ --
S1316/ --
S1054/G9
S1288/ --
S989/G12
S1277/ --
S1293/ --
S1305/ --
Results
Mean levels in the Fourier spectra in the frequency interval 300 -- 900 µHz, corresponding
to white noise, reach 20 µmag for the stars with the lowest noise.
In many stars we see
apparently high levels of non-white noise, but the detailed temporal variation of the noise
is unknown. We are therefore not able to clearly disentangle the noise and stellar signal in
the analysis. However, we do see evidence of excess power in the Fourier spectra, shifting to
lower frequencies for more luminous stars, consistent with expectations (Fig. 1; right panel).
If the observed power excesses were due to stellar oscillations, this result would show great
prospects for asteroseismology in stellar clusters. A more detailed analysis will be given by
Stello et al. (2007).
This paper has been supported by the Astronomical Society of
Acknowledgments.
Australia.
References
Brown T. M., Gilliland R. L., Noyes R. W., Ramsey L. W., 1991, ApJ 368, 599
Gilliland R. L., Brown T. M., Thomson D. T., Schild R. E., Jeffrey W. A., Penprase B. E., 1991, AJ 101,
541
Kjeldsen H., Bedding T. R., 1995, A&A 293, 87
Montgomery K. A., Marschall L. A., Janes K. A., 1993, AJ 106, 181
Sanders W. L., 1977, A&AS 27, 89
Stello D., et al., 2006, MNRAS, 373, 1141
Stello D., et al., 2007, MNRAS, accepted
|
astro-ph/0105183 | 2 | 0105 | 2002-02-07T03:06:03 | Two-Body Relaxation in Cosmological Simulations | [
"astro-ph"
] | It is logically possible that early two-body relaxation in simulations of cosmological clustering influences the final structure of massive clusters. Convergence studies in which mass and spatial resolution are simultaneously increased, cannot eliminate this possibility. We test the importance of two-body relaxation in cosmological simulations with simulations in which there are two species of particles. The cases of two mass ratios, sqrt(2):1 and 4:1, are investigated. Simulations are run with both a spatially fixed softening length and adaptive softening using the publicly available codes GADGET and MLAPM, respectively.
The effects of two-body relaxation are detected in both the density profiles of halos and the mass function of halos. The effects are more pronounced with a fixed softening length, but even in this case they are not so large as to suggest that results obtained with one mass species are significantly affected by two-body relaxation.
The simulations that use adaptive softening are less affected by two-body relaxation and produce slightly higher central densities in the largest halos. They run about three times faster than the simulations that use a fixed softening length. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 5 November 2018
(MN LATEX style file v1.4)
Two-Body Relaxation in Cosmological Simulations
James Binney1 and Alexander Knebe1,2,
1Theoretical Physics, 1 Keble Road, Oxford OX1 3NP
2Centre for Astrophysics & Supercomputing, Swinburne University, Mail # 31, PO Box 218, Hawthorn, VIC 3122, Australia
Received ...; accepted ...
ABSTRACT
It is logically possible that early two-body relaxation in simulations of cosmological
clustering influences the final structure of massive clusters. Convergence studies in
which mass and spatial resolution are simultaneously increased, cannot eliminate this
possibility. We test the importance of two-body relaxation in cosmological simulations
with simulations in which there are two species of particles. The cases of two mass
ratios, √2 : 1 and 4 : 1, are investigated. Simulations are run with both a spatially fixed
softening length and adaptive softening using the publicly available codes GADGET
and MLAPM, respectively.
The effects of two-body relaxation are detected in both the density profiles of
halos and the mass function of halos. The effects are more pronounced with a fixed
softening length, but even in this case they are not so large as to suggest that results
obtained with one mass species are significantly affected by two-body relaxation.
The simulations that use adaptive softening are less affected by two-body relax-
ation and produce slightly higher central densities in the largest halos. They run about
three times faster than the simulations that use a fixed softening length.
Key words: methods: numerical -- galaxies: formation -- cosmology: theory
1
INTRODUCTION
The Cold Dark Matter (CDM) model for the formation of
cosmological structure has enjoyed considerable success over
the last decade and a half. Recent results from the PSCz sur-
vey (Hamilton & Tegmark 2002) underline the ability of this
model to account successfully for observations of the distri-
bution of galaxies and the structure of rich galaxy clusters
within constraints set by measurements of the cosmic back-
ground radiation, the theory of primordial nucleosynthesis
and observations of distant type Ia supernovae.
By contrast with these successes where large-scale phe-
nomena are concerned, there are serious doubts as to
whether the CDM model is compatible with the internal
structures of galaxies (e.g., Spergel & Steinhardt 2000). All
these difficulties can be traced to the cuspy central density
profiles of dark matter halos. This phenomenon was first
identified by Navarro, Frenk & White (1996), who concluded
that the central density diverged with radius as r−1. Sub-
sequent work by Moore et al. (1998, 1999), Ghigna et al.
(2000) and Klypin et al. (2001) suggests that the divergence
is even steeper: ρ ∼ r−1.5.
On sufficiently small scales, discreteness effects must
cause deviations between density profiles in simulations and
those of real dark-matter halos, in which the particle mass
is believed to be tiny. Determining the smallest scale on
which numerical simulations are trustworthy has proved dif-
c(cid:13) 0000 RAS
ficult and has given rise to some controversy. Kravtsov et al.
(1998), Ghigna et al. (2000) and Klypin et al. (2001) have
approached this problem by progressively increasing the res-
olution of simulations started from initial conditions that
sampled the same underlying density field. Kravtsov et al.
(1998) argued that their simulations were converging on a
central density profile of halos that was less cuspy than the
NFW profile. Subsequently, Ghigna et al. (2000) and Klypin
et al. (2001) showed that simulations converge on profiles
more cuspy than the NFW profile if one simultaneously in-
creases both the mass and the spatial resolution; Kravtsov
et al. (1998) had increased only the spatial resolution.
While the results of Ghigna et al. (2000) and Klypin
et al. (2001) suggest that real dark matter halos should
have very cuspy centres as Moore et al. (1998) originally
concluded, they do not establish this claim beyond all rea-
sonable doubt. The reason is that when structure forms
bottom-up, as in the CDM paradigm, and mass and spa-
tial resolution are increased together, the first virialized sys-
tems are few-body systems, regardless of the resolution em-
ployed. In such systems the two-body time is comparable
to the dynamical time, and two-body interactions tend to
make the core denser, and less vulnerable to subsequent
tidal destruction. As clustering progresses, one can imag-
ine that the dynamics is dominated by the interaction of
quasi-particles formed by first-generation systems. The early
clustering of these quasi-particles is again heavily influenced
2
Binney J.J. & Knebe A.
by two-particle interactions, and leads to the formation of a
new generation of quasi-particles, each of which is made up
of quasi-particles of the previous generation. Thus one can
imagine the dynamics right up to the macroscopic scale be-
ing influenced by the two-body interactions that took place
between the real simulation particles as the density field first
went non-linear. Convergence tests of the type performed by
Ghigna et al. (2000) and Klypin et al. (2001) would be pow-
erless to expose discreteness effects if this picture were valid,
because a low-resolution simulation would simply enter the
evolutionary sequence of a high-resolution simulation at the
stage proper to the mass of its simulation particles.
An indication that we should worry about the possi-
bility just described comes from the observation that cuspy
profiles are only obtained when the smoothing kernel used
to calculate gravitational forces from particle positions is
hard. Now, Fig. 10 of Knebe, Green & Binney (2001) shows
that for the values of the softening parameter that yield
cuspy profiles, the net gravitational force on a particle may
be dominated by the force from its nearest neighbour prior
to the virialization of the smallest modelled scales. This cir-
cumstance is worrying, because the essence of collisionless
dynamics is that the forces acting on each particle reflect
the mean density field rather than the chance location of a
neighbour. Moreover, large forces from neighbours can affect
the dynamics of particles more strongly at early times, when
random velocities are small and thus force contributions are
relatively slowly changing.
The standard way of determining the significance for
a simulation of two-body relaxation is to include particles
of more than one mass: if the simulation is collisionless,
the final distributions of the particles will be independent
of mass, whereas the more massive particles will tend to
sink to the bottoms of potential wells if two-body relaxation
is significant. Since no results from multi-mass cosmologi-
cal simulations have appeared, we present here the results
of such experiments. The experiments were conducted with
two radically different simulation codes, the tree-code GAD-
GET (Springel, Yoshida & White 2001) and the adaptive
multigrid code, MLAPM (Knebe et al. 2001). We detect
the effects of two-body relaxation in results produced with
both codes. However, the magnitude of the effect is not so
large as to suggest that the cuspiness of dark-matter halos
is generated by two-body relaxation. We argue instead that
cuspiness arises from violent relaxation, and will invariably
arise when gravity causes particles to cluster collisionlessly
in a cosmological context.
2
INITIAL CONDITIONS
Cosmological simulations are invariably initialized by using
the Zel'dovich approximation to displace particles from an
initial approximately homogeneous distribution (e.g., Efs-
tathiou et al. 1985). The simplest way to approximate a
homogeneous distribution is to place particles at the ver-
tices of a regular grid. Since the regularity of the grid obvi-
ously induces unwanted long-range correlations between the
particles, another computationally more demanding way of
producing an approximately homogeneous mass distribution
has been widely used. This technique involves scattering par-
ticles at random over the computational volume, and then
Figure 1. Power spectra of the initial particle distributions.
moving them for some time under the influence of a ficti-
tious repulsive gravity. The motion of particles away from
one another flattens out Poissonian density inhomogeneities
of the initial distribution to produce a rather homogeneous
glass (White, 1996).
We have initialized simulations containing two parti-
cle species using both of the above approaches to the pro-
duction of an approximately homogeneous distribution. Un-
fortunately, glasses containing two particle species contain
complex short-range correlations because the size of the void
around each particle that is generated by repulsive gravity
depends on the mass of the particle. Consequently, multi-
mass glasses have complex and mass-dependent short-range
correlations that obscure the effects of two-body relaxation.
Since initial conditions based on a regular lattice yield much
cleaner results, we confine the discussion to these simula-
tions. The results obtained with glass-initialized simulations
are consistent with the results presented here.
Our lattice based simulations were initialized as follows.
643 particles of one species were placed on the nodes of
a regular grid with 643 cells covering a physical scale of
15h−1 Mpc on a side. Then 643 particles were placed on the
nodes of the grid that is offset from the first grid by half a
lattice spacing parallel to each axis. Finally Zel'dovich dis-
placements were applied to each mass. In the final configu-
ration, the particles of a given species define a density field
that differs only in its overall normalization from the den-
sity field defined by particles of the other species. Hence the
composite density field defined by both species together is
also the same, up to an overall normalization. The masses of
individual particles were set such that the composite density
was that of the standard CDM model and the masses of the
two species were in the ratio 1 : √2 or 1 : 4. It follows that
the more massive species accounted for 59 per cent and 80
per cent of the total mass in the two cases. Fig. 1 shows the
power spectra of the initial particle distribution. The power
spectra of the light (m1) and the heavy particles are indis-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Two-body relaxation in cosmological simulations
3
Figure 2. Density profiles of the two most massive halos in four simulations. Each column shows results for one of the halos. Full curves
show the number density of the less massive particles, and dashed curves show the number density of the more massive particles.
tinguishable except at the highest wavenumbers, where the
offset between the lattices from which the particles started
gives rise to very slight differences in the way in which arti-
facts arising from the lattices impact on the power spectrum.
3 RESULTS
We evolved the particle distributions that are described by
Fig. 1 in a standard CDM cosmology from redshift z = 35
to the current epoch using both the tree code GADGET
(Springel et al. 2001) and our own multigrid code, MLAPM
(Knebe et al. 2001). The essential difference between these
codes is that whereas GADGET uses a spatially invari-
ant gravitational softening length, MLAPM softens grav-
ity adaptively so as to provide a compromise between en-
hanced force resolution and reduced particle noise in the
force-field. The GADGET's Plummer softening length was
set to ǫ = min[5, 1/a(t)]h−1 kpc, where a < 1 is the scale
factor, so at late times the code used accurately Newtonian
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
forces for distances greater than 5h−1 kpc. MLAPM oper-
ates by refining the grid on which it solves Poisson's equation
whenever the density exceeds a critical value ρref , typically
set to 8 particles per cell. For these experiments the density
compared to ρref was not the density of gravitating mass, but
the number density of particles, irrespective of their mass.
The finest grid employed at the end of the simulation had a
mesh spacing equivalent to 81923 cells in the computational
volume, each cell being 1.8h−1 kpc on a side. On such a mesh
the force law becomes Newtonian for distances >
5h−1 kpc
∼
(Knebe et al., 2001). Hence the spatial resolution provided
by the two codes agreed well in high-density regions.
The GADGET runs required 244 398 and 249 677
timesteps, while the MLAPM runs used the equivalent of
1000 timesteps on the finest domain grid (128 cells on a
side), during which 64 000 timesteps would be taken on an
81923 grid.
Fig. 2 shows the number-density profiles of the less and
more massive particles of the two most massive halos to form
4
Binney J.J. & Knebe A.
Figure 4. Ratio of the number of light to heavy particles in halos
as a function of the total number of particles in the halo.
panel we show for each species the ratio of the actual number
of cluster to that predicted by Press-Schechter theory. The
full histograms, which show the results obtained for the two
mass ratios by MLAPM, are almost indistinguishable. The
results obtained with GADGET agree moderately well with
the MLAPM results only for m2/m1 = 4. The GADGET
results for m2/m1 = √2 show significantly more low-mass
clusters.
Fig. 4 clarifies what is going on by plotting as a func-
tion of the total particle number of a given halo the ratio
of the number of light to heavy particles. Since in the sim-
ulation as a whole there are equal numbers of particles of
each species, this ratio would ideally be unity, independent
of total particle number. The results from MLAPM devi-
ate significantly from this ideal only for npart < 10, where
for both mass ratios clusters have more light than heavy
particles. The results from GADGET, by contrast, reveal a
significant tendency for clusters containing > 10 particles
to have more heavy than light particles. The tendency is
strongest for m2/m1 = 4, as would be expected if it were
the result of two-particle relaxation causing heavy particles
to sink towards the centres of clusters, while light particles
evaporate from them. It is not clear why in all simulations
clusters with only a handful of particles tend to have an
excess of light particles, but this phenomenon may be con-
nected with the fact that in Fig. 1 the power spectrum of
the light particles is higher than that of the heavy particles
at large k. We believe this phenomenon to be of no physical
significance and to arise by chance from the phases of the
waves used to displace the particles.
4 DISCUSSION
Simulations of cosmological clustering with two mass species
have been used to explore the importance of two-body relax-
ation in general cosmological simulations. In the two largest
halos that form in our simulations, mass segregation is a
small but detectable effect in the sense that the difference
in the central densities of particles with masses in the ratio
4 : 1 is comparable to the difference in the central densities
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 3. Mass functions from the simulations. The friends-of-
friends algorithm has been used. The ordinate shows the number
of particles in each object, not the total mass. The lower panel
shows the ratio of the actual number of clusters to predicted by
Press-Schechter theory, and the thin full curve in the upper panel
shows this prediction.
in the simulations. As we will see below in Fig. 4, both these
halos contain about an equal number of light and heavy
particles in every run. For each halo four pairs of profiles
are shown, corresponding to the two mass ratios and two
codes. Since the full curves, which show the profile of the
less massive particles, always lie below the dashed curves,
mass segregation is clearly detected. The effect is slightly
more pronounced for the halo shown on the left than for
that shown on the right. The main difference between the
results obtained with the two codes is that MLAPM pro-
duces a central density that is about a factor 2 larger than
that produced by GADGET, and mass segregation is more
pronounced in the profiles produced by GADGET.
The effects of two-body relaxation on the less massive
halos in the simulations is examined by Fig. 3, which shows
for the four simulations the numbers of halos formed that
contain in excess of a given number of particles (irrespective
of particle mass). The thin full curve in the upper panel
shows the prediction of Press-Schechter theory. In the lower
Two-body relaxation in cosmological simulations
5
early on as part of a low-mass object. The insensitivity of
density profiles to the initial power spectrum must be the
result of a conspiracy, which ensures that as high-k power
is smoothed away, the loss of high-density seeds from early
virialization is counteracted by an increase in the effective-
ness of violent relaxation at late times. The bottom line is
that cuspy density profiles must be considered inevitable so
long as structure formation is dominated by a collisionless
fluid.
ACKNOWLEDGMENTS
We benefited from a valuable conversation with A. Klypin.
REFERENCES
Binney, J., & Tremaine, S. 1987, Galactic Dynamics (Princeton:
Princeton University Press)
Dehnen W., 2000, ApJ Lett. 536, 39
Doroshkevich, A.G., Klypin, A.A., 1981, Soviet Astr., 25, 127
Efstathiou G., Davis M., Frenck C.S., White S.D.M.,
1985, ApJ Suppl. 57, 241
Ghigna S., Moore B., Governato F., Lake G., Quinn T., Stadel
J., 2000, ApJ 544, 616
Hamilton A., Tegmark M.,
2002, MNRAS
in
press
astro-ph/0008392
Klypin A.A., Kravtsov A.V., Valenzuela O., Prada F.,
1999, ApJ 522, 82
Klypin A., Kravtsov A.V., Bullock J.S., Primack J.R.,
2001, ApJ 554, 903
Knebe A., Devriendt J.E.G., Mahmood A., Silk J., 2002, MNRAS,
329 , 813
Knebe A., Green A., Binney J.J., 2001, MNRAS 325, 845
Kravtsov A.V., Klypin A., Bullock J.S., Primack J.R.,
1998, ApJ 502, 48
Moore B., Governato F., Quinn T., Stadel J., Lake G.,
1998, ApJ 499, L5
Moore B., Quinn T., Governato F., Lake G., Stadel J., Lake G.,
1999, MNRAS 310, 1147
Navarro J.F., Frenk C.S., White S.D.M., 1996, ApJ, 462, 563
Spergel D.N., Steinhardt P.J., 2000, Phys. Rev. Lett. 84, 3760
Springel V., Yoshida N., White S.D.M., 2001, NewA 6, 79
van Albada, T.S., 1982, MNRAS 201, 939
White S.D.M., 1996, Cosmology and Large-Scale Structure, Les
Houches Session LX, eds. Schaeffer R., Silk J., Spiro M., Zinn-
Justin J., Elsevier 1996, p. 349
obtained for any one species with the two codes (about a
factor 2). Mass segregation is stronger in simulations run
with GADGET, which has (spatially) fixed softening, than
with MLAPM, which softens adaptively.
In the MLAPM simulations there is no evidence that
two-particle relaxation enhances the fraction of heavy par-
ticles in clusters. The GADGET simulations show a clear
tendency for clusters with more than a handful of particles
to have more massive than light particles. This tendency in-
creases in strength with the mass ratio m2/m1, just as is
expected if it is driven by two-particle relaxation.
Given the clear indication that two-particle relaxation
is more pronounced in the GADGET simulations, it is inter-
esting that it is in the MLAPM simulations that the central
densities of the two most massive clusters are largest (by
about a factor 2).
The major difference between the GADGET and
MLAPM simulations was one of computational cost: using
a small softening length everywhere increases the time re-
quired to run a simulation, both because it increases the
cost of evaluating forces, and, more importantly, because it
forces one to smaller timesteps. Quantitatively, the present
GADGET simulations took longer than the corresponding
MLAPM simulations by a factor about three. Since the qual-
ity of the scientific output from an N-body simulation is
always limited by the number of particles employed, and
most simulators are more limited by the availability of com-
puter time than computer memory, the different speeds of
our simulation amounts to a strong case for the use of adap-
tive softening of the type that MLAPM, and certain tree
codes (Dehnen 2000) provide.
The equipartition time, on which mass segregation de-
velops,
is faster than the two-body relaxation time by
roughly the ratio of the particle masses. By contrast, the
core-collapse timescale, on which the central density of a
halo is increased by two-body relaxation, is typically tens
to hundreds of two-body relaxation times (e.g., Binney &
Tremaine 1987). Hence, our finding that mass segregation
is a marginal effect for mass ratio 4 : 1, strongly suggests
that the cuspiness of dark-matter halos in simulations is not
an artifact produced by two-body relaxation. This conclu-
sion is reinforced by the observation that central densities
are highest in simulations in which mass segregation is least
evident.
If the cuspiness of halos is not caused by high-k power
in the input spectrum, as experiments with WDM suggests
(Knebe et al. 2002), and not produced by two-body relax-
ation as we have argued, what is its origin? It must result
from violent relaxation: in the absence of high-k power, the
quasi-linear regime will continue until massive objects are
ready to virialize. These will collapse by large factors be-
cause they will collapse from smooth and symmetrical initial
conditions. Numerical experiments long ago established that
the peak density in the final virialized entity is comparable
to the peak density achieved during collapse (Doroshkevich
& Klypin 1981; van Albada 1982). Hence, there is an ar-
gument for less high-k power giving rise to higher central
densities.
In the presence of high-k power, violent relaxation will
be unimportant for the formation of massive objects: these
will form hierarchically through a series of mergers, and their
cores are likely to be dominated by material that virialized
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0003242 | 1 | 0003 | 2000-03-16T19:36:18 | Supernovae versus Neutron Star Mergers as the Major r-Process Sources | [
"astro-ph"
] | I show that recent observations of r-process abundances in metal-poor stars are difficult to explain if neutron star mergers (NSMs) are the major r-process sources. In contrast, such observations and meteoritic data on Hf182 and I129 in the early solar system support a self-consistent picture of r-process enrichment by supernovae (SNe). While further theoretical studies of r-process production and enrichment are needed for both SNe and NSMs, I emphasize two possible direct observational tests of the SN r-process model: gamma rays from decay of r-process nuclei in SN remnants and surface contamination of the companion by SN r-process ejecta in binaries. | astro-ph | astro-ph |
Revised version March 6, 2000
Preprint typeset using LATEX style emulateapj v. 04/03/99
SUPERNOVAE VERSUS NEUTRON STAR MERGERS AS THE MAJOR R-PROCESS SOURCES
School of Physics and Astronomy, University of Minnesota, Minneapolis, MN 55455; [email protected]
Y.-Z. Qian
Revised version March 6, 2000
ABSTRACT
I show that recent observations of r-process abundances in metal-poor stars are difficult to explain
if neutron star mergers (NSMs) are the major r-process sources.
In contrast, such observations and
meteoritic data on 182Hf and 129I in the early solar system support a self-consistent picture of r-process
enrichment by supernovae (SNe). While further theoretical studies of r-process production and enrich-
ment are needed for both SNe and NSMs, I emphasize two possible direct observational tests of the SN
r-process model: gamma rays from decay of r-process nuclei in SN remnants and surface contamination
of the companion by SN r-process ejecta in binaries.
Subject headings: Galaxy: evolution -- nuclear reactions, nucleosynthesis, abundances -- supernovae:
general -- stars: neutron
1.
INTRODUCTION
sources.
All of the actinides and approximately half of the stable
nuclei with mass numbers A ∼> 100 in the solar system
were produced by rapid neutron capture, the r-process.
While the astrophysical site of this process remains a
mystery, the main candidate environments are neutrino-
heated ejecta from core-collapse supernovae (SNe; Woosley
& Hoffman 1992; Meyer et al. 1992; Takahashi, Witti, &
Janka 1994; Woosley et al. 1994) and decompressed ejecta
from neutron star mergers (NSMs; Lattimer & Schramm
1974, 1976; Symbalisty & Schramm 1982; Freiburghaus,
Rosswog, & Thielemann 1999b). To be a major source
for the r-process, an environment must satisfy two crite-
rions: one on reproducing the solar r-process abundance
pattern and the other on supplying the total amount of r-
process material in the present Galaxy. Using more phys-
ical parametrizations than the previous approach based
on constant neutron number density and temperature
(e.g., Kratz et al. 1993), several recent studies (Hoffman,
Woosley, & Qian 1997; Meyer & Brown 1997; Freiburghaus
et al. 1999a) derived the r-process conditions that can pro-
duce relative abundance patterns with peaks at A ∼ 130
and 195 as observed in the solar system. At present, it
seems that conditions in models of neutrino-heated ejecta
from SNe are deficient (e.g., Qian & Woosley 1996) while
those in models of decompressed ejecta from NSMs are
promising (e.g., Freiburghaus et al.
1999b) for an r-
process. However, due to the uncertainties in the theoreti-
cal models (e.g., Qian & Woosley 1996; Freiburghaus et al.
1999b), a reliable comparison of the actual conditions in
these two environments with the r-process conditions can-
not be made yet to help establish or discriminate either
environment as the r-process site. Furthermore, for both
environments the models showed that if the ejecta were
composed of r-process material, then the amount provided
by a single event combined with the estimated number of
SNe or NSMs over Galactic history would be adequate to
account for the present Galactic r-process inventory (e.g.,
Mathews & Cowan 1990; Qian & Woosley 1996; Rosswog
et al. 1999). Therefore, the above two criterions can-
not readily identify SNe or NSMs as the major r-process
1
In this Letter I discuss a phenomenological approach to
test SNe and NSMs as the major r-process sources. By
considering mixing of the ejecta from an individual event
with the interstellar medium (ISM), I show that observa-
tions of metal-poor stars are difficult to explain if NSMs
are the major r-process sources (§2). I further show that
a self-consistent picture of r-process enrichment by SNe
based on the same consideration is supported by mete-
oritic data on 182Hf and 129I in the early solar system and
by observations of metal-poor stars (§3). To emphasize
the importance of observations in establishing the major
r-process sources, I conclude with a discussion of two pos-
sible direct tests of the SN r-process model: gamma rays
from decay of r-process nuclei in SN remnants and surface
contamination of the companion by SN r-process ejecta in
binaries (§4).
2. R-PROCESS ABUNDANCES AND NEUTRON STAR
MERGERS
G
I assume that as far as the r-process is concerned, the
mergering of a neutron star (NS) with a black hole (BH)
is similar to that of two neutron stars. These two kinds
of events are collectively referred to as NSMs. The rate
of such events in the Galaxy is quite uncertain. Here I
take an average rate hf NSM
i ∼ (104 yr)−1 over Galac-
tic history, which is at the very high end of various esti-
mates (e.g., Phinney 1991; Bethe & Brown 1998; Arzou-
manian, Cordes, & Wasserman 1999). Numerical simula-
tions of a NS-NS merger event by Rosswog et al. (1999)
showed that a total mass M NSM
∼ 4 × (10−3 -- 10−2)M⊙
of decompressed NS material may be ejected. Then the
grand total from all the past NSMs over Galactic history
of tG ∼ 1010 yr is M NSM
itG ∼ 4 × (103 -- 104)M⊙.
This is roughly equal to the present Galactic r-process in-
ventory X tot
⊙,r ∼ 10−7 is the
total r-process mass fraction of nuclei with A ∼> 100 in
the solar system (Kappeler, Beer, & Wisshak 1989) and
MG ∼ 1011 M⊙ is the total Galactic mass in stars and
gas. So it seems that NSMs could be the major r-process
sources.
⊙,rMG ∼ 104 M⊙, where X tot
ej
ej
hf NSM
G
2
However, the amount of ejecta from a single NSM event
discussed above is too much to explain the observed r-
process abundances in metal-poor stars. This can be seen
by considering mixing of the ejecta from each event with
the ISM. Rosswog et al. (1999) showed that the total en-
ergy of the NSM ejecta is at most comparable to the SN
explosion energy (∼ 1051 erg). Therefore, when its en-
ergy/momentum is dispersed in the ISM, this ejecta can
mix with at most the same amount of material as swept
up by a SN remnant, Mmix ≈ 3 × 104 M⊙ (e.g., Thorn-
ton et al. 1998). Consequently, if all of this ejecta were
r-process material as required to account for the present
Galactic r-process inventory, an ISM enriched by a single
NSM event would have a total r-process mass fraction
⊙,r
ej
X tot
r,NSM = 1.3 × 10−7 M NSM
4 × 10−3 M⊙!(cid:18) 3 × 104 M⊙
Mmix
ej
(cid:19) .
⊙,r
+ X A>130
⊙,r
⊙,r
≈ X A>130
100∼<A≤130
even for the lowest M NSM
(1)
The solar r-process mass fractions of elements with 100 ∼<
A ≤ 130 and A > 130 are X
≈
4 × 10−8 (Kappeler et al. 1989). According to equa-
tion (1), X tot
⊙,r ≈
100∼<A≤130
r,NSM would be approximately equal to X tot
X
of inter-
est. Therefore, whether the NSM ejecta were composed
of r-process elements with 100 ∼< A ≤ 130 or A > 130,
or both groups in solar proportion, equation (1) predicts
that abundance ratios of e.g., Ag (A ∼ 107) and/or Eu
(A ∼ 153) with respect to hydrogen in an ISM enriched
by a single event would be at least comparable to the corre-
sponding solar r-process values (Ag/H)⊙,r and (Eu/H)⊙,r.
This predicted level of r-process enrichment is in disagree-
ment with recent observations of r-process abundances in
metal-poor stars, as Ag/H in stars with [Fe/H] ∼ −1.3 to
−2.2 are ∼ 10 -- 102 times lower than (Ag/H)⊙,r (Crawford
et al. 1998) while Eu/H in stars with [Fe/H] ∼ −3 are
∼ 30 -- 103 lower than (Eu/H)⊙,r (McWilliam et al. 1995;
Sneden et al. 1996, 1998).
The following interpretation of equation (1) gives some
insights into how NSMs would explain the present Galac-
tic r-process inventory and may help appreciate why this
explanation is disfavored by observations. The Galaxy
can be divided into ∼ 3 × 106 units each having a mass
Mmix ≈ 3 × 104 M⊙. This division has a physical mean-
ing as Mmix is the maximum mass within which the ejecta
from an individual NSM event can be distributed. Due
to the rather low rate of NSMs in the Galaxy, on average
at most one NSM event occurred in a unit over Galac-
tic history. Therefore, in order to account for the present
Galactic r-process inventory, each unit would have to be
enriched with an approximately solar r-process mass frac-
tion by a single NSM event. As shown above, this picture
of r-process enrichment is inconsistent with the observed
r-process abundances in metal-poor stars. Furthermore,
due to the high rate of SNe in the Galaxy, many SNe oc-
curred in a unit where a single NSM event also occurred
at sometime in Galactic history (see §4). As Fe enrich-
ment of this unit was provided by these SNe, stars formed
at different times in this unit would have varying [Fe/H]
but either zero or approximately solar r-process mass frac-
tion if NSMs were the major r-process sources. This is in
disagreement with the observed correlation between abun-
dances of r-process elements and Fe at [Fe/H] ∼> −2.5
(Gratton & Sneden 1994; Crawford et al. 1998; see also
McWilliam et al. 1995).
Of course, the above discussion of mixing of the NSM
ejecta and the ISM is oversimplified. For example, one
could imagine that a smaller than average fraction of r-
process ejecta was mixed into the ISM near the edge of a
NSM remnant. In this case stars formed near the edge of
the remnant would have r-process mass fractions smaller
than that given by equation (1). However, one would also
expect that less than average enrichment was not unduly
pervasive and a significant fraction of the metal-poor stars
would have the r-process enrichment indicated by equation
(1). The fact that no such stars have been observed sug-
gests a difficulty of the NSM r-process model in explaining
the observations at low metallicities. Furthermore, even
if the observed r-process abundances in metal-poor stars
could be attributed to pervasive less than average enrich-
ment by NSMs, one would still face the other difficulty in
explaining the observed correlation between abundances
of r-process elements and Fe at [Fe/H] ∼> −2.5. As Fe
enrichment was controlled by SNe occurring at a much
higher rate, widely-varying degrees of mixing of the r-
process ejecta in an already existing NSM remnant with
Fe produced by fresh SNe would result in large scatter in
r-process abundances over a broad range of [Fe/H]. This is
in disagreement with the rather early onset of the correla-
tion between abundances of r-process elements and Fe at
[Fe/H] ≈ −2.5. In summary, observations of metal-poor
stars would be difficult to explain if NSMs were the major
r-process sources.
3. R-PROCESS ENRICHMENT BY SUPERNOVAE
In contrast to the case of NSMs, observations of metal-
poor stars as well as meteoritic data on 182Hf and 129I in
the early solar system support a self-consistent picture of
r-process enrichment by SNe (Qian & Wasserburg 2000,
hereafter QW; Wasserburg & Qian 2000; hereafter WQ).
In this picture the ejecta from each SN event is mixed with
an average mass Mmix ≈ 3 × 104 M⊙ of ISM swept up by
It is as-
the SN remnant (e.g., Thornton et al. 1998).
sumed that the SN rate per unit mass of gas f SN
G /Mgas is
approximately constant over Galactic history (this seems
reasonable as the star formation rate is proportional to the
gas mass). Consequently, an average ISM in the Galaxy
is enriched by newly-synthesized material from SNe at a
frequency
f SN
mix = Mmix
f SN
G
Mgas
= (107 yr)−1
×(cid:18) Mmix
3 × 104 M⊙(cid:19)(cid:20)
f SN
G
(30 yr)−1(cid:21)(cid:18) 1010 M⊙
Mgas (cid:19) , (2)
where f SN
G /Mgas
present Galaxy.
is estimated using quantities for the
Meteoritic data on 182Hf (with a lifetime ¯τ182 = 1.3 ×
107 yr) and 129I (¯τ129 = 2.3 × 107 yr) in the early so-
lar system shed important light on the r-process and
its association with SNe. Wasserburg, Busso, & Gallino
(1996; see also QW) showed that the abundance ratio
(182Hf/180Hf)SSF = 2.4 × 10−4 (Harper & Jacobsen 1996;
Lee & Halliday 1995, 1998) is consistent with common
SNe injecting 182Hf into the ISM at a high frequency
fH ≈ f SN
mix ∼ (107 yr)−1 over a uniform production time
TUP ≈ 1010 yr before solar system formation (SSF). How-
ever, the abundance ratio (129I/127I)SSF = 10−4 (Reynolds
1960; see also Brazzle et al. 1999) must be accounted for
by a different type of SNe occurring at a low frequency
fL ∼ fH/10 ∼ (108 yr)−1 (Wasserburg et al. 1996; QW).
Therefore, the meteoritic data require at least two dis-
tinct types of SN r-process events: the high-frequency H
events producing heavy nuclei with A > 130 including
182Hf and the low-frequency L events producing light nu-
clei with A ≤ 130 including 129I. The r-process produc-
tion in the SN environments associated with the H and
L events was discussed in some detail by Qian, Vogel, &
Wasserburg (1998a).
The above picture of r-process production and enrich-
ment by SNe has clear predictions for r-process abun-
dances resulting from a single event (QW; WQ). For ex-
ample, with an average ISM enriched by the H events at
a frequency fH ∼ (107 yr)−1, the solar r-process abun-
dances of elements with A > 130 such as Eu were pro-
vided by fHTUP ∼ 103 H events. This requires that the
Eu abundance in an ISM enriched by a single H event
must be
log ǫH(Eu) = log ǫ⊙,r(Eu) − log(fHTUP) ∼ −2.5,
(3)
where the log ǫ notation is defined as log ǫ(Eu) ≡
log(Eu/H) + 12 with Eu/H being the abundance ratio of
Eu to hydrogen, and log ǫ⊙,r(Eu) = 0.51 is the value for
solar r-process Eu (Kappeler et al. 1989). Similarly, the
solar r-process abundances of elements with A ≤ 130 such
as Ag were provided by fLTUP ∼ 102 L events. This re-
quires that the Ag abundance in an ISM enriched by a
single L event must be
log ǫL(Ag) = log ǫ⊙,r(Ag) − log(fLTUP) ∼ −0.8,
(4)
where log ǫ⊙,r(Ag) = 1.19 (Kappeler et al. 1989) is used.
The predictions in equations (3) and (4) are in good agree-
ment with observations of very metal-poor stars which
were formed when only a small number of SNe had con-
tributed r-process elements to the ISM. The observed
log ǫ(Eu) values for stars with [Fe/H] ∼ −3 range from
−2.5 to −0.9 (McWilliam et al. 1995; Sneden et al. 1996,
1998), which can be accounted for by ∼ 1 -- 40 H events
with log ǫH(Eu) ∼ −2.5 from a single event (QW; WQ).
In addition, Ag abundances at the level indicated by equa-
tion (4) were observed in HD 2665 and BD +37◦1458 with
[Fe/H] ∼ −2 (Crawford et al. 1998) and in CS 22892 -- 052
with [Fe/H] ∼ −3 (Cowan et al. 1999).
⊙,r
100∼<A≤130
In the above picture of r-process enrichment by SNe the
total mass of r-process elements ejected in an H event must
be X A>130
⊙,r Mmix/(fHTUP) ∼ 10−6 M⊙, while that in an
Mmix/(fLTUP) ∼ 10−5 M⊙.
L event must be X
A total ∼ 10−6 -- 10−5 M⊙ of material can be naturally
provided by the neutrino-heated ejecta from the proto-
neutron star in a SN over a period ∼ 1 -- 10 s (e.g., Qian
& Woosley 1996). However, whether this neutrino-heated
ejecta is composed of r-process material is yet to be shown.
The difference by a factor ∼ 10 in the total amount of
r-process ejecta between the H and L events has been
speculated to indicate that neutrino emission and hence,
3
ejection of r-process material are terminated by the tran-
sition of the proto-neutron star into a BH in the H events
while prolonged ejection is sustained by neutrino emission
from a stable NS in the L events (Qian et al. 1998a).
In summary, despite the lack of a complete model for
successful r-process production in SNe, there is a self-
consistent picture of r-process enrichment by SNe that can
account for the meteoritic data on 182Hf and 129I in the
early solar system and the observed r-process abundances
in metal-poor stars. Furthermore, the observed correla-
tion between abundances of r-process elements and Fe at
[Fe/H] ∼> −2.5 (Gratton & Sneden 1994; Crawford et al.
1998; see also McWilliam et al. 1995) can be understood as
the result of sufficient mixing of r-process products and Fe
from multiple SN events in this picture. In fact, the same
picture was used by Wasserburg & Qian (WQ) as the ba-
sic framework to explain the dispersion in abundances of
heavy r-process elements such as Eu at [Fe/H] ∼ −3.
4. DISCUSSION AND CONCLUSIONS
The amount of ISM to mix with the ejecta from an in-
dividual NSM event is limited by the total energy of the
event. On the other hand, due to the very low rate of
NSMs in the Galaxy, a large amount of r-process ejecta
would be required from each event to account for the
present Galactic r-process inventory. When mixed with
the ISM, this required amount of ejecta would result in
abundances of r-process elements with A ≤ 130 (such as
Ag) and A > 130 (such as Eu) that are much too high (by
factors ∼ 10 -- 102 for Ag and ∼ 30 -- 103 for Eu) compared
with the observed values in metal-poor stars. Further-
more, an average ISM received the ejecta from only ∼ 1
NSM event over Galactic history. If r-process enrichment
of the ISM was provided by NSMs in this way while Fe
enrichment was provided by many SNe, there would be
no correlation between abundances of r-process elements
and Fe, in disagreement with the observed correlation at
[Fe/H] ∼> −2.5. While the complexities in mixing of the
ejecta with the ISM can affect the above considerations in
detail, it is unlikely that they can remove the difficulty of
the NSM r-process model in explaining the observations
of metal-poor stars (especially the rather early onset of
the correlation between abundances of r-process elements
and Fe at [Fe/H] ≈ −2.5). Nevertheless, future numerical
studies of r-process enrichment by NSMs accompanied by
Fe enrichment by SNe should be interesting to pursue and
may give a more definitive answer.
In contrast, a self-consistent picture of r-process enrich-
ment by SNe can be obtained by considering mixing of the
ejecta from an individual event with the ISM. Here an av-
erage ISM is enriched in r-process elements with A > 130
by the H events at a frequency fH ∼ (107 yr)−1 and
in those with A ≤ 130 by the L events at a frequency
fL ∼ (108 yr)−1. This picture can account for the mete-
oritic data on 182Hf and 129I in the early solar system and
the observed r-process abundances in metal-poor stars.
Furthermore, sufficient mixing of r-process products and
Fe from multiple SN events in this picture would result in
the observed correlation between abundances of r-process
elements and Fe at [Fe/H] ∼> −2.5. The same picture was
also used by Wasserburg & Qian (WQ) as the basic frame-
work to explain the dispersion in abundances of heavy r-
4
process elements such as Eu at [Fe/H] ∼ −3. However, a
complete model of r-process and Fe production in the H
and L events is still lacking and should be investigated by
future theoretical studies.
In discussing r-process enrichment by NSMs I have as-
sumed that the maximum amount of ISM to mix with the
ejecta from an individual event is the same as swept up
by a SN remnant. This is because the total energy of the
NSM ejecta seen in numerical simulations (Rosswog et al.
1999) is at most comparable to the SN explosion energy
(∼ 1051 erg). For given conditions of the ISM, the expan-
sion/evolution of the ejecta is essentially governed by its
total energy. The large difference in the amount of ejecta
between a NSM and a SN has no significant effect here as in
both cases the mass of the swept-up ISM soon overwhelms
that of the ejecta while the total energy remains more or
less the same. I note that a small amount (∼< 10−5 M⊙)
of material might be ejected in highly-relativistic jets in a
NS-BH merger event (Janka et al. 1999). However, the
total energy of these jets is ∼< 1051 erg (Janka et al. 1999)
and their existence would not increase significantly the
amount of ISM that could mix with the entire ejecta from
the event. It takes ∼ 106 yr for the energy (∼ 1051 erg) and
the associated momentum of the ejecta to be dispersed in
the ISM (e.g., Thornton et al. 1998), where the next NSM
or SN would occur on a much longer timescale (∼ 1010 yr
for NSMs and ∼ 107 yr for SNe). This leaves substan-
tial time for mixing of the ejecta with the swept-up ISM.
However, the details of the mixing process require further
studies.
If, as argued here, SNe are the major sources for the
r-process, then there are two possible direct observational
tests: gamma rays from decay of r-process nuclei in SN
remnants and surface contamination of the companion by
SN r-process ejecta in binaries. Qian et al. (1998b, 1999)
discussed gamma-ray signals from decay of long-lived r-
process nuclei (with lifetimes ∼ 103 -- 105 yr) in a nearby
SN remnant and from decay of short-lived r-process nu-
clei (with lifetimes ∼ 1 -- 10 yr) produced in a Galactic
SN that may occur in the future. The nuclide 126Sn is
of particular interest (Qian et al. 1998b) as its lifetime
(∼ 105 yr) is much longer than the age (∼ 104 yr) of the
Vela SN remnant at a distance ≈ 250 pc. In the picture
of r-process enrichment by SNe discussed here (see also
QW; WQ) the solar r-process mass fraction of nuclei with
A ≤ 130 resulted from ∼ 102 L events (see §3). So a total
mass X A=126
⊙,r Mmix/102 ∼ 5 × 10−7 M⊙ of 126Sn nuclei are
≈ 1.6 × 10−9
produced in each L event, where X A=126
⊙,r
(Kappeler et al. 1989) is the solar r-process mass frac-
tion of 126Te, the stable daughter of 126Sn.
If the SN
associated with the Vela remnant was an L event, then
decay of 126Sn in this remnant would produce gamma-
ray fluxes ∼ 10−7 γ cm−2 s−1 at energies Eγ = 415, 666,
and 695 keV. Detection of these fluxes would require fu-
ture gamma-ray experiments such as the proposed Ad-
vanced Telescope for High Energy Nuclear Astrophysics
(ATHENA, Kurfess 1994). As the Vela remnant contains
a pulsar, such detection would also provide evidence for
the speculated association between L events and SNe pro-
ducing neutron stars (Qian et al. 1998a).
The other test mentioned above takes advantage of the
occurrence of SNe in binaries. The r-process ejecta from
the SN would contaminate the surface of its binary com-
panion. Some binaries would survive the SN explosion
and acquire a NS or a BH in place of the SN progenitor.
Therefore, an ordinary star observed to be the binary com-
panion of a NS or a BH might show r-process abundance
anomalies on the surface. To estimate the plausible level
of such anomalies, I assume that a fraction ∼ 10−3 of the
entire SN ejecta (∼ 10 M⊙, mostly non-r-process material)
would be intercepted by a low mass (∼ 1 M⊙) companion
and then mixed with ∼ 10−2 M⊙ of the surface material.
If the SN was an H event, ∼ 10−9 M⊙ of r-process ele-
ments with A > 130 would be intercepted, while for an
L event ∼ 10−8 M⊙ of r-process elements with A ≤ 130
would be intercepted (see §3). These quantities are to
be compared with ∼ 4 × 10−10 M⊙ of the corresponding r-
process elements in ∼ 10−2 M⊙ of the surface material in a
companion star of solar r-process composition. So a SN in
a binary could enhance significantly the surface r-process
abundances in the companion star, especially if the SN
was an L event. In view of the large overabundance of O,
Mg, Si, and S recently observed in the companion star of a
probable BH (Israelian et al. 1999), detection of r-process
enhancement in similar binary systems seems promising.
Such detection may also test directly the speculation by
Qian et al. (1998a) that H events are associated with SNe
producing BHs while L events are associated with SNe
producing neutron stars.
I thank Petr Vogel and Jerry Wasserburg for many dis-
cussions on the r-process. I am also grateful to the referee,
Friedel Thielemann, for detailed criticisms that help im-
prove the paper. This work was supported in part by the
Department of Energy under grant DE-FG02-87ER40328.
Arzoumanian, Z., Cordes, J. M., & Wasserman, I. 1999, ApJ, 520,
Harper, C. L., & Jacobsen, S. B. 1996, Geochim. Cosmochim. Acta,
REFERENCES
696
60, 1131
Bethe, H. A., & Brown, G. E. 1998, ApJ, 506, 780
Brazzle, R. H., Pravdivtseva, O. V., Meshik, A. P., & Hohenberg, C.
Hoffman, R. D., Woosley, S. E., & Qian, Y.-Z. 1997, ApJ, 482, 951
Israelian, G., Rebolo, R., Basri, G., Casares, J., & Martin, E. L.
M. 1999, Geochim. Cosmochim. Acta, 63, 739
1999, Nature, 401, 142
Cowan, J. J., Sneden, C., Ivans, I., Burles, S., Beers, T. C., & Fuller,
Janka, H.-Th., Eberl, Th., Ruffert, M., & Fryer, C. L. 1999, ApJ,
G. 1999, BAAS, 31, 930
Crawford, J. L., Sneden, C., King, J. R., Boesgaard, A. M., &
527, L39
Kappeler, F., Beer, H., & Wisshak, K. 1989, Rep. Prog. Phys., 52,
Deliyannis, C. P. 1998, AJ, 116, 2489
Freiburghaus, C., Rembges, J.-F., Rauscher, T., Kolbe, E.,
Thielemann, F.-K., Kratz, K.-L., Pfeiffer, B., & Cowan, J. J.
1999a, ApJ, 516, 381
Freiburghaus, C., Rosswog, S., & Thielemann, F.-K. 1999b, ApJ,
945
Kratz, K.-L., Bitouzet, J.-P., Thielemann, F.-K., Moller, P., &
Pfeiffer, B. 1993, ApJ, 403, 216
Kurfess, J. D. 1994, ATHENA Mission Proposal, NASA New Mission
Concepts in Astrophysics
525, L121
Gratton, R. G., & Sneden, C. 1994, A&A, 287, 927
Lattimer, J. M., & Schramm, D. N. 1974, ApJ, 192, L145
Lattimer, J. M., & Schramm, D. N. 1976, ApJ, 210, 549
Lee, D.-C., & Halliday, A. N. 1995, Nature, 378, 771
Lee, D.-C., & Halliday, A. N. 1998, Lunar and Planetary Science
Rosswog, S., Liebendorfer, M., Thielemann, F.-K., Davies, M. B.,
5
Conference, 29, 1416
Mathews, G. J., & Cowan, J. J. 1990, Nature, 345, 491
McWilliam A., Preston, G. W., Sneden, C., & Searle, L. 1995, AJ,
Benz, W., & Piran, T. 1999, A&A, 341, 499
Sneden, C., Cowan, J. J., Burris, D. L., & Truran, J. W. 1998, ApJ,
496, 235
Sneden, C., McWilliam, A., Preston, G. W., Cowan, J. J., Burris, D.
109, 2757
Meyer, B. S., & Brown, J. S. 1997, ApJS, 112, 199
Meyer, B. S., Howard, W. M., Mathews, G. J., Woosley, S. E., &
Hoffman, R. D. 1992, ApJ, 399, 656
Qian, Y.-Z., Vogel, P., & Wasserburg, G. J. 1998a, ApJ, 494, 285
Qian, Y.-Z., Vogel, P., & Wasserburg, G. J. 1998b, ApJ, 506, 868
Qian, Y.-Z., Vogel, P., & Wasserburg, G. J. 1999, ApJ, 524, 213
Qian, Y.-Z., & Wasserburg, G. J. 2000, Phys. Rep., in press (QW)
Qian, Y.-Z., & Woosley, S. E. 1996, ApJ, 471, 331
Reynolds, J. H. 1960, Phys. Rev. Lett., 4, 8
L., & Armosky, B. J. 1996, ApJ, 467, 819
Symbalisty, E., & Schramm, D. N. 1982, Astrophys. Lett., 22, 143
Takahashi, K., Witti, J., & Janka, H.-Th. 1994, A&A, 286, 857
Thornton, K., Gaudlitz, M., Janka, H.-Th., & Steinmetz, M. 1998,
ApJ, 500, 95
Wasserburg, G. J., Busso, M., & Gallino, R. 1996, ApJ, 466, L109
Wasserburg, G. J., & Qian, Y.-Z. 2000, ApJ, 529, L21 (WQ)
Woosley, S. E., & Hoffman, R. D. 1992, ApJ, 395, 202
Woosley, S. E., Wilson, J. R., Mathews, G. J., Hoffman, R. D., &
Meyer, B. S. 1994, ApJ, 433, 229
|
astro-ph/9910251 | 1 | 9910 | 1999-10-14T18:27:26 | High-Redshift Quasars as Probes of Galaxy and Cluster Formation | [
"astro-ph"
] | Quasars at large redshifts provide a powerful probe of structure formation in the early universe. Several arguments suggest that the formation of ellipticals and massive bulges may have involved an early quasar phase. At very large redshifts, such structures are likely to be found at the highest peaks of the density field, and would thus be highly biased tracers: the earliest (massive) galaxy formation may have occurred in the cores of future rich clusters. Preliminary results from our search for clustered protogalaxies around quasars at z > 4 support this idea. Quasars at even larger redshifts may be an important contributor to the reionisation of the universe, and signposts of the earliest galaxy and cluster formation. | astro-ph | astro-ph |
To appear in "The Hy-Redshift Universe:
Galaxy Formation and Evolution at High Redshift"
eds. A.J. Bunker & W.J.M. van Breugel
Astr. Soc. Pacif. Conf. Ser., in press (1999)
High-Redshift Quasars as Probes
of Galaxy and Cluster Formation
S. G. Djorgovski
Palomar Observatory, Caltech, Pasadena, CA 91125, USA
Abstract. Quasars at large redshifts provide a powerful probe of struc-
ture formation in the early universe. Several arguments suggest that the
formation of ellipticals and massive bulges may have involved an early
quasar phase. At very large redshifts, such structures are likely to be
found at the highest peaks of the density field, and would thus be highly
biased tracers: the earliest (massive) galaxy formation may have occurred
in the cores of future rich clusters. Preliminary results from our search
for clustered protogalaxies around quasars at z > 4 support this idea.
Quasars at even larger redshifts may be an important contributor to the
reionisation of the universe, and signposts of the earliest galaxy and clus-
ter formation.
1.
Introduction
Hy Spinrad always hated quasars (Spinrad 1979). In those paleolithic days (i.e.,
before 1987 or so) it was not yet obvious that a powerful quasar lurks in the heart
of every one of Hy's beloved radio galaxies, but in some sense this does not really
matter:
in the work of the Spinrad School of Observational Cosmology, AGN
have been used simply as means to find stellar populations at large redshifts, in
order to probe their formation and evolution. This is in principle a viable and
sound approach.
In the simplest view, the very existence of luminous quasars at large red-
shifts suggests the existence of their (massive?) host galaxies, at least in the
minds of a vast majority of astronomers today. At z > 4, this has some very
interesting and non-trivial implications for our understanding of galaxy and
structure formation (Turner 1991).
At a slightly more complex level, the observed history of the comoving
number density of quasars may be indicative of the history of galaxy formation
and evolution: the same kind of processes, i.e., dissipative mergers and tidal
interactions, may be fueling both bursts of star formation and AGN activity.
The peak seen in the comoving number density of quasars around z ∼ 2 or 3
(Schmidt, Schneider & Gunn 1995) can then be interpreted in this context: the
1
ostensible decline at high redshifts may be indicative of the initial assembly and
growth of quasar central engines and their host galaxies; whereas the decline
at lower redshifts may be indicative of the decrease in fueling, as galaxies are
carried apart by the universal expansion, as many of the smaller pieces are being
consumed, and as the gas is being converted into stars. Qualitatively similar
predictions are made by virtually all models of hierarchical structure formation
(see, e.g., Cataneo 1999, or Kauffmann & Haehnelt 1999).
Due to their brightness, quasars are much easier to find (per unit telescope
time) than galaxies at comparable redshifts. It then makes sense to use quasars
as probes, or at least as pointers of sites of galaxy formation.
Quasars have been used very effectively as probes of the intergalactic medium,
and indirectly of galaxy formation, through the studies of absorption line sys-
tems. A vast literature exists on this subject, which is beyond the scope of this
review; for good summaries, see, e.g., Rees (1998a) or Rauch (1998).
A good review of the searches for quasars and related topics was given by
Hartwick & Schade (1990). Osmer (1999) provides a modern update. Some of
the issues covered in this review have been described by Djorgovski (1998) and
Djorgovski et al. (1999).
2. Quasars and Galaxy Formation
Possibly the most direct evidence for a close relation between quasars and galaxy
formation is the remarkable correlation between the masses of central black holes
(MBH) in nearby galaxies, and the luminosities (∼ masses) of their old, metal-
rich stellar populations, a.k.a. bulges (Kormendy & Richstone 1995, Magorrian
et al. 1998), with MBH's containing on average ∼ 0.6% of the bulge stellar
mass. The most natural explanation for this correlation is that both MBH's
and the stellar populations are generated through a parallel set of processes,
i.e., dissipative merging and assembly at large redshifts. Quiescent MBH's are
evidently common among the normal galaxies at z ∼ 0, and had to originate at
some point: as they grow by accretion, their formation is the quasar activity.
Quasars may thus be a common phase of the early formation of ellipticals and
massive bulges.
Quasar demographics support this idea. Small & Blandford (1992), Chokshi
& Turner (1992), and Haehnelt & Rees (1993) all conclude that an average L∗-ish
galaxy today should contain an MBH with M• ∼ 107 M⊙ or so. These estimates
(essentially integrating the known AGN radiation over the past history of the
universe) are fully consistent with the actual census of MBH (quasar remnants)
in nearby galaxies.
Two other pieces of fossil evidence link the high-z quasars with the forma-
tion of old, metal-rich stellar populations. First, the analysis of metallicities
in QSO BEL regions indicates super-solar abundances (up to ZQ ∼ 10 Z⊙!)
in quasars at z > 4 (Hamann & Ferland 1993, 1999; Matteucci & Padovani
1993). The only places we know such abundances to occur are the nuclei of
giant elliptical galaxies. Furthermore, abundance patterns in the intracluster x-
ray gas at low redshifts are suggestive of an early, rapid star formation phase in
protoclusters populated by young ellipticals (Loewenstein & Mushotzky 1996).
2
A nearly simultaneous formation of quasars and their host galaxies, or at
least ellipticals and bulges, is consistent with all of these observations, and it fits
naturally in the general picture of hierarchical galaxy and structure formation
via dissipative merging (see, e.g., Norman & Scoville 1988, Sanders et al. 1988,
Carlberg 1990, Hernquist & Mihos 1995, Mihos & Hernquist 1996, Monaco et
al. 1999, Franceschini et al. 1999, etc.).
An extreme case of this idea is that quasars are completely reducible to
ultraluminous starbursts, as advocated for many years by Terlevich and collab-
orators (see, e.g., Terlevich & Boyle 1993, and references therein). Most other
authors disagree with such a view (cf. Heckman 1991, or Williams & Perry 1994),
but (nearly) simultaneous manifestations of both ultraluminous starbursts and
AGN, perhaps with comparable energetics, are clearly allowed by the data. It is
thus also possible that the early AGN can have a profound impact on their still
forming hosts, through the input of energy and momentum (Ikeuchi & Norman
1991, Haehnelt et al. 1998).
3. Quasar (Proto)Clustering and Biased Galaxy Formation
Producing sufficient numbers of massive host galaxies needed to accommodate
the observed populations of quasars at z > 4, say, is not easy for most hierarchi-
cal models: such massive halos should be rare, and associated on average with
∼ 4 to 5-σ peaks of the primordial density field (Efstathiou & Rees 1988, Cole &
Kaiser 1989, Nusser & Silk 1993). It is a generic prediction that for essentially
every model of structure formation such high density peaks should be strongly
clustered (Kaiser 1984). This is a purely geometrical effect, independent of any
messy astrophysical details of galaxy formation, and thus it is a fairly robust
prediction: the formation of the first galaxies (some of which may be the hosts
of high-z quasars) and of the primordial large-scale structure should be strongly
coupled.
Quasars provide a potentially useful probe of large-scale structure out to
very high redshifts. The pre-1990 work has been reviewed by Hartwick & Schade
(1990). A number of quasar pairs on tens to hundreds of comoving kpc scales has
been seen (Djorgovski 1991, Kochanek et al. 1999), and some larger groupings
on scales reaching ∼ 100 Mpc (Crampton et al. 1989, Clowes & Campusano
1991), but all in heterogeneous data sets. Analysis of some more complete
samples did show a clustering signal (e.g., Iovino & Shaver 1988, Boyle et al.
1998). The overall conclusion is that quasar clustering has been detected, but
that its strength decreases from z ∼ 0 out to z ∼ 2, the peak of the quasar era,
presumably reflecting the linear growth of the large-scale structure. However,
if quasars are biased tracers of structure formation at even higher redshifts,
associated with very massive peaks of the primordial density field, this trend
should reverse and the clustering strength should again start increasing towards
the larger look-back times.
The first hints of such an effect were provided by the three few-Mpc quasar
pairs found in the statistically complete survey by Schneider et al. (1994), as
pointed out by Djorgovski et al.
(1993) and Djorgovski (1996), and subse-
quently confirmed by more detailed analysis by Kundic (1997) and Stephens et
al. (1997). A deeper survey for more such pairs by Kennefick et al. (1996) did
3
not find any more, presumably due to a limited volume coverage. La Franca et
al. (1998) find a turn-up in the clustering strength of quasars even at redshifts
as low as z ∼ 2. It would be very important to check these results with new,
large, complete samples of quasars over a wide baseline in redshift.
More recently, observations of large numbers of "field" galaxies at z ∼ 3−3.5
(1998) identified redshift space structures which are almost
by Steidel et al.
certainly the manifestation of biasing. However, the effect (the bias) should
be even stronger at higher redshifts, and most of the earliest massive galaxies
should be strongly clustered. A search for protoclusters around known high-z
objects such as quasars thus provides an important test of our basic ideas about
the biased galaxy formation.
Intriguingly, there is a hint of a possible superclustering of quasars at z > 4,
on scales ∼ 100 h−1 comoving Mpc (cf. Djorgovski 1998). The effect is clearly
present in the DPOSS sample (which is complete, but still with a patchy coverage
on the sky), and in a more extended, but heterogeneous sample of all QSOs at
z > 4 reported to date. The apparent clustering in the complete sample may be
an artifact of a variable depth of the survey, which we will be able to check in
a near future. Or, it could be due to patchy gravitational lensing magnification
of the high-z quasars by the foreground large-scale structure; again, we will
be able to test this hypothesis using the DPOSS galaxy counts. But it could
also represent real clustering of high-density peaks in the early universe, only
∼ 0.5 − 1 Gyr after the recombination. The observed scale of the clustering is
intriguing: it is comparable to that corresponding to the first Doppler peak seen
in CMBR fluctuations, and to the preferred scales seen in some redshift surveys
(e.g., Broadhurst et al. 1990; Landy et al. 1996). More data are needed to check
on this remarkable result.
4. Quasar-Marked Protoclusters at z > 4?
Any single search method for high-z protogalaxies (PGs) has its own biases, and
formative histories of galaxies in different environments may vary substantially.
For example, galaxies in rich clusters are likely to start forming earlier than in
the general field, and studies of galaxy formation in the field may have missed
possible rare active spots associated with rich protoclusters.
We are conducting a systematic search for clustered PGs, by using quasars
at z > 4 as markers of the early galaxy formation sites (ostensibly protocluster
cores). The quasars themselves are selected from the DPOSS survey (Djorgovski
et al. 1999, and in prep.; Kennefick et al. 1995). They are purely incidental to
this search: they are simply used as beacons, pointing towards the possible sites
of early, massive galaxy formation.
The first galaxy discovered at z > 3 was a quasar companion (Djorgovski
et al. 1985, 1987). A Lyα galaxy and a dusty companion of BR 1202 -- 0725 at
z = 4.695 have been discovered by several groups (Djorgovski 1995, Hu et al.
1996, Petitjean et al. 1996), and a dusty companion object has been found in
the same field (Omont et al. 1995, Ohta et al. 1995). Hu & McMahon (1996)
also found two companion galaxies in the field of BR 2237 -- 0607 at z = 4.55.
We have searches to various degrees of completeness in about twenty QSO
fields so far (Djorgovski 1998; Djorgovski et al., in prep.). Companion galaxies
4
PSS 0030+1702 (z=4.305)
PSS 0117+1552 (z=4.275)
PSS 0248+1802 (z=4.465)
PSS 1721+3256 (z=4.031)
Figure 1.
Examples of clustered companion protogalaxies in the
fields of 4 DPOSS quasars at z > 4. These are deep, R-band Keck
images centered on the quasars. The fields shown are 54 arcsec square.
Some of the spectroscopically confirmed companions are labeled with
the arrows.
have been found in virtually every case, despite very incomplete coverage. They
are typically located anywhere between a few arcsec to tens of arcsec from the
quasars, i.e., on scales ∼ 100+ comoving kpc. We also select candidate PGs
by using deep BRI imaging over a field of view of several arcmin, probing
∼ 10 comoving Mpc (∼ cluster size) projected scales. This is a straightforward
extension of the method employed so successfully to find the quasars themselves
at z > 4 (at these redshifts, the continuum drop is dominated by the Lyα forest,
rather than the Lyman break, which is used to select galaxies at z ∼ 2 − 3.5).
The candidates are followed up by multislit spectroscopy at the Keck, which is
still in progress as of this writing.
As of the mid-1999, about two dozen companion galaxies have been con-
firmed spectroscopically. Their typical magnitudes are R ∼ 25m, implying con-
tinuum luminosities L ≤ L∗. The Lyα line emission is relatively weak, with
typical restframe equivalent widths ∼ 20 − 30 A, an order of magnitude lower
5
than what is seen in quasars and powerful radio galaxies, but perfectly reason-
able for the objects powered by star formation. There are no high-ionization
lines in their spectra, and no signs of AGN. The SFR inferred both from the
Lyα line, and the UV continuum flux is typically ∼ 5 − 10 M⊙/yr, not corrected
for the extinction, and thus it could easily be a factor of 5 to 10 times higher.
Overall, the intrinsic properties of these quasar companion galaxies are very
similar to those of the Lyman-break selected population at z ∼ 3 − 4, except of
course for their special environments and somewhat higher look-back times.
There is a hint of a trend that the objects closer to the quasars have stronger
Lyα line emission, as it may be expected due to the QSO ionization field. In
addition to these galaxies where we actually detect (presumably starlight) con-
tinuum, pure Lyα emission line nebulae are found within ∼ 2 − 3 arcsec for
several of the quasars, with no detectable continuum at all. The Lyα fluxes are
exactly what may be expected from photoionization by the QSO, with typical
LLyα ∼ a few ×1043 erg/s. They may represent ionized parts of still gaseous
protogalaxy hosts of the quasars. We can thus see and distinguish both the
objects powered by the neighboring QSO, and "normal" PGs in their vicinity.
The median projected separations of these objects from the quasars are ∼ a
few ×100h−1 comoving kpc, an order of magnitude less than the comoving r.m.s.
separation of L∗ galaxies today, but comparable to that in the rich cluster cores.
The frequency of QSO companion galaxies at z > 4 also appears to be an order
of magnitude higher than in the comparable QSO samples at z ∼ 2 − 3, the peak
of the QSO era and the ostensible peak merging epoch. However, interaction
and merging rates are likely to be high in the densest regions at high redshifts,
which would naturally account for the propensity of some of these early PGs to
undergo a quasar phase, and to have close companions.
The implied average star formation density rate in these regions is some
2 or 3 orders of magnitude higher than expected from the limits estimated for
these redshifts by Madau et al. (1996) for f ield galaxies, and 1 or 2 orders of
magnitude higher than the measurements by Steidel et al. at z ∼ 4, even if
we ignore any SFR associated with the QSO hosts (which we cannot measure,
but is surely there). These must be very special regions of an enhanced galaxy
formation in the early universe.
It is also worth noting that (perhaps coincidentally) the observed comoving
number density of quasars at z > 4 is roughly comparable to the comoving den-
sity of very rich clusters of galaxies today. Of course, depending on the timescales
involved, there must be some protoclusters without observable quasars in them,
and some where more than one AGN is present (an example may be the obscured
companion of BR 1202 -- 0725).
5. Towards the Renaissance at z > 5: the First Quasars
and the First Galaxies
The remarkable progress in cosmology over the past few years, reviewed by
several speakers at this meeting, has pushed the frontiers of galaxy and structure
formation studies out to z > 5. Half a dozen galaxies, two QSOs (cf. Fan
et al. 1999), and one radio galaxy are now known at z ≥ 5, with the most
distant confirmed object at z = 5.74 (Hu et al. 1999). Remarkably, there is
6
no convincing evidence yet for a high-z decline of the comoving star formation
rate density out to z > 4 (Steidel et al. 1999). Moreover, the universe at z ∼ 5
appears to be already fully reionised (Songaila et al. 1999, Madau et al. 1999),
implying the existence of a substantial activity in a population of sources at
even higher redshifts.
These observational results pose something of a challenge for the models
of galaxy formation. Essentially in all modern models, the first subgalactic
fragments with masses ≥ 106 M⊙ begin to form at z ∼ 10 − 30, and the universe
becomes reionised at z ∼ 8 − 12 (see, e.g., Gnedin & Ostriker 1997, Miralda-
Escude & Rees 1997, or Rauch 1998 and references therein). This corresponds to
a time interval of only about ∼ 0.5 −1 Gyr for a reasonable range of cosmologies.
What is not known is what are the first or the dominant ionisation sources
which break the "dark ages": primordial starbursts or primordial AGN? This is
one of the fundamental questions in cosmology today, and it dominates many of
the discussions about the NGST (see, e.g., Rees 1998b, Haiman & Loeb 1998, or
Loeb 1999). Optical searches for quasars at z > 5 have been reviewed by Osmer
(1999). There are exciting new prospects of detecting such a population in x-
rays using CXO (Haiman & Loeb 1999). The value of such quasars as probes
of the earliest phases of galaxy and structure formation during the reionisation
era at z ∼ 10 × 2±1 cannot be overstated.
Some numerical simulations suggest that an early formation of quasars, at
z ∼ 8, say, is viable in the framework of the currently popular hierarchical models
with dissipation (cf. Katz et al. 1994). It is even possible that a substantial
amount of QSO activity may predate the peak epoch of star formation in galaxies
(Silk & Rees 1998). A catastrophic gravitational collapse of a massive primordial
star cluster may be the most natural way of forming the first MBHs, but a variety
of other mechanisms have been proposed (e.g., Loeb 1993, Umemura et al. 1993,
Loeb & Rasio 1994, etc.). Future observations will tell whether such primordial
fireworks marked the end of the dark ages in the universe.
Acknowledgments.
It is a pleasure to acknowledge the work of my collab-
orators, R. Gal, R. Brunner, R. de Carvalho, S. Odewahn, and the rest of the
DPOSS QSO search team. I also wish to thank the staff of Palomar and Keck
observatories for their expert assistance during our observing runs. This work
was supported in part by the Norris Foundation and by the Bressler Foundation.
Ivan King and his LOC crew brought this meeting into existence; thank you all.
Finally, many thanks to Hy for introducing me to the joys of low-S/N astronomy
and letting me play with the big toys: it was fun (most of the time)!
References
Boyle, B., Croom, S., Smith, R., Shanks, T., Miller, L., & Loaring, N. 1998,
preprint [astro-ph/9805140]
Broadhurst, T., Ellis, R., Koo, D., & Szalay, A. 1990, Nat, 343, 726
Cataneo, A. 1999, MNRAS in press [astro-ph/9907335]
Carlberg, R. 1990, ApJ, 350, 505
Chokshi, A., & Turner, E. 1992, MNRAS, 259, 421
Clowes, R., & Campusano, L. 1991, MNRAS, 249, 218
7
Cole, S. & Kaiser, N. 1989, MNRAS, 237, 1127
Crampton, D., Cowley, A., & Hartwick, F.D.A. 1989, ApJ, 345, 59
Djorgovski, S., Spinrad, H., McCarthy, P., & Strauss, M. 1985, ApJ, 299, L1
Djorgovski, S., Strauss, M. , Perley, R., Spinrad, H., & McCarthy, P. 1987, AJ,
93, 1318
Djorgovski, S. 1991, in The Space Distribution of Quasars, ed. D. Crampton,
ASPCS, 21, 349
Djorgovski, S., Thompson, D., & Smith, J. 1993, in First Light in the Universe,
eds. B. Rocca-Volmerange et al., Gif sur Yvette: Eds. Fronti`eres, p.67
Djorgovski, S.G. 1995, in Science with the VLT, eds. J.R. Walsh & I.J. Danziger,
Berlin: Springer Verlag, p. 351
Djorgovski, S., Pahre, M., Bechtold, J., & Elston, R. 1996, Nat, 382, 234
Djorgovski, S.G. 1996, in New Light on Galaxy Evolution, IAU Symp. 171, eds.
R. Bender & R. Davies, Dordrecht: Kluwer, p.277
Djorgovski, S.G. 1998,
in Fundamental Parameters in Cosmology, Rec. de
Moriond, eds. Y. Giraud-Heraud et al., Gif sur Yvette: Eds. Fronti`eres,
p. 313 [astro-ph/9805159]
Djorgovski, S.G., Odewahn, S.C., Gal, R.R., Brunner, R., & de Carvalho, R.R.
1999, in Photometric Redshifts and the Detection of High Redshift Galax-
ies, eds. R. Weymann et al., ASPCS in press [astro-ph/9908142]
Djorgovski, S.G., Gal, R. R., Odewahn, S. C., de Carvalho, R. R., Brunner, R.,
Longo, G., & Scaramella, R. 1999, in Wide Field Surveys in Cosmology,
eds. S. Colombi et al., Gif sur Yvette: Eds. Fronti`eres, p.89 [astro-
ph/9809187]
Efstathiou, G., & Rees, M. 1988, MNRAS, 230, P5
Fan, X. et al. (the SDSS collaboration) 1999, AJ, 118, 1
Franceschini, A., Hasinger, G., Miyaji, T., & Malquori, D. 1999, MNRAS in
press [astro-ph/9909290]
Gnedin, N., & Ostriker, J. 1997, ApJ, 486, 581
Haehnelt, M., & Rees, M. 1993, MNRAS, 263, 168
Haehnelt, M., Natarajan, P., & Rees, M. 1998, MNRAS, 300, 817
Haiman, Z., & Loeb, A. 1998, ApJ, 503, 505
Haiman, Z., & Loeb, A. 1999, ApJ, 521, L9
Hamann, F., & Ferland, G. 1993, ApJ, 418, 11
Hamann, F., & Ferland, G. 1999, ARAA in press
Hartwick, F.D.A, & Schade, D. 1990, ARAA, 28, 437
Heckman, T. 1991, in Massive Stars in Starbursts, eds. C. Leitherer et al., STScI
Symposium No. 5, Cambridge: Cambridge Univ. Press, p.289
Hernquist, L., & Mihos, C. 1995, ApJ, 448, 41
Hu, E., McMahon, R., & Egami, E. 1996, ApJ, 459, L53
Hu, E., & McMahon, R. 1996, Nat, 382, 231
Hu, E., McMahon, R., & Cowie, L. 1999, ApJL in press [astro-ph/9907079]
Ikeuchi, S., & Norman, C. 1991, ApJ, 375, 479
8
Iovino, A., & Shaver, P. 1988, ApJ, 330, L13
Kaiser, N. 1984, ApJ, 284, L9
Katz, N., Quinn, T., Bertschinger, E., & Gelb, J. 1994, MNRAS, 270, L71
Kauffmann, G., & Haehnelt, M. 1999, MNRAS in press [astro-ph/9906493]
Kennefick, J.D., Djorgovski, S.G., & de Carvalho, R. 1995, AJ, 110, 2553
Kennefick, J.D., Djorgovski, S.G., & Meylan, G. 1996, AJ, 111, 1816
Kochanek, C., Falco, E., & Munoz, J. 1999, ApJ, 510, 590
Kormendy, J., & Richstone, D. 1995, ARAA, 33, 581
Kundic, T. 1997, ApJ, 482, 631
La Franca, F., Andreani, P., & Cristiani, S. 1998, ApJ, 497, 529
Landy, S., Shectman, S., Lin, H., Kirshner, R., Oemler, A., & Tucker, D. 1996,
ApJ, 456, L1
Loeb, A. 1993, ApJ, 403, 542
Loeb, A., & Rasio, F. 1994, ApJ, 432, 52
Loeb, A. 1999, this volume
Loewenstein, M., & Mushotzky, R. 1996, ApJ, 466, 695
Madau, P., Ferguson, H., Dickinson, M., Giavalisco, M., Steidel, C., & Fruchter,
A. 1996, MNRAS, 283, 1388
Madau, P., Haardt, F., & Rees, M. 1999, ApJ, 514, 648
Magorrian, J. et al. 1998, AJ, 115, 2285
Matteucci, F,. and Padovani, P. 1993, ApJ, 419 485
Mihos, C., & Hernquist, L. 1996, ApJ, 464, 641
Miralda-Escude, J., & Rees, M. 1997, ApJ, 478, L57
Monaco, P., Salucci, P., & Danese, L. 1999, MNRAS in press [astro-ph/9907095]
Norman, C., & Scoville, N. 1988, ApJ, 332, 124
Nusser, A., & Silk, J. 1993, ApJ, 411, L1
Ohta, K., Yamada, T., Nakanishi, K., Kohno, K,. Akiyama, M., & Kawabe, R.
1996, Nat, 382, 426
Omont, A., Petitjean, P., Guilloteau, S., McMahon, R., Solomon, P., & Pecontal,
E. 1996, Nat, 382, 428
Osmer, P. 1999, this volume
Petitjean, P., Pecontal, E., Vals-Gabaud, D., & Charlot, S. 1996, Nat, 380, 411
Rauch, M. 1998, ARAA, 36, 267
Rees, M. 1998a, in Structure and Evolution of the Intergalactic Medium from
QSO Absorption Systems, eds. P. Petitjean & S. Charlot, Gif sur Yvette:
Eds. Frontieres, p. 19
Rees, M. 1998b, preprint [astro-ph/9809029]
Sanders, D., Soifer, B.T., Elias, J., Neugebauer, G., & Matthews, K. 1988, ApJ,
328, L35
Schmidt, M., Schneider, D., & Gunn, J. 1995, AJ, 110, 68
Schneider, D., Schmidt, M., & Gunn, J. 1994, AJ 107, 1245
Silk, J., & Rees, M. 1998, A&A, 331, L1
9
Small, T., & Blandford, R. 1992, MNRAS, 259, 725
Songaila, A., Hu, E., Cowie, L., & McMahon, R. 1999, ApJL in press
Spinrad, H. 1979, private communication
Steidel, C., Giavalisco, M., Pettini, M., Dickinson, M., & Adelberger, K. 1996,
ApJ, 462, L17
Steidel, C., Adelberger, K., Dickinson, M., Pettini, M., & Kellogg, M. 1998,
ApJ, 492, 428
Steidel, C., Adelberger, K., Giavalisco, M., Dickinson, M., & Pettini, M. 1999,
ApJ, 519, 1
Stephens, A., Schneider, D., Schmidt, M., Gunn, J., & Weinberg, D. 1997, AJ,
114, 41
Terlevich, R., & Boyle, B. 1993, MNRAS, 262, 491
Turner, E. 1991, AJ, 101, 5
Umemura, M., Loeb, A., & Turner, E. 1993, ApJ, 419, 459
Williams, R., & Perry, J. 1994, MNRAS, 269, 538
10
|
astro-ph/9601027 | 1 | 9601 | 1996-01-08T12:50:46 | Future Cosmic Microwave Background Constraints to the Baryon Density | [
"astro-ph"
] | We discuss what can be learned about the baryon density from an all-sky map of the cosmic microwave background (CMB) with sub-degree angular resolution. With only minimal assumptions about the primordial spectrum of density perturbations and the values of other cosmological parameters, such a CMB map should be able to distinguish between a Universe with a baryon density near 0.1 and a baryon-dominated Universe. With additional reasonable assumptions, it is conceivable that such measurements will constrain the baryon density to an accuracy similar to that obtained from BBN calculations. | astro-ph | astro-ph |
Future Cosmic Microwave Background Constraints to the
Baryon Density
M. Kamionkowski1, G. Jungman2, A. Kosowsky3,4, and D. N. Spergel5,6
1Department of Physics, Columbia University, New York, NY 10027
2Department of Physics, Syracuse University, Syracuse, NY 13244
3Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138
4Department of Physics, Harvard University, Cambridge, MA 02138
5Department of Astrophysical Sciences, Princeton University,
Princeton, NJ 08544
6Department of Astronomy, University of Maryland, College Park, MD
20742
Abstract. We discuss what can be learned about the baryon density
from an all-sky map of the cosmic microwave background (CMB) with
sub-degree angular resolution. With only minimal assumptions about the
primordial spectrum of density perturbations and the values of other cos-
mological parameters, such a CMB map should be able to distinguish be-
tween a Universe with a baryon density near 0.1 and a baryon-dominated
Universe. With additional reasonable assumptions, it is conceivable that
such measurements will constrain the baryon density to an accuracy sim-
ilar to that obtained from BBN calculations.
The current range for the baryon-to-photon ratio allowed by big-bang nu-
cleosynthesis (BBN) is 0.0075 ∼< Ωbh2
∼< 0.024 (Copi et al. 1995). This gives
Ωb ∼< 0.1 for the range of acceptable values of h, which implies that if Ω = 1,
as suggested by inflationary theory (or even if Ω ∼> 0.3 as suggested by cluster
dynamics), then the bulk of the mass in the Universe must be nonbaryonic. On
the other hand, X-ray -- cluster measurements might be suggesting that the ob-
served baryon density is too high to be consistent with BBN (see, e.g., Felten &
Steigman 1995 and references therein); this becomes especially intriguing given
the recent measurement of a large primordial deuterium abundance in quasar
absorption spectra (Hogan & Ruger 1995). The range in the BBN prediction
can be traced primarily to uncertainties in the primordial elemental abundances.
There is, of course, also some question as to whether the X-ray -- cluster measure-
ments actually probe the universal baryon density. For these reasons and more,
it would clearly be desirable to have an independent measurement of Ωbh2.
Here, we evaluate the precision with which the baryon-to-photon ratio,
Ωbh2, can be determined with high-resolution CMB maps (Bennett et al. 1995;
Janssen et al. 1995; Bouchet et al. 1995). We work within the context of mod-
els with adiabatic primordial density perturbations, although similar arguments
apply to isocurvature models as well, and we expect the power spectrum to dis-
1
tinguish clearly the two classes of models (Crittenden & Turok 1995). (More
details may be found in Jungman et al. 1995a,b.)
A given cosmological theory makes a statistical prediction about the dis-
tribution of CMB temperature fluctuations, expressed by the angular power
spectrum
C(θ) ≡ h[∆T ( m)/T0][∆T (n)/T0]i m·n=cos θ ≡ Xℓ
(2ℓ + 1)CℓPℓ(cos θ)/(4π),
(1)
where ∆T (n)/T0 is the fractional temperature perturbation in the direction n,
Pℓ are the Legendre polynomials, and the brackets represent an ensemble average
over all angles and observer positions. Since we can observe from only a single lo-
cation in the Universe, the observed multipole moments C obs
ℓ will be distributed
about the mean value Cℓ with a "cosmic variance" σℓ ≃ p2/(2ℓ + 1)Cℓ.
We consider an experiment which maps a fraction fsky of the sky with
a gaussian beam with full width at half maximum θfwhm and a pixel noise
σpix = s/√tpix, where s is the detector sensitivity and tpix is the time spent
observing each θfwhm × θfwhm pixel. We adopt the inverse weight per solid
angle, w−1 ≡ (σpixθfwhm/T0)2, as a measure of noise that is pixel-size inde-
pendent (Knox 1995). Current state-of-the-art detectors achieve sensitivities of
s = 200 µK√sec, corresponding to an inverse weight of w−1 ≃ 2 × 10−15 for a
one-year experiment. Realistically, however, foregrounds and other systematic
effects may increase the noise level; conservatively, w−1 will likely fall in the
range (0.9 − 4) × 10−14. Treating the pixel noise as gaussian and ignoring any
correlations between pixels, the C obs
ℓ will be distributed about the Cℓ with a
standard error
σℓ = [(2ℓ + 1)fsky/2]−1/2hCl + (wfsky)−1eℓ2σ2
bi ,
(2)
where σb = 7.4 × 10−3(θfwhm/1◦).
Given a spectrum of primordial density perturbations, the Cℓ are obtained
by solving the coupled equations for the evolution of perturbations to the space-
time metric and perturbations to the phase-space densities of all particle species
in the Universe. We consider models with initial adiabatic density perturba-
tions filled with photons, neutrinos, baryons, and collisionless dark matter; this
includes all inflation-based models. The CMB power spectrum depends upon
many parameters. Here, we include the following set: the total density Ω; the
Hubble constant, H0 = 100 h km sec−1 Mpc−1; the density of baryons in units
of the critical density, Ωbh2; the cosmological constant in units of the critical
density, Λ; the power-law indices of the initial scalar- and tensor-perturbation
spectra, nS and nT ; the amplitudes of the scalar and tensor spectra, parame-
terized by Q, the total CMB quadrupole moment, and r = Q2
S, the ratio of
the squares of the tensor and scalar contributions to the quadrupole moment;
the optical depth to the surface of last scatter, τ ; the deviation from scale in-
variance of the scalar perturbations, α ≡ dn/d ln k; and the effective number of
light-neutrino species at decoupling, Nν . Thus for any given set of cosmological
parameters, s = {Ω, Ωbh2, h, nS , Λ, r, nT , α, τ, Q, Nν}, we can calculate the mean
multipole moments Cℓ(s).
We now wish to determine the precision with which CMB maps will be
able to determine Ωbh2 without making any assumptions about the values of
T /Q2
2
Figure 1.
The standard error on Ωbh2.
the other undetermined parameters. The answer will depend on the measure-
ment errors σl, and on the underlying cosmological theory.
If the actual pa-
rameters describing the Universe are s0, then the probability distribution for
observing a CMB power spectrum which is best fit by the parameters s is
P (s) ∝ expn− 1
approximately by
2 (s − s0) · [α] · (s − s0)o where the curvature matrix [α] is given
(3)
αij = Xℓ
1
σ2
ℓ " ∂Cℓ(s0)
∂si
∂Cℓ(s0)
∂sj
#
with σℓ as given in Eq. (2). The covariance matrix [C] = [α]−1 is an estimate
of the standard errors that would be obtained from a maximum-likelihood fit
to data: the error in measuring the parameter si (obtained by integrating over
all the other parameters) is approximately C1/2
. If some of the parameters are
known, then the covariance matrix for the others is determined by inverting the
submatrix of the undetermined parameters.
ii
Fig. 1 displays the standard error in Ωbh2 as a function of the beam width
θfwhm for different noise levels and for fsky = 1. The underlying model as-
sumed here for the purpose of illustration is "standard CDM," given by s =
{1, 0.01, 0.5, 1, 0, 0, 0, 0, 0, QCOBE , 3}, where QCOBE = 20 µK is the COBE nor-
malization (G´orski et al. 1994). The solid curves show the C1/2
Ωbh2,Ωbh2 obtained by
inversion of the full 11× 11 curvature matrix [α] for w−1 = 2× 10−15, 9× 10−15,
and 4 × 10−14. These are the sensitivities that can be attained at the given
noise levels with the assumption of uniform priors (that is, including no infor-
mation about any parameter values from other observations). The dotted curves
show the C1/2
Ωbh2,Ωbh2 obtained by inversion of the Ωbh2-Q submatrix of [α]; this
is the error in Ωbh2 that could be obtained if all other parameters except the
3
normalization were fixed, either from other observations or by assumption. Re-
alistically, the precision obtained will fall somewhere between these two sets of
curves. The results for a mapping experiment which covers only a fraction fsky
of the sky can be obtained by replacing w → wfsky and scaling by f −1/2
[c.f.,
sky
Eq. (2)].
The implications of CMB maps for the baryon density depend quite sen-
minimal assumptions) at least be able to rule out a baryon-dominated Universe
sitively on the experiment. As long as θfwhm ∼< 0.5, the CMB should (with
(Ωb ∼> 0.3) and therefore confirm the predictions of BBN. With angular res-
olutions that approach 0.1◦ [which might be achievable, for example, with a
ground-based interferometry map (Myers 1995) to complement a satellite map],
a CMB map would provide limits to the baryon-to-photon ratio that were com-
petitive with BBN. Furthermore, if other parameters can be fixed, the CMB
might be able to restrict Ωbh2 to a small fraction of the range currently allowed
by BBN.
Moreover, the CMB will also provide information on several other param-
eters (Jungman et al. 1995a,b). Most significantly, the total density Ω can be
determined to better than 10% with minimal assumptions and perhaps better
than 1%.
Acknowledgments. This work was supported in part by the D.O.E. under
contracts DEFG02-92-ER 40699 and DEFG02-85-ER 40231, by the Harvard
Society of Fellows, by the NSF under contract ASC 93-18185, and by NASA
under contract NAG5-3091 and NAGW-2448, and by NASA under the MAP
Mission Concept Study Proposal.
References
Bennett, C. L. et al. 1995, NASA Mission Concept Study
Bouchet, F. R. et al. 1995, astro-ph/9507032
Copi, C. J., et al. 1995, astro-ph/9508029
Crittenden, R. G. & Turok, N. 1995, Phys. Rev. Lett., 75, 2642
Felten, J. E. & Steigman, G. 1995, in Proc. St. Petersburg Gamow Seminar,
St. Petersburg, Russia, 12 -- 14 September 1994, A. M. Bykov and A.
Chevalier, Sp. Sci Rev. (Dordrecht: Kluwer)
G´orski, K. M. et al. 1994, ApJ, 430, L89
Hogan, C. J. & Ruger, A. 1995, this proceedings
S. T. Myers, private communication
Janssen, M. A. et al. 1995, NASA Mission Concept Study
Jungman, G., Kamionkowski, M., Kosowsky, A., & Spergel, D. N. 1995a, astro-
ph/9507080, Phys. Rev. Lett., in press
Jungman, G., Kamionkowski, M., Kosowsky, A., & Spergel, D. N. 1995b, in
preparation
Knox, L. 1995, Phys. Rev. D, 52, 4307
4
|
0711.2686 | 2 | 0711 | 2008-02-12T13:53:54 | Observational constraints on late-time Lambda(t) cosmology | [
"astro-ph",
"gr-qc",
"hep-th"
] | The cosmological constant, i.e., the energy density stored in the true vacuum state of all existing fields in the Universe, is the simplest and the most natural possibility to describe the current cosmic acceleration. However, despite its observational successes, such a possibility exacerbates the well known cosmological constant problem, requiring a natural explanation for its small, but nonzero, value. In this paper we study cosmological consequences of a scenario driven by a varying cosmological term, in which the vacuum energy density decays linearly with the Hubble parameter. We test the viability of this scenario and study a possible way to distinguish it from the current standard cosmological model by using recent observations of type Ia supernova (Supernova Legacy Survey Collaboration), measurements of the baryonic acoustic oscillation from the Sloan Digital Sky Survey and the position of the first peak of the cosmic microwave background angular spectrum from the three-year Wilkinson Microwave Anisotropy Probe. | astro-ph | astro-ph |
Observational constraints on late-time Λ(t) cosmology
S. Carneiro∗1†, M. A. Dantas2‡, C. Pigozzo1§ and J. S. Alcaniz2¶
1Instituto de F´ısica, Universidade Federal da Bahia, Salvador - BA, 40210-340, Brasil
2Departamento de Astronomia, Observat´orio Nacional, Rio de Janeiro - RJ, 20921-400, Brasil
(Dated: November 7, 2018)
The cosmological constant Λ, i.e., the energy density stored in the true vacuum state of all
existing fields in the Universe, is the simplest and the most natural possibility to describe the current
cosmic acceleration. However, despite its observational successes, such a possibility exacerbates the
well known Λ problem, requiring a natural explanation for its small, but nonzero, value. In this
paper we study cosmological consequences of a scenario driven by a varying cosmological term, in
which the vacuum energy density decays linearly with the Hubble parameter, Λ ∝ H. We test
the viability of this scenario and study a possible way to distinguish it from the current standard
cosmological model by using recent observations of type Ia supernova (Supernova Legacy Survey
Collaboration), measurements of the baryonic acoustic oscillation from the Sloan Digital Sky Survey
and the position of the first peak of the cosmic microwave background angular spectrum from the
three-year Wilkinson Microwave Anisotropy Probe.
I.
INTRODUCTION
The nature of the mechanism behind the current cos-
mic acceleration constitutes a major problem nowadays
in Cosmology [1]. Even though almost all observational
data available so far are in good agreement with the sim-
plest possibility, i.e., a vacuum energy plus cold dark
matter (ΛCDM) scenario, it is becoming rather consen-
sual that in order to better understand the nature of the
dark components of matter and energy, one must also
consider more complex models as, for instance, scenarios
with interaction in the dark sector [2].
In this regard, the simplest examples of interacting
dark matter/dark energy models are scenarios with vac-
uum decay [Λ(t)CDM]. In reality, Λ(t)CDM cosmologies
constitute the special case in which the ratio of the dark
energy pressure to its energy density, ω, is exactly −1
[3, 4]. This kind of model may be based on the idea that
dark energy is the manifestation of vacuum quantum fluc-
tuations in the curved space-time, after a renormalization
in which the divergent vacuum contribution in the flat
space-time is subtracted. The resulting effective vacuum
energy density will depend on the space-time curvature,
decaying from high initial values to smaller ones as the
universe expands [5]. As a result of conservation of to-
tal energy, implied by Bianchi identities, the variation of
vacuum density leads either to particle production or to
an increasing in the mass of dark matter particles, two
general features of decaying vacuum or, more generally,
of interacting dark energy models [6].
Naturally, the precise law of vacuum density variation
depends on a suitable derivation of the vacuum contri-
∗ICTP Associate Member
†[email protected]
‡[email protected]
§[email protected]
¶[email protected]
bution in the curved background, which is in general a
difficult task.
In this regard, however, a viable possi-
bility is initially to consider a de Sitter space-time and
estimate the renormalized vacuum contribution with help
of thermodynamic reasonings as, e.g., those in line with
the holographic conjecture [7]. The resulting ansatz is
Λ ≈ (H + m)4 − m4, where H is the Hubble parameter
and m is a cutoff imposed to regularize the vacuum con-
tribution in flat space-time (the next step is to consider
a quasi-de Sitter background, with a slowly decreasing
H. In this case, the above ansatz may be considered a
good approximation, but with the vacuum density decay-
ing with time.). In the early-time limit, with H >> m,
we have Λ ≈ H 4. By using this scaling law for the vac-
uum density and introducing a relativistic matter com-
ponent, some of us have obtained from the Einstein equa-
tions an interesting solution with the following features
[8]. Firstly, the Universe undergoes an empty, quasi-de
Sitter phase, with H ≈ 1, which in the asymptotic limit
of infinite past tends to de Sitter solution with H = 1
(in Planck units). However, at a given time, the vacuum
undertakes a fast phase transition, with Λ decreasing to
nearly zero in a few Planck times, producing a consid-
erable amount of radiation.
In this sense, this can be
understood as a non-singular inflationary scenario, with
a semi-eternal quasi-de Sitter phase originating a radia-
tion dominated universe.
The subsequent evolution of vacuum energy essentially
depends on the masses of the produced particles. If we
apply the energy-time uncertainty relation to the pro-
cess of matter production, we will conclude that mas-
sive particles can be produced only at very early times,
when H is very high. Therefore, baryonic particles and
massive dark matter (as supersymmetric particles and
axions) stopped being produced before the time of elec-
troweak phase transition. On the other hand, the late-
time production of photons and massless neutrinos must
be forbidden by some selection rule, otherwise the Uni-
verse would be completely different from the one observed
today. Thus, if no other particle is produced, the vac-
uum density stops decaying at very early times, and for
late times we have a standard universe, with the pres-
ence of a genuine cosmological constant (some thermody-
namic considerations, in line with the holographic prin-
ciple, permit to infer the value of this constant, leading
to Λ ≈ m6 [5, 8]).
In order to have a decaying vacuum density at the
present time, the produced dark particles should have
a mass as small as 10−65g1. Here, we have considered
the possibility of a late-time decaying Λ, and compared
the consequent cosmological scenario with the constraints
imposed by current observations [10, 11]. In such a limit,
H << m and Λ ≈ m3H, leading again to Λ ≈ m6 in the
de Sitter limit. From a qualitative point of view, we have
found no important difference between this Λ(t) scenario
and the flat ΛCDM model [10]. After the phase tran-
sition described above the Universe enters a radiation-
dominated phase, followed by a matter-dominated era
long enough for structure formation, which tends asymp-
totically to a de Sitter universe with vacuum dominating
again. The only important novelty, related to matter
production, is a late-time suppression of the density con-
trast of matter, which may constitute a potential solution
to the cosmic coincidence problem [12]. Moreover, the
analysis of the redshift-distance relation for supernovae
of type Ia, particularly with the Supernova Legacy Sur-
vey (SNLS) data set, has shown a good fit, with present
values of H and the relative density of matter in accor-
dance with other observations [11].
In this paper, we go a little further in our investiga-
tion and study new observational consequences of the
Λ(t)CDM scenario described above. We use to this end
distance measurements from type Ia supernovae (SNe
Ia), measurements of the baryonic acoustic oscillations
(BAO) and the position of the first peak of the cosmic
microwave background (CMB). We show that, besides
the interesting cosmic history of this class of Λ(t)CDM
models, a conventional, spatially flat ΛCDM model is
only slightly favored over them by the current observa-
tional data.
II. THE MODEL
In a spatially flat FLRW space-time the Friedmann and
the conservation equations can be written, respectively,
as2
and
ρT = 3H 2 ,
ρT + 3H(ρT + pT ) = 0 ,
(1)
(2)
1 Some authors associate this mass to the graviton in a de Sitter
background with H and Λ of the orders currently observed [9].
2 We work in units where MPlanck = (8πG)−1/2 = c = 1.
2
where ρT and pT are the total energy density and pres-
sure, respectively. If we consider that the cosmic fluid is
composed of matter with energy density ρm and pressure
pm, and of a time-dependent vacuum term with energy
density ρΛ = Λ and pressure pΛ = −Λ, we obtain
ρm + 3H(ρm + pm) = − Λ ,
(3)
which shows that matter is not independently conserved,
with the decaying vacuum playing the role of a matter
source. Throughout our analysis we assume that baryons
are independently conserved at late times, being not pro-
duced at the expenses of the decaying vacuum. This
amounts to saying that we will postulate, in addition
to Eq.
(3), a conservation equation for baryons, i.e.,
ρb + 3H(ρb + pb) = 0, where ρb and pb refer to baryon
density and pressure.
From the above equations and considering our late-
time ansatz Λ = σH (σ is a positive constant of the
order of m3), the evolution equation reads
2 H + 3H 2 − σH = 0 .
(4)
Now, with the conditions H > 0 and ρm > 0, the
integration of the above equation leads to the following
expression for the scale factor,
a(t) = C [exp (σt/2) − 1]
3 ,
2
(5)
where C is the first integration constant and the second
one was chosen so that a = 0 at t = 0. From these equa-
tions, it is straightforward to show that at early times
the above expression reduces to the Einstein-de Sitter
solution whereas at late times it tends to the de Sitter
universe.
As stated earlier, Λ = σH and, therefore, ρm = 3H 2 −
σH. By using solution (5), one can also show that
and
ρm =
σ2C3
3a3 +
σ2C3/2
3a3/2
,
Λ =
σ2
3
+
σ2C3/2
3a3/2 .
(6)
(7)
The above expressions can be easily understood as fol-
lows. The first terms in the r.h.s. are the expected scaling
laws for matter density and the cosmological constant in
the case of a non-decaying vacuum while the second ones
are related to the time variation of the vacuum density
and the concomitant matter production. As expected,
at early times matter dominates, with its density scaling
with a−3, and the matter production process is negligi-
ble. On the other hand, at late times the vacuum term
dominates, as should be in a de Sitter universe.
From Eqs. (1), (6) and (7), the evolution of the Hubble
parameter as a function of the redshift can be written as
H(z) = H0h1 − Ωm + Ωm(1 + z)3/2i ,
(8)
where H0 and Ωm are, respectively, the present values of
the Hubble parameter and of the relative energy density
of matter. Note that, due to the matter production, this
expression is rather different from that found in the con-
text of the standard ΛCDM case. In particular, if Λ = 0
and Ωm = 1, we obtain H(z) = H0(1 + z)3/2, leading to
ρm = 3H 2
0 (1 + z)3, as expected for the Einstein-de Sitter
scenario.
[13], N = 115. As discussed in Ref.
[11], the best-fit
values for this analysis is obtained for h = 0.70 and Ωm =
0.33, with reduced χ2
r = 1.01. At 95% of confidence
level, we also find 0.69 < h < 0.71 and 0.28 < Ωm <
0.37. The 2-dimensional contours shown in the Figures
are obtained from the traditional frequentist confidence
intervals (based on the ∆χ2 approach and assuming that
errors are normally distributed).
III. OBSERVATIONAL CONSTRAINTS
B. Baryonic acoustic oscillations
3
Extending and updating previous results [11], we study
in this Section some observational consequences of the
class of Λ(t)CDM scenarios discussed above. Note that,
similarly to the standard ΛCDM case, this class of models
has only two independent parameters, H0 and Ωm [see
Eq. (8)]. The best-fit values for these quantities will be
determined on the basis of a statistical analysis of recent
type Ia supernovae (SNe Ia) measurements, as given by
the SNLS Collaboration [13], the distance to the baryonic
acoustic oscillations (BAO) from the Sloan Digital Sky
Survey (SDSS) [14], and the position of the first peak
in the spectrum of anisotropies of CMB radiation from
the three-year Wilkinson Microwave Anisotropy Probe
(WMAP) [15] (for more details on the statistical analysis
discussed below we refer the reader to Ref. [16]).
A. SNe Ia observations
The predicted distance modulus for a supernova at red-
shift z, given a set of parameters s, is
µp(zs) = m − M = 5 logdL + 25,
(9)
where m and M are, respectively, the apparent and ab-
solute magnitudes, the complete set of parameters is
s ≡ (H0, Ωm), and dL stands for the luminosity distance
(in units of megaparsecs),
dL = c(1 + z)Z 1
x′
dx
x2H(x; s)
,
(10)
with x′ = a(t)
a0
variable, and H(x; s) the expression given by Eq. (8).
= (1 + z)−1 being a convenient integration
We estimated the best fit to the set of parameters s by
using a χ2 statistics, with
χ2 =
N
Xi=1
p(zs) − µi
(cid:2)µi
o(z)(cid:3)2
σ2
i
,
(11)
(µB ) + σ2
p(zs) is given by Eq. (9), µi
where µi
o(z) is the extinction
corrected distance modulus for a given SNe Ia at zi, and
σ2
i = σ2
int, where σ2
(µB ) is the variance in the
individual observations and σ2
int stands for the intrinsic
dispersion for each SNe Ia absolute magnitude. Since
we use in our analysis the SNLS collaboration sample
The use of BAO to test dark energy models is usually
made by means of the parameter A, i.e., [14]
A = DV pΩmH 2
0
zc
,
(12)
where z = 0.35 is the typical redshift of the sample and
DV is the dilation scale, defined as
DV (z) = (cid:20)DM (z)2 zc
H(z)(cid:21)1/3
,
with the comoving distance DM given by
DM (z) = cZ z
0
dz′
H(z′)
.
(13)
(14)
An important aspect worth emphasizing at this point
is that the value of A is obtained from the data in the
context of a ΛCDM model, and can be considered a good
approximation only for some class of dark energy models
[17]. In particular, two conditions are implicitly assumed
to be valid [14]. First, the evolution of matter density
perturbations during the matter-dominated era must be
similar to the ΛCDM case, at least until the character-
istic redshift z = 0.35. Second, the comoving distance
to the horizon at the time of equilibrium between mat-
0 )−1. However,
ter and radiation must scale with (ΩmH 2
none of the above conditions are satisfied in the present
model because of the matter production associated to the
vacuum decay. As will be shown in a forthcoming pub-
lication [12], if matter is homogeneously produced there
is a suppression in its density contrast for z < 5, that is,
after the period of galaxy formation (this will eventually
imply a higher value of Ωm in order to fit the observed
mass power spectrum).
On the other hand, as radiation is independently con-
served, its relative energy density for z >> 1 is given
by Ωrz4, where Ωr is its present value. With the help
in the same limit z >> 1,
of (8) we can see that,
the relative density of matter is Ω2
mz3 (with the ex-
tra factor Ωm being due to the matter production be-
tween t(z) and the present time). By equating the
two densities, we obtain the redshift of equilibrium be-
tween matter and radiation, given by zrm = Ω2
m/Ωr.
Therefore, after including conserved radiation into (8)
(see equation (17) below) and taking z >> 1, we have
4
0.80
0.75
0.70
h
0.65
0.80
0.75
0.70
h
0.65
0.60
0.0
0.1
0.2
0.3
m,o
0.4
0.5
0.6
0.60
0.0
0.1
0.2
0.4
0.5
0.6
0.3
m,o
FIG. 1: Left) Superposition of the confidence regions in the Ωm − h plane for our analysis of SNe Ia, BAO and CMB. Right)
The same for a spatially flat ΛCDM model.
rH (zrm) ≈ cpρr/6(ΩmH0)−2, where ρr is the present
value of the radiation density (while in the ΛCDM case
we would obtain rH (zrm) ≈ cpρr/6(ΩmH 2
0 )−1, as stated
above). Thus, one can see that the parameter A is not
appropriate to test the model, and we will use instead the
dilation scale DV , which is weakly sensitive to the cos-
mological evolution before z = 0.35. By combining our
function H(z), given by (8), into (13)-(14) we can find
the region in the Ωm − H0 plane which gives the observed
value DV = (1370 ± 64) Mpc (1σ) [14]. The BAO bands
in the Ωm − H0 parametric space are shown in Fig. 1.
parameters at high redshifts, due to the process of matter
production. Therefore, in order to perform a correct test,
we have to explicitly calculate the acoustic scale in the
model and then compare with the measured position of
the peak.
The acoustic scale, defined as the ratio between the
comoving distance to the surface of last scattering and
the radius of the acoustic horizon at that time, is then
given by
C. The first peak of CMB
lA =
πR zls
R ∞
zls
0
cs
c
dz
H(z)
dz
H(z)
,
The two tests we have previously described depend on
the physics at low and intermediate redshifts (until z ∼ 1)
and will lead, as we will see, to very similar results. As a
complementary test, involving high-z measurements, let
us consider the position of the first peak in the spectrum
of CMB anisotropies. Since in the present model there
is no production of baryonic matter or radiation and the
spatial curvature is null, we expect that a correct posi-
tion of the first peak is enough to guarantee a spectral
profile similar to the ΛCDM one, provided the spectrum
of primordial fluctuations is the same.
In the context of a large class of dark energy models
this test is performed by comparing the predicted shift
parameter with the ΛCDM value. However, this is only
valid if the acoustic horizon at the time of last scattering
is the same [18]. This is not true in the present model
because, as we have already shown, for the same values of
H0 and Ωm we have different expressions for cosmological
where zls is the redshift of last scattering, and
cs = c(cid:18)3 +
9
4
Ωb
Ωγz(cid:19)−1/2
is the sound velocity. Here, Ωb and Ωγ are, respectively,
the present relative energy densities of baryons and pho-
tons.
The function H(z) to be used in Eq. (15) must now
include radiation. As this component is independently
conserved, scaling with a−4, the appropriate generaliza-
(15)
(16)
5
0.74
0.72
0.70
h
0.68
SNLS+BAO
SNLS+BAO+CMB
0.74
0.72
0.70
h
0.68
SNLS (dashed contours)
SNLS+BAO+CMB
SNLS+BAO
SNLS (dashed contours)
0.66
0.1
0.2
0.3
m,o
0.4
0.5
0.66
0.1
0.2
0.3
m,o
0.4
0.5
FIG. 2: Left) Confidence regions (68.3%, 95.4% and 99.7%) in the Ωm − h plane for a joint analysis involving SNe Ia, BAO
and CMB data. As indicated, the dashed lines stand for the SNe Ia results of Ref. [11]. In both panels, the largest contours
represent the joint analysis of SNe Ia plus BAO measurements whereas the smallest ones arises when the CMB data are included
in the analysis. Right) The same for a spatially flat ΛCDM model.
tion of Eq. (8) is given by3
H(z)
H0
≈ (cid:26)h1 − Ωm + Ωm(1 + z)3/2i2
+ Ωr(1 + z)4(cid:27)1/2
.
(17)
Therefore, apart from our two free parameters H0 and
Ωm, in order to determine the acoustic scale we need the
present values of the energy densities of radiation, pho-
tons and baryons, as well as an expression for zls. Since
radiation and baryons are independently conserved, and
we want to preserve the spectrum profile as well as the nu-
cleosynthesis constraints, we will take for these densities
the same values obtained from CMB observations in the
context of the ΛCDM model [19]: Ωγh2 ≈ 2.45 × 10−5,
Ωrh2 ≈ 4.1 × 10−5 and Ωbh2 ≈ 0.02.
means of the current fitting formula [18], and then we
impose that the optical depth has to be the same in both
models, that is,
Z zls
0
Γ(z)
H(z)
dz
1 + z
ls
= Z z∗
0
Γ(z)
H ∗(z)
dz
1 + z
,
(18)
where Γ is the rate of photon scattering and H ∗ is the
Hubble parameter in the ΛCDM context. For the current
intervals of H0 and Ωm, the relative differences between
zls and z∗
ls are typically as small as 1%.
For a scale invariant ΛCDM model with spectral index
n = 1, the position of the first peak after including the
effect of plasma driving is given by the fitting expression
[20]
Concerning zls, its value is not in principle the same
as in the ΛCDM case. To determine its value for a given
pair (Ωm,H0) we will proceed as follows. First, we obtain
the redshift of last scattering z∗
ls in the ΛCDM model by
where
l1 = lA(1 − δ1),
δ1 = 0.267(cid:16) r
0.3(cid:17)0.1
,
(19)
(20)
3 The inclusion of a conserved component of radiation changes the
dynamics and, consequently, the evolution of Λ(z) and ρm(z).
Thus, rigorously speaking the generalization of H(z) would re-
quire a reanalysis of the dynamics. Nevertheless, as Ωr ≈ 10−4,
when the decaying vacuum and matter production begin to be
important, the radiation term is negligible. For this reason, Eq.
(17) can be considered a good approximation. Indeed, a numeri-
cal analysis in the range 0 < z < 104 showed a difference between
Eq. (17) and the exact H(z) as small as 0.01%.
with r ≡ ργ(zls)/ρm(zls). Since the plasma driving de-
pends essentially on pre-recombination physics, we will
therefore assume that the above fitting formulae are a
good approximation to the present Λ(t) model. In our
case, the parameter r is given by
r =
Ωγ
Ω2
m
zls.
(21)
Finally, from the above results we can determine the
region of the Ωm-H0 plane for which the first peak has the
position currently observed by WMAP, i.e., l1 = 220.8 ±
0.7 (1σ) [15]. As in the case of BAO, the CMB bands in
the Ωm − H0 parametric space is shown in Fig. 1.
D. Results and discussions
The superposition of the confidence regions and bands
corresponding to our analysis of SNe Ia, BAO and CMB
observations is shown in Figure 1. For the sake of com-
parison, the right panel shows the same analysis for the
standard ΛCDM model. From these analysis, we note
that, although not very restrictive and parallel to the
SNe Ia contours, the BAO bands can be statistically
combined with SNe Ia data providing constraints on the
Ωm − H0 plane. At 2σ level, we find h = 0.70 ± 0.01 and
Ωm = 0.32 ± 0.05, with reduced χ2
r ≃ 1.00. The inclusion
of CMB data into our analysis in turn makes the com-
plete joint analysis more restrictive but also shows that
the concordance is as good as in the standard ΛCDM
case, and that the Λ(t)CDM model discussed here can-
not be ruled out.
It can be anticipated from Figure 1
that the data will prefer higher values of the matter den-
sity parameter, while the Hubble parameter is expected
to be slightly smaller than the current accepted values.
In order to confirm this qualitative discussion, Figure 2
shows the results of our joint statistical analysis (SNe Ia
+ BAO + CMB). At 2σ level, we find h = 0.69±0.01 and
Ωm = 0.36 ± 0.01, with reduced χ2
r = 1.01. For the sake
of comparison, the same analysis for the ΛCDM case is
also shown in the right panel of Fig. 2.
Finally, with the above best-fit values for the param-
eters, we can calculate the total expanding age for this
class of Λ(t)CDM models, given by [10, 11]
t0 = H −1
0
2
3 ln(Ωm)
Ωm − 1
≃ 15.0Gyr ,
(22)
as well as the redshift of transition from a decelerated
expansion to the current accelerating phase, i.e., zT ≈
1.3. Note that both values are slightly higher than, but
of the same order of, the standard ones.
IV. FINAL REMARKS
By using the most recent cosmological observations,
we have discussed the observational viability of a class
of Λ(t)CDM scenarios in which Λ ∝ H, as well as a
possible way to distinguish it from the standard ΛCDM
model in what concerns the general characteristics of the
predicted cosmic evolution. As discussed earlier, these
Λ(t)CDM models have some interesting features as, e.g.,
the association of dark energy with vacuum fluctuations,
the circumvention of the cosmological constant problem
by subtracting the flat space-time contribution from the
curved space-time vacuum density, and the possible (but
not necessary) link between dark matter and massive
gravitons.
6
We have presented some quantitative results which
clearly show that, even in the current stage of the Uni-
verse evolution, our decaying vacuum scenario is very
similar to the standard one. We have statistically tested
the viability of the model by using recent SNe Ia observa-
tions and measurements of the BAO and the first peak of
the CMB spectrum. At 95.4% c.l., a joint analysis involv-
ing SNe Ia + BAO provides the intervals 0.69 ≤ h ≤ 0.71
and 0.28 ≤ Ωm ≤ 0.37, which are in good agreement
with the values of the Hubble and the matter density
parameters obtained from independent analysis [21, 22].
When the position of the first peak of CMB anisotropies
is included in the analysis, the best-fit value for the rel-
ative matter density is increased, Ωm ≃ 0.36. This re-
sult cannot rule out the model, and it may be indicating
that, besides the interesting cosmic history of this class
of Λ(t)CDM models, a conventional, spatially flat ΛCDM
model is only slightly favored over them by the current
observational data. Still on the best-fit for Ωm, we note
that such a higher matter density is something to be more
investigated by means of other cosmological or dynam-
ical tests as, e.g., the predicted mass power spectrum
in the context of the model. As we have discussed ear-
lier, a higher matter density is necessary to compensate
the late-time suppression of the density contrast owing
to matter production [12]. Another possibility will be
provided by future supernovae observations, since the
present model starts to diverge from the standard one
for higher redshifts [11].
Finally, we should also emphasize two aspects to be
considered before a definite conclusion about the com-
parison between the present model and the flat standard
scenario. First, that in our study of CMB we have used
parameter values and fitting formulae that are strictly
correct for the ΛCDM case, particularly the expression
giving the position of the first peak for a given acoustic
scale [equation (19)]. In spite of our reasons to consider
that use as a good approximation, it can lead to bias
in our results.
If that is the case, only a more com-
plete analysis of CMB would rule out or corroborate the
model. The second aspect is of theoretical character. As
discussed in the Introduction, our ansatz for the variation
of Λ, if genuine, is a good approximation only for quasi-
de Sitter backgrounds. Naturally, our Universe, although
dominated by a cosmological term, is far away from the
asymptotic de Sitter state, with matter still giving an
important contribution to the cosmic fluid. Only a more
profound theoretical study, on the basis of quantum field
theories in the expanding background, could establish the
degree of applicability and limits of that approximation.
Acknowledgements
partially supported by CNPq and CAPES. JSA is also
supported by FAPERJ No. E-26/171.251/2004.
The authors are grateful to Deepak Jain, Max Tegmark
and Raul Abramo for useful references. This work is
7
[1] V. Sahni and A. A. Starobinsky, Int. J. Mod. Phys. D9,
373 (2000); P. J. E. Peebles and B. Ratra Rev. Mod.
Phys. 75, 559 (2003); T. Padmanabhan, Phys. Rept.
380, 235 (2003); E. J. Copeland, M. Sami and S. Tsu-
jikawa, Int. J. Mod. Phys. D15, 1753 (2006); J. S. Al-
caniz, Braz. J. Phys. 36, 1109 (2006).
[2] L. Amendola, Phys. Rev. D 62,
043511 (2000);
J. M. F. Maia and J. A. S. Lima, Phys. Rev. D 65,
083513 (2002); T. Koivisto, Phys. Rev. D 72, 043516
(2005); S. Lee, G. C. Liu and K. W. Ng, Phys. Rev.
D 73, 083516 (2006); M. C. Bento, O. Bertolami and
N. M. C. Santos, arXiv:gr-qc/0111047; O. Bertolami,
F. Gil Pedro and M. Le Delliou, Phys. Lett. B 654, 165
(2007); F. E. M. Costa, J. S. Alcaniz and J. M. F. Maia,
arXiv:0708.3800 [astro-ph].
[3] M. Ozer and M. O. Taha, Phys. Lett. B171, 363 (1986);
Nucl. Phys. B287, 776 (1987); O. Bertolami, Nuovo Ci-
mento Soc. Ital. Fis. B93, 36 (1986).
[4] K. Freese et al, Nucl. Phys. B287, 797 (1987); W.
Chen and Y-S. Wu, Phys. Rev. D41, 695 (1990); M.
S. Berman, Phys. Rev. 43, 1075 (1991); D. Pav´on, Phys.
Rev. D43, 375 (1991); J. C. Carvalho, J. A. S. Lima and
I. Waga, Phys. Rev. D46 2404 (1992); A. I. Arbab and
A. M. M. Abdel-Rahman, Phys. Rev. D50, 7725 (1994);
J. A. S. Lima and J. M. F. Maia, Phys. Rev D49, 5597
(1994); J. A. S. Lima and M. Trodden, Phys. Rev. D53,
4280 (1996); J. M. Overduin and F. I. Cooperstock, Phys.
Rev. D58, 043506 (1998); J. M. Overduin, Astrophys. J.
517, L1 (1999); J. Shapiro and J. Sola, Phys. Lett B475,
236 (1999); M. V. John and K.B. Joseph, Phys. Rev.
D61, 087304 (2000); O. Bertolami and P. J. Martins,
Phys. Rev. D61, 064007 (2000); R. G. Vishwakarma,
Gen. Rel. Grav. 33, 1973 (2001); A. S. Al-Rawaf, Mod.
Phys. Lett. A14, 633 (2001); R. Schutzhold, Phys. Rev.
Lett. 89, 081302 (2002); Int. J. Mod. Phys. A17, 4359
(2002); M. K. Mak, J. A. Belinchon, and T. Harko, IJMP
D11, 1265 (2002); I. L. Shapiro and J. Sola, JHEP 0202,
006 (2002); W. Zimdahl and D. Pav´on, Gen. Rel. Grav.
35, 413 (2003); M. R. Mbonye, IJMP A18, 811 (2003);
J. S. Alcaniz and J. M. F. Maia, Phys. Rev. D67, 043502
(2003); I. L. Shapiro, J. Sola, C. Espana-Bonet, and
P. Ruiz-Lapuente, Phys. Lett. B574, 149 (2003); J. V.
Cunha and R. C. Santos, IJMP D13, 1321 (2004); R.
Opher and A. Pellison, Phys. Rev. D70, 063529 (2004);
R. Horvat, Phys. Rev. D70, 087301 (2004); C. Espana-
Bonet et al. JCAP 0402, 006 (2004); P. Wang and X.
Meng, Class. Quant. Grav. 22, 283 (2005); S. Carneiro
and J. A. S. Lima, IJMP A20 2465 (2005); S. Carneiro,
IJMP D14, 2201 (2005); I. L. Shapiro, J. Sola, and H.
Stefancic, JCAP 0501, 012 (2005); E. Elizalde, S. Nojiri,
S. D. Odintsov, and P. Wang, Phys. Rev. D71, 103504
(2005); R. Aldrovandi, J. P. Beltr´an Almeida, and J. G.
Pereira, Grav. & Cosmol. 11, 277 (2005); F. Bauer, Class.
Quant. Grav. 22, 3533 (2005); B. Wang, Y. Gong, and
E. Abdalla, Phys. Lett. B624, 141 (2005); J. Sola and
H. Stefancic, Mod. Phys. Lett. A21, 479 (2006); J. D.
Barrow and T. Clifton, Phys. Rev. D73, 103520 (2006);
B. Wang, C. Y. Lin, and E. Abdalla, Phys. Lett. B637,
357 (2006); A. E. Montenegro Jr. and S. Carneiro, Class.
Quant. Grav. 24, 313 (2007).
[5] S. Carneiro. Int. J. Mod. Phys. D12, 1669 (2003).
[6] J. S. Alcaniz and J. A. S. Lima, Phys. Rev. D72, 063516
(2005).
[7] R. Bousso, Rev. Mod. Phys. 74, 825 (2002); T. Padman-
abhan, Phys. Rept. 406, 49 (2005).
[8] S. Carneiro, Int. J. Mod. Phys. D15, 2241 (2006); J.
Phys. A40, 6841 (2007).
[9] See, for example, M. Novello and R. P. Neves, Class.
Quant. Grav. 20, L67 (2003).
[10] H. A. Borges and S. Carneiro, Gen. Rel. Grav. 37, 1385
(2005).
[11] S. Carneiro, C. Pigozzo, H. A. Borges and J. S. Alcaniz,
Phys. Rev. D74, 23532 (2006) [arXiv:astro-ph/0605607].
[12] H. A. Borges, S. Carneiro, J. C. Fabris and C. Pigozzo,
arXiv:0711.2689 [astro-ph], to appear in Phys. Rev. D.
[13] P. Astier et al., Astron. Astrophys. 447, 31 (2006).
[14] D. J. Eisenstein et al., Astrophys. J. 633, 560 (2005).
[15] G. Hinshaw et al., Astrophys. J. Suppl. 170, 288 (2007);
D. N. Spergel et al., Astrophys. J. Suppl. 170, 377 (2007).
[16] T. Padmanabhan and T. R. Choudhury, Mon. Not.
R. Astron. Soc. 344, 823 (2003); P. T. Silva and
O. Bertolami, Astrophys. J. 599, 829 (2003); Z.-
H. Zhu and M.-K. Fujimoto, Astrophys. J. 585, 52
(2003);
J. S. Alcaniz, Phys. Rev. D 65, 123514
(2002) [arXiv:astro-ph/0202492]; S. Nesseris and L.
Perivolaropoulos, Phys. Rev. D 70, 043531 (2004); J. S.
Alcaniz and N. Pires, Phys. Rev. D 70, 047303 (2004)
[arXiv:astro-ph/0404146]; J. S. Alcaniz, Phys. Rev. D 69,
083521 (2004) [arXiv:astro-ph/0312424]; T. R. Choud-
hury and T. Padmanabhan, Astron. Astrophys. 429, 807
(2005); J. S. Alcaniz and Z.-H. Zhu, Phys. Rev. D 71,
083513 (2005) [arXiv:astro-ph/0411604]; D. Rapetti, S.
W. Allen, and J. Weller, Mon. Not. R. Astron. Soc.
360, 555 (2005); Z. K. Guo, Z. H. Zhu, J. S. Al-
caniz and Y. Z. Zhang, Astrophys. J. 646, 1 (2006)
[arXiv:astro-ph/0603632]; L. Samushia and B. Ratra, As-
trophys. J. 650, L5 (2006); M. A. Dantas, J. S. Alcaniz,
D. Jain and A. Dev, Astron. Astrophys. 467, 421 (2007)
[arXiv:astro-ph/0607060].
[17] See, for instance, M. Doran, S. Stern and E. Thommes,
JCAP 0704, 015 (2007).
[18] O. Elgaroy and T. Multamaki, astro-ph/0702343.
[19] P. de Bernardis et al., Nature (London) 404, 955 (2000);
S. Hanany et al., Astrophys. J. Lett. 545, L5 (2000); S.
Padin et al., ibid, 549, L1 (2001); D. N. Spergel et al.,
Astrophys. J. Suppl. 148, 175 (2003).
[20] W. Hu et al., Astrophys. J. 549, 669 (2001).
[21] W. L. Freedman et al., Astrophys. J. 553, 47 (2001).
[22] H. A. Feldman, R. Juszkiewicz, P. Ferreira et al., Astro-
phys. J. 596, L131 (2003).
|
astro-ph/0410184 | 1 | 0410 | 2004-10-07T09:55:58 | A new multiparametric topological method for determining the primary cosmic ray mass composition in the knee energy region | [
"astro-ph"
] | The determination of the primary cosmic ray mass composition from the characteristics of extensive air showers (EAS), obtained at an observation level in the lower half of the atmosphere, is still an open problem. In this work we propose a new method of the Multiparametric Topological Analysis and show its applicability for the determination of the mass composition of the primary cosmic rays at the PeV energy region. | astro-ph | astro-ph | A new Multiparametric Topological
method for determining the primary
cosmic ray mass composition in the knee
energy region
M.Ambrosio (1), C.Aramo (1,2), D.D'Urso (3), A.D.Erlykin (1,4),
F.Guarino (1,2), A.Insolia (3)
(1) INFN, section of Napoli, Napoli, Italy
(2) Dept. of Physical Sciences, Univ. of Napoli Federico II, Napoli, Italy
(3) Dept. of Physics and Astronomy, Univ. of Catania and INFN, Section
of Catania, Catania, Italy
(4) P.N.Lebedev Physical Institute, Moscow, Russia
Abstract
The determination of the primary cosmic ray mass composition from
the characteristics of extensive air showers (EAS), obtained at an ob-
servation level in the lower half of the atmosphere, is still an open
problem. In this work we propose a new method of Multiparametric
Topological Analysis and show its applicability for the determination
of the mass composition of the primary cosmic rays at the PeV energy
region.
4
0
0
2
t
c
O
7
1
v
4
8
1
0
1
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1E-mail address: [email protected]
1
1
Introduction
It is clear that due to large fluctuations in the longitudinal development of
extensive air showers (EAS) and relatively weak sensitivity of EAS char-
acteristics to the mass of the parent particle, the observed parameters of
showers initiated by particles with different primary masses largely overlap.
In order to estimate the primary mass in the case of the individual shower or
to estimate the mean mass composition at a certain energy we have to use
as many observables as possible. To process the big amount of information
various multiparametric and non-parametric methods are employed. Among
them are multivariate fitting, KNN-method, Bayesian approach, neural net
analysis and others. In [1] we proposed a new method of the multiparametric
topological analysis (MTA) for the study of the mean mass composition on
the basis of the EAS longitudinal development data. We tested this method
using the data simulated in the EeV energy region for the Pierre Auger ex-
periment and were encouraged by its accuracy and the easy applicability.
In this contribution we test this method at the lower, PeV energy region
close to the well known 'knee' in the primary cosmic ray energy spectrum
(∼3 PeV). Experimental arrays for the EAS study in this energy region are
usually more compact and are able to measure the total size of electron Ne,
muon Nµ, hadron Nh and Cherenkov light Nph components of the shower.
The typical example of such a complex EAS array is KASCADE [2]. Here
we use only two parameters: Ne and Nµ, although the method is easily
generalized for the larger number of observables.
2 The simulated data
In what follows, we assume that the primary energy estimate for the ob-
served events is accurate at a few percent level. In fact, the primary energy
estimator for KASCADE is N tr
µ - the total number of low energy (> 0.23
GeV) muons collected in the interval of distances between 40 and 200 m from
the shower core. It has been shown that this number depends just on the
primary energy and not on the primary mass. The accuracy of the primary
energy estimate in the knee region made with this parameter is determined
mainly by the N tr
µ reconstruction error and is about 5% [3]
The data set used to implement and test the new method consists of
20000 vertical cascades produced by particles with the fixed energy of 0.5
PeV, 8000 cascades for 1 PeV and 2000 cascades for 5 PeV, simulated using
the CORSIKA program (version 6.023, [4]) with the QGSJET interaction
model. Simulations were performed using Naples PCs . The primary nuclei
were P , He, O and F e, each of them initiating 5000, 2000 and 500 cascades
for 0.5, 1 and 5 PeV primary energy respectively. The CORSIKA output
provides various characteristics of the shower: the total number of different
particles, their lateral distribution and arrival times at different observation
levels.
3 Multiparametric Topological Analysis (MTA)
The MTA method, when applied to the simulated data, relies on a topolog-
ical analysis of correlations between the most significant parameters of the
shower development in the atmosphere. In principle this method could be
used also with a greater number of parameters, however, in this paper we
restrict ourself to the simple case of two parameters only: logNe and logNµ
- the total number of electrons and muons at the sea level. A scatter plot of
these two parameters has been built using the 4 × 4000 showers for 0.5 PeV,
4 × 1000 for 1 PeV and 4 × 400 for 5 PeV primary energy. Figure 1 shows
the scatter plot for only proton and iron induced showers of 1 PeV primary
energy as well as their projected distributions.
It can be seen that the populations arising from two nuclei are quite
well separated in the logNµ − logNe plane, but not in the projected one-
dimensional plots. The scatter plot of four primary nuclei: P, He, O and F e
shows much less separation even in the two-dimensional plot. The idea of
the MTA method is to divide the entire area of the scatter plot into cells
whose dimensions are defined by the accuracy with which the parameters
can be measured. For this work, we adopted the accuracy and the size of the
cell equal to ∆(logNµ) = 0.05 and ∆(logNe) = 0.02, which corresponds to
the experimental errors of KASCADE array [2]. In each cell we can find the
total number of showers Ni
F e showers
induced by P, He, O and Fe respectively, and then derive the associated
frequencies: pi
tot
which can be interpreted as the probability for a real shower falling into
the ith cell to be initiated by proton, helium, oxygen or iron primary nuclei.
In other words, in the case of an experimental data set of Nexp showers,
tot, as the sum of Ni
O and Ni
O=Ni
O/Ni
F e=Ni
F e/Ni
He=Ni
He/Ni
P , Ni
He, Ni
tot and pi
P =Ni
P /Ni
tot, pi
tot, pi
Figure 1: (a) logNµ vs.
logNe scatter plot for proton (star) and iron (circle) induced
showers at 1 PeV primary energy. b) Projected distributions of the scatter plot in (a) for
protons (dashed line) and iron nuclei (full line)
it may be seen as composed of a mixture of Nexp× pP proton showers,
Nexp× pHe helium showers, Nexp× pO oxygen showers and Nexp× pF e iron
induced showers, where pA = Σipi
A. As a probe set of 'experimental' showers
4 subsets of showers, different from those used for the determination of
probabilities pA, have been used. Each subset consists of 500 showers for
0.5 PeV, 200 showers for 1 PeV and 50 showers for 5 PeV primary energy.
For each individual shower in a given subset the partial probabilities pi
P ,
pi
He , pi
F e have been read from the relevant cell i. The sum of such
probabilities over the entire subset permits to estimate the probability for
a shower of a given nature to be identified as a shower generated by the P ,
He, O or F e primary particle. This probability is shown in Figure 2 at E =
1 PeV. One can see that the method works quite well being able to attribute
the highest probability to the correct nuclei.
O and pi
4 Determination of the primary mass composition
The obtained mean probabilities Pij for cascades induced by type i nuclei
to be identified as those induced by j nuclei for the pure primary mass
composition can be used for the reconstruction of the mixed primary mass
composition as the coefficients in the system of linear equations:
Figure 2: Application of the MTA method to the logNµ − logNe scatter plot - mean
probability Pij for the cascades induced by i nucleus: P, He, O and Fe (specified in
headers) being identified as induced by j nucleus, indicated at abscissa, at E = 1 PeV.
n′
P = nP · PP →P + nHe · PHe→P + nO · PO→P + nF e · PF e→P
n′
He = nP · PP →He + nHe · PHe→He + nO · PO→He + nF e · PF e→He
n′
O = nP · PP →O + nHe · PHe→O + nHe · PO→O + nHe · PF e→O
n′
F e = nP · PP →F e + nHe · PHe→F e + nO · PO→F e + nF e · PF e→F e
where nP , nHe, nO and nF e are the true numbers, which determine the pri-
mary mass composition in the sample of N = nP + nHe + nO + nF e cascades,
which are observed as n′
F e, due to a misclassification.
P , n′
He, n′
O and n′
In order to invert the problem and to reconstruct the abundances nAi
in the primary mass composition from the observed abundances n′
Ai and
the known probabilities Pij, we can apply any method capable to solve the
inverse problem taking into account possible errors of the observed distri-
bution and the constraint:
Σ4
i=1nAi = Σ4
i=1n′
Ai = N
(1)
The observed abundances were simulated using a set of 4×800 cascades
different from those used for the determination of Pij. At the moment the
result of the solution for the uniform primary mass composition at E = 1
PeV, with all abundances pA = nA
N = 0.25, reconstructed by MTA method
is pP = 0.26, pHe = 0.22, pO = 0.27 and pF e = 0.25.
As an example of the non-uniform mass composition we have taken the
case with pP = pF e = 0.1, pHe = pO = 0.4, which resembles the mass
composition of cosmic rays at the knee energy of ∼3 PeV [5]. The total
amount of cascades used for the simulation of this distribution is 2000 (200
P + 800 He+ 800 O +200 F e). The result of the solution for this case is
pP = 0.14, pHe = 0.34, pO = 0.42 and pF e = 0.10 at E = 1 PeV .
It is seen that absolute values of the abundance are reconstructed satis-
factorily.
For the time being the values of the probabilities Pij are assumed to be
known precisely. A more accurate solution of the system, taking propely
into account the errors of the probability, is under study.
5 Application for experimental cascades
It has to be stressed that all the results outlined above are biased by the
fact that we are dealing with simulated 'noiseless' data and more realistic
testing has to be performed using data which take into account the varying
primary energy and inclination angle, the instrumental signature (noise)
and the various sources of errors (uneven and incomplete sampling, etc.).
Application of MTA is straightforward - the relevant simulations should be
made with the largest possible statistics.
The preliminary study of the effect of the experimental errors shows
rather strong dependence of the probabilities Pij on the adopted errors.
In Figure 3 we show the probabilities Pii for the correct identification of
nuclei as a function of the primary energy. The experimental accuracies
of the shower size determination have been taken as ∆(logNe) = 0.02 and
∆(logNµ) = 0.05. It is seen that while the energy dependence is week the
reduction of probability Pii is substantial.
Figure 3: The probability of the correct identification of the primary nucleus Pii as
the function of its energy without (full circles) and with (open circles) an account of
experimental errors: ∆(logNe) = 0.02 and ∆(logNµ) = 0.05. The primary nuclei are:
(a)-P ; (b)-He ; (c)-O ; (d)-Fe.
6 Conclusions
We proposed and tested the Multiparametric Topological Analysis method
for the determination of the mean primary cosmic ray mass composition
on the basis of measurements of just two parameters of the observed EAS
- the total number of electrons and muons. Definitely the method needs
further development before being used for the processing of real experimental
cascades, but the first results are very encouraging. The next efforts should
include the increase of simulation statistics, the study of the effect of the
shower energy spectrum and angular distribution, the accurate estimate
of the reconstruction errors. The great advantage of the proposed MTA
method is that it is extremely easy to use and generalize for the larger
number of observables.
Acknowledgements
Authors thank K.H.Kampert and L.Perrone for useful discussions.
References
[1] Ambrosio M. et al., 2004, Talk at CRIS2004, to be published in
Nucl.Phys.B (Proc.Suppl.), 2004
[2] Antoni T. et al., 2003, Nucl. Instr. Meth.,A513, 490
[3] Weber J. et al., 1997, 25th ICRC, Durban, 6, 153
[4] Heck D. et al. 1998, FZKA Report Forschungszentrum Karlsruhe 6019
[5] Kampert K.-H. et al., 2004, astro-ph/0405608
|
astro-ph/0305183 | 1 | 0305 | 2003-05-12T09:29:50 | Jets and Black Holes in Hypernova Explosions | [
"astro-ph"
] | Bipolar explosion models for hypernovae (very energetic supernovae), associated with the black hole formation, are presented. The model features are as follows: (1) Fe-peak elements are ejected at high velocities to reach the surface region, while dense O-rich materials occupy the central region at low velocities. (2) The abundance ratios (Zn, Co)/Fe are enhanced while the ratios (Mn, Cr)/Fe are suppressed. The overall abundances are consistent with those observed in extremely metal-poor stars formed at the very early phase of the Galaxy, suggesting significant contribution of such explosions to the early Galactic chemical evolution. (3) Certain relations among the optical luminosity, the chemical composition, and the property of the central remnant are predicted. Such relations may serve for identifying possible gravitational wave emitters in the future. | astro-ph | astro-ph |
Progress of Theoretical Physics Supplement
1
Jets and Black Holes in Hypernova Explosions
Keiichi Maeda∗) and Ken'ichi Nomoto
Department of Astronomy, University of Tokyo, Tokyo 113-0033, Japan
Bipolar explosion models for hypernovae (very energetic supernovae), associated with
the black hole formation, are presented. The model features are as follows: (1) Fe-peak ele-
ments are ejected at high velocities to reach the surface region, while dense O-rich materials
occupy the central region at low velocities. (2) The abundance ratios (Zn, Co)/Fe are en-
hanced while the ratios (Mn, Cr)/Fe are suppressed. The overall abundances are consistent
with those observed in extremely metal-poor stars formed at the very early phase of the
Galaxy, suggesting significant contribution of such explosions to the early Galactic chemical
evolution. (3) Certain relations among the optical luminosity, the chemical composition, and
the property of the central remnant are predicted. Such relations may serve for identifying
possible gravitational wave emitters in the future.
§1.
Introduction
Discovery of a new class of supernovae (SNe), called 'hypernovae', which have
distinctly large kinetic energies (E51 = E/1051ergs ∼> 10) compared with previously
known supernovae (E51 ∼ 1) is one of the most exciting topics in recent studies on
supernovae. For a review of hypernovae, see Refs. 9), 10).
Modeling spectra and light curves of these objects has uncovered an interest-
ing relation between progenitor's masses (MMS) and outcomes (E, M (56Ni)). (See
Ref. 10) and references therein.) There is an apparent transition around 20−25M⊙ in
these properties, from normal SNe to two branches of hypernovae and faint SNe. The
modeling of hypernovae suggests that E and M (56Ni) are larger for larger MMS in
this 'hypernova branch'. The transition implies that the explosion mechanism of stars
with MMS ∼> 20 − 25M⊙ might be different from that of stars with MMS ∼< 20M⊙,
especially for hypernovae given their large energies. (For faint SNe, the small E may
simply be attributed to a large gravitational binding energy.)
In this paper, we present bipolar explosion models for hypernovae. We describe
the hydrodynamic and nucleosynthetic results, and predict how the nucleosynthetic
yield, the optical luminosity, and the property of the central remnant are related.
§2. Hydrodynamics
The main ingredient of our models is a pair of jets propagating through a stellar
mantle. At the beginning of each calculation, the central part (Mr ≤ MREM0) of
a progenitor star is displaced by a compact remnant. The jets are injected at the
interface with the opening half-angle being θjet. The energy injected by the jets is
Ejet = α Maccretionc2). The
assumed to be proportional to the accretion rate (i.e.,
models are summarized in Table 1. The outcome of the explosion depends on the
∗) E-mail: [email protected]
2
K. Maeda and K. Nomoto
Table I. Models and Results. Masses are in solar mass unit (M⊙), and θjet is in degree. Abundance
ratios are normalized at the solar values, i.e., [X/Y] ≡ log 10(X/Y) − log 10(X/Y)⊙.
Model MMS MREM0
40A
40B
40C
40D
25A
25B
1.5
1.5
1.5
3.0
1.0
1.0
40
40
40
40
25
25
α
0.01
0.01
0.05
0.01
0.01
0.01
θjet Etotal MREM M (56Ni)
1.07E-1
15
8.11E-2
45
2.40E-1
15
15
6.28E-8
7.81E-2
15
45
1.51E-1
10.9
1.2
32.4
8.5
6.7
0.6
5.9
6.8
2.9
10.5
1.9
1.5
[S/Si]
-0.46
-0.54
-0.30
-1.1
-0.28
-0.26
[C/O]
-1.3
-1.2
-1.3
-1.0
-0.80
-0.82
Fig. 1. Velocity distribution of Model 40A at 1.5 second after the initiation of the jets. The right
panel shows that in the central region on an expanded scale. The reference arrow at the upper
right represents 2 × 109cm s−1.
interaction between the jets and the stellar mantle, and on the accretion rate which
is affected by the interaction itself. We follow the hydrodynamic evolution and
nucleosynthesis in two dimensions.
Figure 1 shows a snapshot of the velocity distribution of Model 40A at 1.5
seconds after the initiation of the jets. The jets propagate through the stellar mantle,
depositing their energies into ambient matter at the working surface. The bow shock
expands laterally to push the stellar mantle sideways, which reduces the accretion
rate. The strong outflow occurs along the z-axis (the jet direction), while matter
accretes from the side. As the accretion rate decreases, the jets are turned off. Then
the inflow along the r-axis turns to the weak outflow.
The outcome is a highly aspherical explosion. The accretion forms a central
dense core. Densities near the center become much higher than those in spherical
models. This feature is indeed what is suggested by the spectroscopic6) and light
curve8), 3) modeling of hypernovae.
Other hydrodynamic properties are as follows (Table 1): (1) A more massive
star makes a more energetic explosion. The reason is that a more massive star has
Jets and Black Holes in Hypernova Explosions
3
Fig. 2. Left: Distributions of 56Ni (which decays into 56Fe: filled circles) and 16O (dots). The
mass elements in which the mass fraction of each isotope exceeds 0.1 are plotted. Right: Mass
fractions of selected isotopes in the velocity space along the z-axis of Model 40A.
a stronger gravity to make the accretion rate higher. This is consistent with the
relation seen in the hypernova branch.10) (2) A more massive star forms a more
massive compact remnant. The remnant's mass increases as the accretion feeds
it. The final mass MREM reaches typically > 5M⊙ for a 40M⊙ star, and ∼ 2M⊙
for a 20M⊙ star. The bipolar models provide the way of explosions with black hole
formation in a consistent manner. Given the discovery of the evidence of a hypernova
explosion that accompanied formation of a black hole of ∼ 5M⊙ (X-ray Nova Sco;
Refs 1), 11)), it offers an interesting possibility.
§3. Nucleosynthesis
Relatively high temperatures along the z-axis and low temperatures along the
r-axis have significant effects on nucleosynthesis. It results in highly aspherical dis-
tribution of nucleosynthetic products as shown in Figure 2. The distribution of 56Ni
(which decays into 56Fe) is elongated along the z-axis. Such a configuration, i.e.,
high velocity Fe and low velocity O, has been suggested to be responsible for the
feature in the late phase spectra of SN1998bw, where the OI] 6300 was narrower
than the FeII] 5200 blend.7), 2)
Along the z-axis, heavy isotopes which are produced with high temperatures
T9 ≡ T /109K ∼> 5 are blown up to the surface. As a result, 64Ge, 59Cu, 56Ni, 48Cr,
and 44Ti (which decay into 64Zn, 59Co, 56Fe, 48Ti, and 44Ca, respectively) are ejected
at the highest velocities. Isotopes which are synthesized with somewhat lower tem-
peratures (T9 = 4 − 5) are first pushed aside as the jets propagate, then experience
circulation to flow into behind the working surface. 55Co, 52Fe (which decay into
55Mn and 52Cr, respectively), 40Ca, 32S, and 28Si are therefore ejected at the inter-
mediate velocities. Isotopes which are not synthesized but are only consumed during
4
K. Maeda and K. Nomoto
O Ne
Na Al
Mg
Si
S
P
Ti
Ar Ca
Cr Fe
Ni
Zn
Cl
Sc
K
Co
Mn
V
Cu
Na
O
Ne
Mg Si
Ca
Ti
Cr
Fe
Mn
P
Ar
S
Cl
K
Sc
V
Ni
Co
C
C
Fig. 3.
Isotropic yields of the bipolar model 40A (upper) and a spherical model (lower) with E51 =
10, M (56Ni) = 0.1M⊙, and MMS = 40M⊙.
the explosion are accreted from the side. 24Mg, 16O, and 12C occupy the innermost
region at the lowest velocities.
The distribution of isotopes as a function of the velocity shows inversion as com-
pared with conventional spherical models. This affects the overall abundance pat-
terns in the whole ejecta as shown in Figure 3. As noted above, materials which expe-
rience higher T9 are preferentially ejected along the z-axis, while materials with lower
T9 accrete from the side in the bipolar models. Zn and Co are ejected at higher veloc-
ities than Mn and Cr, so that the latter accrete onto the central remnant more easily.
As a consequence, [Zn/Fe] and [Co/Fe] ([X/Fe] ≡ log 10(X/Fe) − log 10(X/Fe)⊙) are
enhanced, while [Mn/Fe] and [Cr/Fe] are suppressed.
Finally, we predict the relation among the optical luminosity, nucleosynthetic
yield, and MREM in the bipolar models. Figure 4 shows M (56Ni) as a function of
MREM. For a given progenitor, they show a clear anti-correlation. As the optical
luminosity is basically proportional to M (56Ni), we predict a smaller MREM for a
brighter supernova in the context of the bipolar explosion. The ratios among various
elements depend on MREM. For example, [O/C] and [S/Si] are smaller for larger
MREM of the same progenitor.
§4. Conclusions
We have presented hydrodynamic and nucleosynthetic features of the bipolar
models. We have discussed advantages of our models in explaining the observed
features of hypernovae, which conventional spherical models fail to reproduce.
Jets and Black Holes in Hypernova Explosions
5
Fig. 4. The mass of ejected 56Ni as a function of MREM.
The bipolar models predict the following unique features. (1) Iron peak elements
(e.g., Zn, Co, Fe) are blown up to the surface. The elements near the surface may be
easily accelerated and ejected as cosmic rays. (2) (Zn, Co)/Fe are enhanced, while
(Mn, Cr)/Fe are suppressed. These trends are seen in the abundances in extremely
metal-poor stars, which likely lock up the information on chemical composition of
supernovae at the earliest phase of the Galaxy. This might indicate a significant
contribution of bipolar supernovae/hypernovae to the early Galactic chemical evo-
lution.4), 5), 10) (3) Certain relations among the optical luminosity, chemical compo-
sition, and MREM are predicted. As MREM is probably an indicator of gravitational
emissions from the event, we expect the relation among the optical luminosity, chem-
ical composition, and gravitational wave. Such relations may serve for identifying
gravitational wave emitters in the future, through observations by different methods.
References
1) G. Israelian, R. Rebolo, G. Basri et al., Nature 401 (1999), 142
2) K. Maeda, T. Nakamura, K. Nomoto, P. Mazzali, F. Patat, I. Hachisu, Astrophys. J. 565
(2002), 405
3) K. Maeda, P.A. Mazzali, J. Deng, K. Nomoto, Y. Yoshii, H. Tomita, Y. Kobayashi, As-
trophys. J., 593, in press
4) K. Maeda and K. Nomoto, Nucl. Phys. A (2003), 718C, 167
5) K. Maeda and K. Nomoto, in 'From Twilight to Highlight: The Physicas of Supernovae,'
eds. W. Hillebrandt and B. Leibundgut (Berlin; Springer, 2003), 373
6) P. Mazzali, K. Iwamoto, K. Nomoto, Astrophys. J. 545 (2000), 407
7) P. Mazzali, K. Nomoto, F. Patat, K. Maeda, Astrophys. J. 559 (2001), 1047
8) T. Nakamura, P. Mazzali, K. Nomoto, K. Iwamoto, Astrophys. J. 550 (2001), 991
9) K. Nomoto, K. Maeda, H. Umeda, T. Ohkubo, J. Deng, P. Mazzali, in 'A massive Star
Odyssey, from Main Sequence to Supernova,' eds. V.D. Hucht, A. Herrero and C. Esteban,
(San Francisco; ASP, 2003), 395, astro-ph/0209064
10) K. Nomoto, K. Maeda, H. Umeda, T. Ohkubo, J. Deng, P. Mazzali, Prog. Theor. Phys.
Suppl. (2003), in press
11) Ph. Podsiadlowski, K. Nomoto, K. Maeda, T. Nakamura, P. Mazzali, B. Schmidt, Astro-
phys. J. 567 (2002), 491
|
astro-ph/0210329 | 1 | 0210 | 2002-10-15T12:55:02 | Large-Scale Structure in the NIR-Selected MUNICS Survey | [
"astro-ph"
] | The Munich Near-IR Cluster Survey (MUNICS) is a wide-area, medium-deep, photometric survey selected in the K' band. The project's main scientific aims are the identification of galaxy clusters up to redshifts of unity and the selection of a large sample of field early-type galaxies up to z < 1.5 for evolutionary studies. We created a Large Scale Structure catalog, using a new structure finding technique specialized for photometric datasets, that we developed on the basis of a friends-of-friends algorithm. We tested the plausibility of the resulting galaxy group and cluster catalog with the help of Color-Magnitude Diagrams (CMD), as well as a likelihood- and Voronoi-approach. | astro-ph | astro-ph |
Large-Scale Structure in the NIR-Selected MUNICS
Survey
C. S. Botzler1, J. Snigula1, R. Bender1, N. Drory1, G. Feulner1,
G. J. Hill2, U. Hopp1, C. Maraston1, C. Mendes de Oliveira3
1 University Observatory Munich (USM), Germany
2 University of Texas at Austin, USA
3 Instituto Astronomico e Geofisico, Sao Paulo, Brazil
Abstract. The Munich Near-IR Cluster Survey (MUNICS) is a wide-area, medium-
deep, photometric survey selected in the K' band. The project's main scientific aims
are the identification of galaxy clusters up to redshifts of unity and the selection of
a large sample of field early-type galaxies up to z < 1.5 for evolutionary studies.
We created a Large Scale Structure catalog, using a new structure finding technique
specialized for photometric datasets, that we developed on the basis of a friends-
of-friends algorithm. We tested the plausibility of the resulting galaxy group and
cluster catalog with the help of Color-Magnitude Diagrams (CMD), as well as a
likelihood- and Voronoi-approach.
Keywords: galaxies: photometric redshift, survey, cluster finding: friends-of-friends
1. Motivation
The MUNICS survey was created in order to identify clusters of galaxies
at high redshifts by detecting their luminous early-type galaxy popula-
tion and to use the resulting mass function of clusters for cosmological
tests (Bahcall et al., 1997). Another aim of the survey was to utilize
field- and cluster-galaxies over a wide range of redshifts as a laboratory
for galaxy evolution (Drory et al., 2001a; Feulner et al., 2002).
2. MUNICS in Brief
Our survey covers an area of roughly 1 deg2 in the near-IR J and K'
bands, supplemented with 0.6 deg2 of V, R, and I band imaging. The
galaxy catalog was selected in K' with a 50% completeness at K' ∼
19.5m (Drory et al., 2001b). Redshifts were determined from V, R,
I, J and K' band photometry using a photometric redshift technique
(Bender et al., 2001), that makes use of model SEDs (Maraston, 1998)
and empirical templates, and was calibrated with ∼ 500 spectroscopic
redshifts (Feulner et al., 2002) (see Fig. 1, left panel). The catalog
contains ∼ 5000 galaxies with z ≥ 0.4.
c(cid:13) 2019 Kluwer Academic Publishers. Printed in the Netherlands.
Botzler.tex; 3/01/2019; 17:56; p.1
<∆z> = -0.006
σ(∆z) = 0.075
2
100
80
60
40
20
y
c
n
e
u
q
e
r
f
<z> = 0.77
3.5
3.0
2.5
2.0
1.5
1.0
0.5
'
K
-
J
0
-1.0
-0.5
0.0
0.5
1.0
0.0
15
16
17
18
19
20
∆z/(1+z)
K'
Figure 1. left panel: Histogram of the redshift errors. The rms scatter is consistent
with a Gaussian (dotted line: best-fit Gaussian) of a width σ = 0.075 and an
insignificant mean deviation from the unity relation of h∆zi = −0.006. right panel:
J-K' vs. K' color-magnitude diagram of one group at hzi = 0.77. The filled boxes
denote the galaxies that were found to be structure members by our FOF. The open
squares denote all galaxies found within 5 core-radii (rc = 0.125 h−1
100 Mpc)(at hzi)
of the cluster center. The group members roughly trace a red-sequence.
3. The Extended Friends-of-Friends Algorithm
So far, friends-of-friends (FOF) algorithms (Huchra and Geller, 1982)
were designed to find number density enhancements in spectroscopic
redshift surveys or numerical simulations (Nolthenius and White, 1987).
Due to the comparatively large uncertainties in the photometric red-
shifts, the original FOF-technique had to be modified for application
to MUNICS, resulting in our "extended friends-of-friends" algorithm.
This modified algorithm takes the photometric redshift errors into
account. It looks for friends compatible with an a priori redshift on a
redshift grid, spanning the entire depth of the survey and finally unifies
the structures found in individual redshift bins, if structures have at
least one member in common.
We found 104 structures within 0.4 ≤ z ≤ 1.1 in our survey field,
which corresponds to a (flux-limited) number density of ∼ 8 · 10−5
structures Mpc−3 within 0.5 ≤ z ≤ 0.8 for an ΩM = 0.3, ΩΛ = 0.7, and
h100 = 0.7 cosmology.
Botzler.tex; 3/01/2019; 17:56; p.2
Large Scale Structure in MUNICS
3
1500
1000
y
500
]
g
e
d
[
c
e
D
65.65
65.60
65.55
65.50
500
1000
1500
0.6
0.5
0.4
0.3
0.2
x
RA [deg + 179]
Figure 2. left panel: Voronoi tessellations and extended-FOF group members for
all galaxies with 0.46 ≤ z ≤ 0.74 of a MUNICS field. Galaxies with Voronoi areas
smaller than the mean area of a random field Voronoi tessellation (i.e. overdense)
are shown as large symbols. Galaxies in underdense regions are denoted by small
symbols. Squares denote non-FOF objects. Members of FOF structure are plotted as
triangles. The graphic shows that they lie preferably in overdense regions, suggesting
consistency with the Voronoi approach. right panel: Likelihood distribution and
extended-FOF group members for all galaxies within 0.45 ≤ z ≤ 0.75 of the same
MUNICS field as seen in Fig. 2, right panel. The likelihood is shown in grey-scales,
with white symbolizing the highest probability. Members of our FOF structures
are plotted as open black triangles. The graphic shows the tendency for the FOF
structures to lie in areas of increased likelihood.
4. Testing the Structures
4.1. Color-Magnitude Diagrams
In a J-K' vs. K' CMD, evolved members of a bound structure are
expected to be lying roughly on a horizontal line, the "red sequence"
(see Fig. 1, right panel). About 50% of our CMDs look reasonable for
groups or clusters.
4.2. Voronoi Tessellation
The inverse of the area of a Voronoi cell around a galaxy is a measure
for the local density of the galaxy distribution and can be used for
finding accumulations of galaxies and finally clusters (Ramella et al.,
2001).
We prepared Voronoi tessellations in preselected photometric red-
shift bins for our MUNICS dataset and compared the resulting density
charts with the results of our FOF algorithm (see Fig. 2, left panel).
Botzler.tex; 3/01/2019; 17:56; p.3
4
4.3. Likelihood Approach
We furthermore determined the probability that a given point on a
uniform grid in redshift-space is the center of a galaxy cluster, taking
into account the known distributions of galaxies in the MUNICS survey.
To estimate this likelihood, we assumed Gaussian probability distribu-
tions for a typical cluster in redshift, projected position and J-K' color.
We included empirical knowledge about typical cluster core radii (rc =
0.125 h−1
100 Mpc; Bahcall, 1977) and velocity dispersions (σV = 103 km
s−1), as well as the individual errors of the MUNICS galaxies and SSP
model colors (Maraston, 1998) to define the rms of the distributions
(see Fig. 2, right panel).
Acknowledgements
The MUNICS project was suported by the Deutsche Forschungsge-
meinschaft, Sonderforschungsbereich 375.
References
Bahcall, N. A.: 1977, 'Clusters of galaxies'. ARA&A 15, 505 -- 540.
Bahcall, N. A., X. Fan, and R. Cen: 1997, '"Constraining Omega with Cluster
Evolution"'. ApJ 485, L53+.
Bender, R., I. Appenzeller, A. Bohm, N. Drory, K. J. Fricke, A. Gabasch, J. Heidt, U.
Hopp, K. Jager, M. Kummel, D. Mehlert, C. Mollenhoff, A. Moorwood, H. Nick-
las, S. Noll, R. P. Saglia, W. Seifert, S. Seitz, O. Stahl, E. Sutoriaus, T. Szeifert,
S. J. Wagner, and B. L. Ziegler: 2001, '"The FORS Deep Field: Photometric
Data and Photometric Redshifts"'.
In: S. Christiani (ed.): ESO/ECF/STScI
Workshop on Deep Fields. Berlin, pp. 327+, Springer.
Drory, N., R. Bender, J. Snigula, G. Feulner, U. Hopp, C. Maraston, G. J. Hill, and
C. M. de Oliveira: 2001a, 'The Munich Near-Infrared Cluster Survey: Number
Density Evolution of Massive Field Galaxies to z1.2 as Derived from the K-
Band-selected Survey'. ApJ 562, L111 -- LL114.
Drory, N., G. Feulner, R. Bender, C. S. Botzler, U. Hopp, C. Maraston, C. Mendes
de Oliveira, and J. Snigula: 2001b, '"The Munich Near-Infrared Cluster Survey
-- I. Field selection, object extraction, and photometry"'. MNRAS 325, 550 -- 562.
in
Feulner, G., N. Drory, U. Hopp, R. Bender, J. Snigula, and G. J. Hill: 2002.
preparation.
Huchra, J. P. and M. J. Geller: 1982, 'Groups of galaxies. I - Nearby groups'. ApJ
257, 423 -- 437.
Maraston, C.: 1998, 'Evolutionary synthesis of stellar populations: a modular tool'.
MNRAS 300, 872 -- 892.
Nolthenius, R. and S. D. M. White: 1987, 'Groups of galaxies in the CfA survey and
in cold dark matter universes'. MNRAS 225, 505 -- 530.
Ramella, M., W. Boschin, D. Fadda, and M. Nonino: 2001, 'Finding galaxy clusters
using Voronoi tessellations'. A&A 368, 776 -- 786.
Botzler.tex; 3/01/2019; 17:56; p.4
|
0801.2159 | 1 | 0801 | 2008-01-14T21:05:27 | A High-Resolution Spectrum of the Highly Magnified Bulge G-Dwarf MOA-2006-BLG-099S | [
"astro-ph"
] | We analyze a high-resolution spectrum of a microlensed G-dwarf in the Galactic bulge, acquired when the star was magnified by a factor of 110. We measure a spectroscopic temperature, derived from the wings of the Balmer lines, that is the same as the photometric temperature, derived using the color determined by standard microlensing techniques. We measure [Fe/H]=0.36 +/-0.18, which places this star at the upper end of the Bulge giant metallicity distribution. In particular, this star is more metal-rich than any bulge M giant with high-resolution abundances. We find that the abundance ratios of alpha and iron-peak elements are similar to those of Bulge giants with the same metallicity. For the first time, we measure the abundances of K and Zn for a star in the Bulge. The [K/Mg] ratio is similar to the value measured in the halo and the disk, suggesting that K production closely tracks alpha production. The [Cu/Fe] and [Zn/Fe] ratios support the theory that those elements are produced in Type II SNe, rather than Type Ia SNe. We also measured the first C and N abundances in the Bulge that have not been affected by first dredge-up. The [C/Fe] and [N/Fe] ratios are close to solar, in agreement with the hypothesis that giants experience only canonical mixing. | astro-ph | astro-ph | A High-Resolution Spectrum of the Highly Magnified Bulge
G-Dwarf MOA-2006-BLG-099S1
Jennifer A. Johnson2, B. Scott Gaudi2, Takahiro Sumi3, Ian A. Bond4 and Andrew Gould2
8
0
0
2
n
a
J
4
1
]
h
p
-
o
r
t
s
a
[
1
v
9
5
1
2
.
1
0
8
0
:
v
i
X
r
a
ABSTRACT
We analyze a high-resolution spectrum of a microlensed G-dwarf in the Galac-
tic bulge, acquired when the star was magnified by a factor of 110. We measure
a spectroscopic temperature, derived from the wings of the Balmer lines, that is
the same as the photometric temperature, derived using the color determined by
standard microlensing techniques. We measure [Fe/H]=0.36 ± 0.18, which places
this star at the upper end of the Bulge giant metallicity distribution. In partic-
ular, this star is more metal-rich than any bulge M giant with high-resolution
abundances. We find that the abundance ratios of alpha and iron-peak elements
are similar to those of Bulge giants with the same metallicity. For the first time,
we measure the abundances of K and Zn for a star in the Bulge. The [K/Mg]
ratio is similar to the value measured in the halo and the disk, suggesting that K
production closely tracks α production. The [Cu/Fe] and [Zn/Fe] ratios support
the theory that those elements are produced in Type II SNe, rather than Type Ia
SNe. We also measured the first C and N abundances in the Bulge that have not
been affected by first dredge-up. The [C/Fe] and [N/Fe] ratios are close to solar,
in agreement with the hypothesis that giants experience only canonical mixing.
Subject headings: gravitational lensing -- stars: abundances -- Galaxy: abun-
dances -- Galaxy: bulge -- Galaxy: evolution
1This paper includes data gathered with the 6.5 meter Magellan Telescopes located at Las Campanas
Observatory, Chile.
2Department of Astronomy, Ohio State University, 140 W. 18th Ave., Columbus, OH 43210, USA;
jaj,gaudi,[email protected]
3Solar-Terrestrial Environment Laboratory Nagoya University, Nagoya, Japan;[email protected]
u.ac.jp
4Institute
for
Information and Mathematical Sciences, Massey University, Auckland, New
Zealand;[email protected]
-- 2 --
1.
Introduction
The Galactic Bulge underwent an intense burst of star formation early in the formation
of the Galaxy, leading to a very different stellar population and chemical evolution history
than found in the Milky Way disk or halo (e.g., Ortolani et al. 1995; Zoccali et al. 2003;
McWilliam & Rich 1994). In particular, massive stars may dominate the pollution at almost
all metallicities, leading to unique abundance patterns (e.g., Lecureur et al. 2007). Similar
events are thought to mark the formation of other galactic spheroids, making the Bulge
stellar population a template for interpreting extragalactic observations. As a result of its
unique formation history in the Galaxy, the Bulge has been the subject of intensive study.
The detection of RR Lyrae stars (Baade 1946) first indicated that the Bulge contained
old stars. With deeper photometry, the main sequence turnoff (MSTO) of the Bulge was
detected. Terndrup (1988) found a mean age of 11-14 Gyr for stars in Baade's Window, with
a negligible fraction of stars with ages < 5 Gyr. Photometry reaching more than 2 magnitudes
below the MSTO with Hubble Space Telescope confirmed the generally old nature of the Bulge
(Ortolani et al. 1995; Holtzman et al. 1998), although Feltzing & Gilmore (2000) included a
reminder that a young metal-rich population has similar MSTO colors and luminosities to
an older, more metal-poor population. Therefore, deriving ages reliably from photometry
of the MSTO requires adequate knowledge of the Bulge metallicity distribution function
(MDF), specifically for the MSTO stars. However, because of the faintness of those stars,
the measurement of the Bulge MDF has historically relied on giants.
Since the discovery of both M giants and RR Lyr stars, it has been known that the
Bulge giants span a wide range in metallicity. Low-dispersion spectra provided the first
quantitative measure of the MDF (Whitford & Rich 1983; Rich 1988). Sadler et al. (1996)
measured indices from low-dispersion spectra of 268 K giants (both red clump stars and first
ascent giants) to derive a mean metallicity1) h[Fe/H]i = −0.11, with a dispersion of 0.46
dex. Recalibration by Fulbright et al. (2006) based on high-resolution spectra of 15 stars
in common with the Sadler et al. (1996) sample reduced the mean metallicity to −0.22.
Ram´ırez et al. (2000) measured h[Fe/H]i = −0.21 from low-dispersion near-infrared spectra
for 72 M giants in the inner Bulge. The good agreement between Ram´ırez et al. (2000) result
and the recalibrated Sadler et al. (1996) result is somewhat surprising. At the bright end
of the giant branch, only metal-rich first ascent giants become M giants. However, both K
giants and M giants become red clump stars, and lower luminosity metal-rich giants are K
stars as well. Therefore, the inclusion of red clump stars and fainter giants in the Sadler et al.
(1996) sample make the biases in their sample more similar to those of the Ram´ırez et al.
1We adopt the usual spectroscopic notation that [A/B] ≡ log10(NA/NB)⋆ -- log10(NA/NB)⊙
-- 3 --
(2000) study.
Zoccali et al. (2003) measured both the MDF of the bulge and the age of the stars
using deep photometry of the Bulge in the optical and near-infrared wavelengths. They
constructed the MK and (V-K)0 color magnitude diagram for a low-reddening window at
(l, b)=(0.277, −6.167). Because the slope of the red giant branch (RGB) in these colors
depends on the metallicity, RGB stars with different metallicities fall on different parts of
the color-magnitude diagram. They could therefore use their photometry to derive the MDF
of 503 giants by comparing the positions of bright (MK < −4.5) RGB stars with globular
cluster fiducials of known metallicity. After correcting for small biases in their MDF caused
by their magnitude cutoff, they find an MDF with a peak at [M/H]= −0.1, a sharp cutoff
at [M/H]= −0.2 and few stars with [M/H]< −1. Adopting the metallicities derived from
the giants for the MSTO dwarfs, they estimated that the Bulge is coeval with the halo and
argued that the lack of stars above the prominent MSTO of the Bulge ruled out a significantly
younger population.
These measurements of the Bulge giant MDF can be improved by metallicity measure-
ments from high-dispersion measurements of many stars in several fields throughout the
Bulge. Recently, Lecureur et al. (2008) and Zoccali et al. (2008) have obtained a total of
∼ 1000 K giant spectra at (R∼20,000) in 4 Bulge windows. They confirm the metal-rich
nature of the Bulge.
With such a metal-rich population having been reached so quickly after star formation
began, the Bulge is expected to have a distinct chemical evolution compared to other Galactic
populations, such as the halo or the disk because, for example, the contributions of longer-
lived polluters, such as Type Ia supernovae (SNe) or low-mass AGB stars, should be small.
Indeed, McWilliam & Rich (1994) measured high [α/Fe] ratios in giants, in particular, high
[Mg/Fe] for [Fe/H] values up to solar, arguing for little Type Ia SN contribution of Fe com-
pared to the thick or thin disks. Fulbright et al. (2007) confirmed the overall enhancement
in [Mg/Fe] and strengthened the conclusion of McWilliam & Rich (1994) that the other α
abundance ratios do not track [Mg/Fe] exactly.
[O/Fe], [Si/Fe], [Ca/Fe] and [Ti/Fe] begin
to decrease around [Fe/H]=0, while [Mg/Fe] does so at supersolar [Fe/H]. Fulbright et al.
(2007) suggested that metallicity-dependent Type II SN yields could explain the different
behaviors of the α elements. McWilliam et al. (2007) explained the low [O/Mg] ratios in
metal-rich Bulge giants through a different metallicity-dependent mechanism: Wolf-Rayet
winds leading to less effective O production in metal-rich massive stars. By [Fe/H]∼0.2,
all the [α/Fe] ratios have begun to decline, probably indicating the introduction of large
amounts of Fe from Type Ia SNe (e.g., Cunha & Smith 2006).
Several studies have looked at the abundances of the light elements Na and Al in the
-- 4 --
Bulge, two elements whose production should depend on the metallicity of the massive stars
that exploded as Type II SNe. Cunha & Smith (2006) measured Na in 7 K and M gi-
ants and found the predicted increase in [Na/Fe] and [Na/O] in the most metal-rich stars.
Lecureur et al. (2007) found supersolar [Al/Fe] at all metallicities and, for [Fe/H] >0, en-
hanced [Na/Fe] compared to the ratios in disk stars. Interestingly, at higher metallicities,
the scatter in [Al/Fe] and [Na/Fe] increased and became larger than could be explained by
observational errors. Interpreting the Na and Al abundances as the result of Type II SN
production may be problematic. The surface abundances of Al and Na have been shown
to be increased by large amounts of internal mixing in metal-poor globular cluster stars
(e.g. Shetrone 1996), where the products of proton-capture reactions deep inside the star
are mixed up to the surface, leading to enhancements in these two elements. However,
Lecureur et al. (2007) argued that the C and N abundances in the giants they studied were
consistent with only mild mixing and, therefore, that the high Na and Al had to be due to
the overall chemical evolution of the Bulge. Cunha & Smith (2006) also found evidence for
mild mixing in giants, affecting C and N, but not O, Na, or Al.
Finally, there is little information on the neutron-capture elements in the Bulge. The
absorption lines for these elements are concentrated in the blue part of the optical spectrum,
where the crowding from Fe, CN, and other lines is severe. Near-IR spectra have essentially
no lines of these elements. While there are a few lines of Ba in the red, these lines in
metal-rich giants are so saturated that reliable measurements are very difficult. As a result,
only the neutron-capture element Eu has published results so far. McWilliam & Rich (1994)
found [Eu/Fe]>0 in Bulge giants, which is likely because of the production of Eu in the
r-process. Measuring additional neutron-capture element abundances would test this idea,
because the r-process is better at making some elements (e.g., Eu) than others (e.g., Ba).
Our knowledge of the metallicity and abundance ratios of Bulge stars has generally been
confined to the bright giants, which are usually the only ones accessible to high-resolution
observations. But studying the dwarfs has several advantages. Their abundances of the
elements such as C and N are unaffected by dredge-up processes on the giant branch. We
can measure elements, such as S and Zn, that not measured in giants, because the hotter
temperatures of the dwarfs decrease the strength of CN and increase the strength of certain
atomic lines.
In addition, it is critical to know the metallicity of stars at the MSTO to
accurately measure the ages from their color and luminosity. Finally, individual ages can be
assigned to dwarf stars near the turnoff.
The advent of large surveys to identify and follow-up microlensing events, such as the
-- 5 --
Microlensing Observations in Astrophsycis (MOA) collaboration2, the Optical Gravitational
Lens Experiment3 (OGLE), the Microlensing Follow Up Network4 (µFUN) and the Prob-
ing Lensing Anomalies Network5 (PLANET), provides an opportunity to study otherwise
unobservable Bulge dwarfs. During high-magnification microlensing events, it is possible to
obtain high-resolution, high signal-to-noise ratio spectra of faint stars with a huge savings
in observing time: a factor A × 100.4∆m where A is the magnification and ∆m is the number
of magnitudes below sky of the unmagnified star. In Johnson et al. (2007) , we reported the
detailed abundances for a highly-magnified Bulge dwarf, OGLE-2006-BLG-265S, which from
a 15 minute exposure at magnification A ∼ 135, was shown to be one of the most metal-rich
stars ever observed. It also provided the first measurements of S and Cu in the Bulge. Here
we present a high-resolution spectrum of the Bulge G-dwarf MOA-2006-BLG-99S, taken at
magnification A = 110.
Finally, we respond the challenge: "Ask not what microlensing can do for stellar spec-
troscopy -- ask what stellar spectroscopy can do for microlensing." There is one important
way that the spectroscopic study of bulge dwarfs can benefit microlensing. Whenever a source
approaches or transits a "caustic" (line of infinite magnification) caused by the lens, one can
measure ρ, the ratio of the angular source radius to angular Einstein radius ρ = θ∗/θE, from
the microlens lightcurve. Then θ∗ is inferred from the dereddened color and magnitude of
the source to yield θE = θ∗/ρ, which in turn provides important constraints on the lens
properties. Because spectroscopy is not normally available for these microlensed sources,
the dereddened color and magnitude are estimated by comparing the source position on
an instrumental color-magnitude diagram with that of the clump and then assuming that
the Bulge clump is similar to the local clump as measured by Hipparcos (Yoo et al. 2004a).
This procedure undoubtedly suffers some statistical errors and could suffer systematic er-
rors as well. For example, the Bulge clump may have a different color from the local one.
High-resolution spectra of an ensemble of microlensed bulge sources will test this proce-
dure for both statistical and systematic errors. For OGLE-2006-BLG-265S, the standard
microlensing procedure yielded (V − I)0 = 0.63 ± 0.05, whereas Johnson et al. (2007) ob-
tained (V − I)0 = 0.705 ± 0.04 from high-resolution spectroscopy. This difference hints at a
possible discrepancy, but only by repeating this procedure on a number of dwarfs can this
be confirmed.
2http://www.massey.ac.nz/∼iabond/alert/alert.html
3http://www.astrouw.edu.pl/∼ ogle/ogle3/ews/ews.html
4http://www.astronomy.ohio-state.edu/∼microfun/
5http://planet.iap.fr/
-- 6 --
2. Observations
MOA-2006-BLG-99 was alerted as a probable microlensing event toward the Galactic
Bulge (J2000 RA = 17:54:10.99, Dec = −35:13:38.0; l = −4.48, b = −4.78) by MOA on 22
July 2006. On 23 July, MOA issued a further alert that this would be a high-magnification
event, with A > 100. Intensive photometric observations were then carried by several collab-
orations, including µFUN, primarily with the aim of searching for planets (Mao & Paczy´nski
1991; Griest & Safizadeh 1998; Udalski et al. 2005; Gould et al. 2006). Results of that search
will be presented elsewhere. The event actually peaked on 23 July (HJD 2453940.349) at
Amax ∼ 350. One of us (BSG) happened to be at the Clay 6.5-m Telescope at the Magellan
Observatory when he received the flurry of µFUN emails describing this event. He then
interrupted his normal program to obtain a 20-minute exposure of this event at the begin-
ning of the night, just after peak, when the magnification was A = 215. Unfortunately, the
atmospheric transparency was poor, and the observation had to be interrupted after 1089
seconds when the cloud cover became too thick. Conditions cleared several hours later, and
he obtained two 20-minute exposures centered at UT 02:09:36 and UT 02:30:34 24 July,
when the magnification was A = 113 and A = 107. We base our results on these higher
quality spectra.
The observations of MOA-2006-BLG-99 were made using the Magellan Inamori Kyocera
Echelle (MIKE) double spectrograph (Bernstein et al. 2003) mounted on the Clay telescope
on Las Campanas. with seeing of 0.7 -- 1.0 arcsec. We used the 1.0 arcsec slit, which produces
R∼25,000 on the blue side and R∼19,000 on the red side.
3. Data Reduction
The data were reduced using the MIKE Python data reduction pipeline (D. Kelson,
2007, private communication), with the exception of the bluest orders containing the CH
and CN lines, which were reduced using the IRAF6 echelle package. The bias and overscan
were subtracted. The wavelength calibration was derived from Th-Ar data. Flatfields were
taken through a diffusor slide to create "milky flats" that made the orders sufficiently wide
to get good flatfields along the edges of the orders of the data frames. Parts of orders
with overlapping wavelength coverage were coadded together before analysis. These overlap
regions are larger for the bluer parts of the spectrum. Over the wavelength region where
6IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the As-
sociation of Universities for Research in Astronomy, Inc., under cooperative agreement with the National
Science Foundation.
-- 7 --
individual lines were measured (5300-8000A), the signal-to-noise (S/N) is ∼ 30. The CN
and CH bandhead regions had lower S/N (S/N∼10).
4. Abundance Analysis
We analyzed both the spectrum of MOA-2006-BLG-99S and the spectrum of the Sun
(Kurucz et al. 1984) using the same set of lines, line parameters, and model atmosphere
grids. We used TurboSpectrum (Alvarez & Plez 1998), a 1-dimensional LTE code, to derive
abundances. TurboSpectrum uses the recent treatment of damping from Barklem et al.
(2000). We interpolated the model atmosphere grid of ATLAS9 models7 updated with new
opacity distribution functions (Castelli & Kurucz 2003). The abundances of most elements
were determined from analysis of the equivalent widths (EWs). The EWs for both MOA-
2006-BLG-99S and the Sun are presented in Table 1. We restricted the analysis to lines
with EW≤ 150mA. The EWs were measured using SPECTRE8 (C. Sneden, 2007, private
communication). We compared synthetic with observed spectra to determine abundances for
lines that were blended, lines that had substantial hyperfine splitting and for C and N that
were measured from CH and CN lines, respectively. We used the solar atlas of Moore et al.
(1966) to check for blending with telluric features and eliminated the few lines that were
affected.
The linelists for CH and CN are from B. Plez (2006, private communication). The effect
of hyperfine splitting (HFS) was included for Sc, Mn, Cu and Ba. The HFS constants were
taken from the sources listed in Johnson et al. (2006). HFS information was not available
for the Na or Al lines, so neither this study nor the literature studies we use for comparison
can correct for those effects. Therefore, the comparison between [Na/Fe] and [Al/Fe] val-
ues for MOA-2006-BLG-99S and the literature values is robust, but the absolute values of
these abundance ratios for all studies have a systematic error. Table 1 lists the transition
probabilities (listed as log gf -values) and sources for all the atomic lines.
4.1. Atmospheric Parameters
We measured a Teff=5800K using the wings of the H-α and H-β lines (Figure 4.1).
Our uncertainty in the temperature is based on the uncertainty in this fit, in particular
7http://kurucz.harvard.edu/grids.html
8http://verdi.as.utexas.edu/spectre.html
-- 8 --
the uncertainty in the level of the continuum. Temperatures derived from the Balmer lines
for metal-rich dwarfs show good agreement with temperatures derived from the infrared
flux method and from V − K colors (Mashonkina et al. 1999; Barklem et al. 2002). For
MOA-2006-BLG-99S, we compared the temperature derived from the Balmer lines with
that derived from excitation equilibrium of the Fe I lines. The temperature derived from
the Fe I lines was very uncertain because the abundances derived from individual lines were
poorly determined because of the S/N of the spectrum. The excitation equilibrium favored
a higher temperature (+200K) with an uncertainty of 300K.Given the uncertainties, this
temperature is in agreement with the Balmer line temperature. Next, we measured the
microturbulent velocity, ξ, by ensuring that the abundances derived from the Fe I lines
do not depend on their reduced EW. Our best fit value was ξ=1.5 km/s. Changing ξ by
±0.3 km/s gave marginal fits to the data, and we adopt that as our uncertainty in ξ. The
gravity was measured by ionization balance for Fe I and Fe II and for Ti I and Ti II. As
a sanity check, in Figure 2, we compare the atmosphere parameters for MOA-2006-BLG-
99S, OGLE-2006-BLG-265S, and the Sun to the parameters from the Yonsei-Yale (Yi et al.
2001; Demarque et al. 2004) isochrones. Finally, the metallicity of the model atmosphere,
[m/H], was set equal to the [Fe/H] given by the Fe I lines. Because the log g of the model
atmosphere affects the [m/H] and the [m/H] affects the abundances (and therefore the log
g measurement), we iterated to find a solution for which Fe I and Fe II were equal and the
[m/H] of the model was equal to [Fe I/H].
Using standard microlensing techniques (e.g., Yoo et al. 2004a), µFUN determined
that the dereddened color and magnitude of the source were (V − I)0 = 0.69 ± 0.05,
I0 = 18.17 ± 0.10. The error is due to possible differential reddening between the mi-
crolensed source and the red clump, which is assumed to have the same (V − I)0 = +1.00
as the local Hipparcos clump.
If (as seems likely) the source lies at approximately the
Galactocentric distance, then its absolute magnitude is MV = 4.5 and its radius is 1.2 R⊙.
That is, it is a solar-type star. We combined the µFUN color and the Ram´ırez & Mel´endez
(2005) color-temperature relation (an update of the Alonso et al. (1996) relation) to derive
an estimate of the temperature of 5806± 200 K. This photometric temperature agrees with
the temperature from the Balmer lines. For OGLE-2006-BLG-265S, the spectroscopic and
photometric temperatures were different, suggesting possible systematic errors in estimating
the dereddened color and magnitude of the source star. As outlined in the Introduction,
quantifying any systematic errors is important for obtaining the most accurate information
about microlensing events, and additional data will be crucial for quantifying any systematic
error.
-- 9 --
Fig. 1. -- Fits to the H-α line for three different temperatures: 5600K, 5800K and 6000K.
The atmosphere with Teff=5800K is the best fit to the hydrogen line.
-- 10 --
Fig. 2. -- The position of MOA-2006-BLG-99S (square), OGLE-2006-BLG-265S (circle) and
the Sun (triangle) in the H-R diagram. We also show isochrones from Yi et al. (2001). The
solid lines show isochrones for [Fe/H]=0.385 and for ages 2, 3, 4, 5 and 7 Gyr. The dashed
line shows a 5 Gyr isochrone for a solar metallicity.
-- 11 --
4.2. Error Analysis
Our uncertainties are 200K for Teff and 0.3 km/s for ξ. The distribution of EWs and
excitation potential for the Fe I lines were sufficiently uncorrelated that ξ did not depend on
Teff. Our measurements of Teff and ξ therefore did not depend on accurate assessments of the
other model atmosphere parameters, namely log g and [m/H]. Our measurement of log g and
[m/H], on the other hand, depended strongly on the other model atmosphere parameters.
Because we set the gravity by ionization equilibrium, log g depends on Teff , ξ, and [m/H],
as well as the abundance measured from the Fe I and Fe II lines. Therefore, the uncertainty
in log g depends on the uncertainties in those quantities. The uncertainty in [m/H], in turn,
depends on the uncertainty in Teff , log g, and ξ, as well as the scatter in the abundance
given by different Fe I lines for a particular model atmosphere. The standard error of the
mean for the gravity derived from the 35 Fe I lines for a single model atmosphere was 0.04
dex, and for the 4 Fe II lines was 0.14. We adopt 0.04 dex as the uncertainty from EW
and log gf errors for Fe I for inclusion in the [m/H] uncertainty. Adding the Fe I and Fe II
uncertainties in quadrature give us 0.15 dex as our uncertainty in the difference between the
Fe I and the Fe II logǫ9 values arising from the EW and oscillator strength uncertainties.
The total uncertainty in log g is 0.66 dex and in [m/H] is 0.21 dex.
We ran the Fe I and Fe II EWs through a series of models: ± 200K, ± 0.3 km/s, ±0.3
dex for log g, and 0.13 dex in [m/H] (smaller because of the limits of the Kurucz grid), and
calculated the difference in Fe abundance with these different model atmospheres.
Finally, we calculate uncertainties using modified equations (A5) and (A20) from McWilliam et al.
log g, and ξ :
log g, Teff:
[m/H] , ξ
(1995). We considered the covariance between Teff :
:
[m/H]. The covariances were calculated by a Monte Carlo technique. For
example, to calculate the covariance between Teff and log g, we first found ∂ log g/∂Teff and
noted the remaining scatter that was caused by uncertainty in ξ, etc. Next, we randomly
picked 1000 Teff from a Gaussian distribution with a σ of 200K. We used the derivative to
calculate log g and then added an extra random ∆log gdrawn from a Gaussian distribution
with a σ equal to the uncertainty from non-Teff causes. We calculated the covariance using
these 1000 Teff -log g pairs. A similar calculation was done for the other covariances.
[m/H] , log g:
9log ǫ(A) ≡ log10(NA/NH) + 12.0
-- 12 --
5. Results
In Table 2, we summarize the abundances measured for 17 elements in MOA-2006-BLG-
99S. We include both logǫ and its error, as well as [X/Fe] and its error. To give an idea of the
uncertainty due to scatter from the lines, rather than from atmospheric parameters, we give
σ, the rms of abundances derived from individual lines as well as the number of lines. We
also give our measurements of the solar abundances, which we will use to calculate ratios.
For reference, we include the Grevesse & Sauval (1998) solar abundances in the final column.
5.1. Metallicity
We measure [Fe/H]= 0.36 ± 0.18 for MOA-2006-BLG-99S. In Johnson et al. (2007), we
measured [Fe/H]= 0.56 ± 0.19 for the dwarf OGLE-2006-BLG-265S. The stars that are mi-
crolensed are unbiased in metallicity. The criterion for spectroscopic follow-up is that the
unmagnified source be faint enough to be a Bulge dwarf, regardless of color. Therefore,
especially considering the large and variable reddening toward the Bulge, we are not biased
in our high-resolution follow-up toward high metallicity sources. The high metallicities of
these two dwarfs is surprising given that work on giants has indicated an average metallicity
near solar.
In Figure 3, we compare the metallicity distribution function (MDF) of the
two dwarfs with several MDFs based on studies of giant stars. The MDF of Rich et al.
(2007), which is based on high-resolution analysis of M giants and should be biased toward
the highest metallicity objects, lacks any giants as metal-rich as these dwarfs. This is partic-
ularly notable, since the work of Sadler et al. (1996) on K giants with low-dispersion spectra
shows some extremely metal-rich stars. A complicating factor is the possible presence of a
metallicity gradient in the Bulge. In Figure 4, we compare the metallicities of the dwarfs
with MDFs derived by Zoccali et al. (2008) from high-resolution spectra for giants in three
Bulge fields: 4◦, 6◦, and 12◦away from the Galactic center. The inner field is more metal-
rich than the outer. However, these two dwarfs are located 6.5◦(MOA-2006-BLG-99S) and
4.9◦(OGLE-2006-BLG-265S) away from the Galactic Center, and therefore gradients cannot
explain their anomalously high metallicities.
These results hint that the MDFs of the Bulge giants and dwarfs may be different.
Whether this is true can be established by more observations of Bulge dwarfs and by res-
olution of the discrepancies among the MDFs derived for giants, particularily between the
low-dispersion and high-dispersion studies. Because the microlensed dwarfs are found at a
range of distances from the Galactic Center, comparison of the giant and dwarf MDFs also
depends on measuring the metallicity gradient (and the size of deviations from that gradient)
in the Bulge. Ideally, the MDF for giants in the same field as the microlensed dwarf would
-- 13 --
be measured.
5.2. Comparison with Isochrones and the Age of MOA-2006-BLG-99S
With our measurements of Teff, log g, and [Fe/H] from spectroscopy, we can compare
the position of MOA-2006-BLG-99S on the Hertzsprung-Russell diagram with theoretical
isochrones (Fig. 2). We use the Yonsei-Yale isochrones (Yi et al. 2001; Demarque et al.
2004) with [Fe/H]=0.385, [α/Fe]=0 for comparison. The best fit age for this star is ∼5 Gyr.
Instead of log g, we can also use the I-band magnitude to plot the star on the H-R diagram.
We adopt a distance of 8.5 kpc, placing this dwarf on the far side of the Bulge, its most
likely position because the optical depth for lensing is larger there. Figure 5 shows that a
similar age is obtained. Indeed, shifting the position of MOA-2006-BLG-99S in the vertical
direction, by changing its distance or luminosity has little effect on the young age that we
derive for this star, because, at this metallicity, there should be no stars this hot with ages ≥
6 Gyr. However, the uncertainty in the temperature combined with the effect that changing
the temperature has on the derived metallicity produce large uncertainties in the age. If we
instead adopt the Teff on the lower edge of our range (5600K), the metallicity calculated
from the Fe I lines drops to [Fe/H]=0.16 dex. Using isochrones of this metallicity, the new
Teff gives an age of ∼9 Gyrs.
Improvements in the accuracy of temperatures are needed
to get better age constraints for Bulge dwarfs, but in principal, ages can be measured for
individual stars near the main-sequence turnoff.
5.2.1. Mixing in Giants in the Bulge
As stars move up the giant branch, they pass through first dredge-up, which brings up
material that has been processed in the CN cycle. The C and N abundances measured in
giants no longer represent the original C and N endowments of the stars, although C+N
will remain constant as long as only material processed in the CN cycle, and not the ON
cycle, is mixed to the surface. Cunha & Smith (2006) and Cunha et al. (2007) measured C
and N in giants in the Bulge. They found that the giants lie to the N-rich side of the line
defined by the C/N ratio of the Sun (Figure 6). Cunha et al. (2007) concluded that a small
amount of mixing had occurred in the giants. This conclusion is only valid if the original
abundances in the giants lie close to the line. Otherwise, if the Bulge dwarfs have non-solar
C/N ratios, the C/N ratios measured in Bulge giants could imply either no mixing (and a
N-rich original composition) or substantial mixing (and a C-rich original composition). The
abundances of C and N for MOA-2006-BLG-99S are also plotted on Figure 8 and shows that
-- 14 --
Fig. 3. -- The MDF for MOA-2006-BLG-99S and OGLE-2006-BLG-265S compared to the
MDF from Sadler et al. (1996), which was measured on K and M giants and with the MDFs
of Ram´ırez et al. (2000), Rich & Origlia (2005), and Rich et al. (2007) for M giants. The
MDF of the dwarfs is shifted to higher metallicities compared to the M giants, which is
surprising since the most metal-rich stars should end their red giant phase as M giants
rather than K giants.
-- 15 --
Fig. 4. -- The MDF for MOA-2006-BLG-99S and OGLE-2006-BLG-265S compared to the
MDF derived by Zoccali et al. (2008) from high-dispersion spectra of giants in three fields:
Baade's Window at 4◦ as well as a 6◦ and 12◦ field. The fraction of the two-bulge-dwarf
sample has been scaled down from 0.5 to fit clearly on the graph. The average metallicity
of the bulge giants decreases as the distance from the Galactic center. The dwarfs are more
than 4◦ away from the center, making their high metallicities even more surprising.
-- 16 --
Fig. 5. -- The position of MOA-2006-BLG-99S (square) in the H-R diagram (MI-Teff). MI
was calculated using the I0 magnitude and assuming a distance of 8.5 kpc. The Teff is the
temperature from the Balmer lines. We also show isochrones from Yi et al. (2001) The solid
lines show isochrones for [Fe/H]=0.385 and for ages 2, 3, 4, 5 and 7 Gyr. The dashed line
shows a 5 Gyr isochrone for a solar metallicity.
-- 17 --
the assumption of Cunha et al. (2007) is justified and that the expected amount of mixing
has occurred in the giants.
5.3. Lithium
Li has been created since the Big Bang by stellar nucleosynthesis and by cosmic ray
spallation. A Li abundance for a star in the Bulge, measuring how fast Li was made in the
early Galaxy, would be very interesting. However, most stars no longer have the same amount
of Li on their surfaces as was present in their natal gas clouds. Li is easily burned during
pre-main sequence and main-sequence phases of stars and is either destroyed throughout the
convective envelope during the RGB phase or (for a brief time) created in the star itself and
dredged up. We have no detection of Li in this star, only a 3-σ upper limit of logǫ(Li)=1.84
dex based on a χ2 fit to the data (see Johnson et al. 2007 for more details). In Figure 7 we
show this upper limit compared with Li measurements in open cluster stars having a range
of ages as well as field stars from Lambert & Reddy (2004). We also include the upper limit
from OGLE-2006-BLG-265S. Lower values can be expected in field stars because of astration
on the main-sequence and because they are often older than the clusters featured in Figure 7
and were formed out of gas that had not been polluted by as much Li. Figure 7 shows that
the Li upper limits in the bulge dwarfs are consistent with the upper limits in field dwarfs.
A dwarf with Teff> 7000K is probably needed to measure the amount of Li produced by
spallation in the Bulge.
5.4. Chemical Evolution of the Bulge
The Bulge has a different star formation history than the halo/disk. The ratios of
Type II/Type Ia pollution or Type II/AGB star pollution at a given [Fe/H] are therefore
different as well, and the abundance ratios reflect this. We compare the abundances for both
MOA-2006-BLG-99S and OGLE-2006-BLG-265S with Bulge giants and field stars from the
thick/thin disk and halo from literature sources. In Table 3, we summarize the literature
sources we use for each element.
5.4.1. Carbon and nitrogen
For many elements, observing red giant stars in the Bulge is an effective method of
measuring their abundances. However, as shown in §5.2.1, internal mixing on the RGB
-- 18 --
Fig. 6. -- The C and N abundance of MOA-2006-BLG-99S (filled red square) compared
with giants in the Bulge. Open green triangles show old giants from Cunha & Smith (2006)
and filled blue triangles show young supergiants from Cunha et al. (2007). The open black
circle shows the position of the Sun and the straight solid line marks where C/N ratio is
solar. The curved line represents constant C+N. The C and N abundances in the giants
can be explained by conversion of some C to N in CN processing from the solid line. The
data for MOA-2006-BLG-99S show that the solid line is a reasonable representation of the
"primordial" (unaffected by internal mixing) C/N ratio in the Bulge.
-- 19 --
logǫ(Li) for the Hyades and M67 from Balachandran (1995) and for field
Fig. 7. -- Teffvs.
stars from Lambert & Reddy (2004) compared with the upper limit for MOA-2006-BLG-99S
and OGLE-2006-BLG-265S. The bulge dwarf Li limits are consistent with disk stars, which
is not surprising given the age and temperature of these stars.
-- 20 --
alters the abundances of C and N. The abundances of C and N that we measure in MOA-
2006-BLG-99S therefore represent the first observations of the primordial C and N produced
by the chemical evolution of the Bulge.
MOA-2006-BLG-99S has [C/Fe]=0.04±0.22 and [N/Fe]=−0.06±0.43 (Fig. 8). The solar
values of [C/Fe] and N/Fe] show that C and N production kept pace with the Fe production
in the Bulge. There are many sources of C in the Universe (Gustafsson et al. 1999, e.g.).
Type II SNe and AGB stars certainly contribute substantial amounts of C and N; the roles of
novae and Wolf-Rayet stars are less clear. These contributions, whatever they are, track the
production of Fe in the chemical evolution of the Bulge. Finally, we attempted to measure
the 12C/13C ratio, which is sensitive to the source of C, being low for low-mass AGB stars
and high for Type II SNe. We could only set an uninteresting limit (12C/13C ¿ 1) on this
very interesting number.
5.5. Sodium and Aluminum
The abundances of Na and Al are elevated in Bulge giants (McWilliam & Rich 1994;
Lecureur et al. 2007) compared to disk stars. Metal-rich Type II SNe are predicted to
produce more of the odd-Z elements such as Na and Al than metal-poor Type II SNe and
could potentially be the explanation of the difference between the Bulge and the disk. In
Figure 9, we show the [Na/Fe] and [Al/Fe] value for MOA-2006-BLG-99S and OGLE-2006-
BLG-265S. These two unmixed stars are on the lower end of the scatter seen in the giants.
This could be a hint that the larger [Na/Fe] and [Al/Fe] values seen in giants are due to
internal mixing, but because the dwarf values fall within the scatter outlined in the scatter
indicates that more measurements in dwarf stars are needed before any differences in the
distribution of [Na/Fe] and [Al/Fe] in dwarfs and giants can be seen.
5.6. The α elements
Figure 10 shows the [O/Fe], [Mg/Fe], [Si/Fe], and [Ca/Fe] for MOA-2006-BLG-99S
compared with halo stars, thin and thick disk stars, and Bulge giants. The metallicity
of MOA-2006-BLG-99S is in the range of metallicities measured for Bulge giants, allowing
direct comparison of [α/Fe] ratios between dwarfs and giants. The agreement is good, as
both the giants and the dwarf have [α/Fe] below the high [α/Fe] values of the more metal-
poor ([Fe/H]≤0) stars. This decline in all [α/Fe] ratios suggests that Fe from Type Ia SNe is
being added and that the more metal-rich stars formed sufficiently later to have this ejectum
-- 21 --
Fig. 8. -- [C/Fe] and [N/Fe] for MOA-2006-BLG-99S (filled black square) compared to Bulge
giants (filled blue circles) and disk dwarfs (open red triangles). The low [C/Fe] values for
the Bulge giants are the result of internal mixing. The [C/Fe] and [N/Fe] values in MOA-
2006-BLG-99S, on the other hand, are the result of the pollution of the gas of the Bulge by
previous generations of stars. The solar ratios for these elements impose constraints on the
inefficiency of C and N production in the Bulge.
-- 22 --
Fig. 9. -- [Na/Fe] and [Al/Fe] for MOA-2006-BLG-99S and OGLE-2006-BLG-265S (filled
black squares) compared to Bulge giants (filled blue circles) and field stars (open red tri-
angles). The dwarfs fall within the distribution of [Na/Fe] and [Al/Fe] values seen in the
giants, but at the lower edge of that distribution.
in their gas.
-- 23 --
5.7. Potassium
K is an odd-Z element and is predicted in nucleosynthesis models to be underproduced
relative to the α elements in metal-poor SNe. There are few measurements in the literature,
and those that exist show the opposite trend of increasing [K/Fe] with decreasing [Fe/H]
(Gratton & Sneden 1987; Chen et al. 2000; Cayrel et al. 2004). However, the only K line
available for study in most stars, the resonance line at 7698A, is affected by non-LTE effects,
and these corrections have not yet been applied to large samples. Zhang et al. (2006) derived
NLTE corrections for each star in their sample. The [K/Fe] values were still supersolar at
low metallicities, with the thin disk stars showing a drop in [K/Fe] for[Fe/H]≥ −1. The
[K/Fe] ratios in the thick disk stars remain high. However, the [K/Mg] ratios showed much
smaller variations among the different Galactic populations. They argued that the constant
[K/Mg] ratio ([K/Mg]=−0.08 ± 0.01) in the stars indicated that the nucleosynthesis of K is
closely coupled to that of the α-elements, which is somewhat surprising given the theoretical
predictions.
We measured the 7698A line in MOA-2006-BLG-99 and the Sun. We also measured the
K abundance in OGLE-2006-BLG-265S using TurboSpectrum and the model atmosphere
described in Johnson et al. (2007). The [K/Fe] we measure for OGLE-2006-BLG-265S is
−0.08 ± 0.20 Figure 11 compares the results for the Bulge to the Zhang et al. (2006) results.
We applied no NLTE correction, but assumed that the LTE abundances in the Sun and
MOA-2006-BLG-99S would be affected by the same amount.
The [K/Mg] values are 0.07±0.23 for MOA-2006-BLG-99S and 0.12±0.27 for OGLE-
2006-BLG-265S; the [K/Mg] abundance is still within a narrow range, even in this very
different chemical evolution history.
5.8.
Iron-peak elements
The abundances for Ti, Sc, Mn and Ni are shown in Figure 12. McWilliam & Rich
(1994) found that [Ti/Fe] behaves like [O/Fe] in the Bulge, with supersolar [Ti/Fe] ratios
for many stars, followed by [Ti/Fe] decreasing to solar for [Fe/H> 0. The abundance of Ti
is also enhanced in halo stars, leading it to be classified as an "α-element" for observational
purposes.
In MOA-2006-BLG-99S, [Ti/Fe] is close to solar, in line with the α-elements
discussed above. The ratios of [Sc/Fe] and [Ni/Fe] are observed to be close to solar for a
-- 24 --
Fig. 10. -- [O/Fe], [Mg/Fe], [Si/Fe] and [Ca/Fe] for MOA-2006-BLG-99S and OGLE-2006-
BLG-265S (filled black squares) compared to Bulge giants (filled blue circles) and field stars
(open red triangles). The [α/Fe] ratios in MOA-2006-BLG-99S agree well with the values
measured in giants of similar metallicity.
-- 25 --
Fig. 11. -- [K/Fe] for MOA-2006-BLG-99S and OGLE-2006-BLG-265S (filled black squares)
compared to field stars (open red triangles). The [K/Fe] values in the dwarfs fall in the range
seen in the disk stars.
-- 26 --
wide range of populations: halo, thick and thin disk and Bulge. The data for MOA-2006-
BLG-99S show that abundances in this Bulge dwarf agree with this picture.
For [Fe/H] < −1.5 in the Galactic halo/disk, there is a plateau at [Mn/Fe]∼ −0.5
(McWilliam et al. 1995, 2003). Because Type Ia SNe have not polluted the most metal-poor
stars in the Galaxy, we can derive the ratio of Mn/Fe produced in (metal-poor) Type II SNe
from this plateau.
[Mn/O] starts to rise before [O/Fe] starts to drop in the disk. Because
the drop in [O/Fe] signals the onset of substantial Type Ia SN contribution, the rise in Mn
relative to O cannot be due to Type Ia SNe, but rather to increased production of Mn by
more metal-rich Type II SNe (Feltzing et al. 2007). McWilliam et al. (2003) also measured
Mn in 13 stars in the Sagittarius dwarf galaxy. These stars have substantial Type Ia SNe
contributions to their gas, but [Mn/Fe] values about 0.2 dex below the trend seen in the
Galactic disk, providing additional evidence that metal-rich Type II SNe are responsible for
Mn production. The near solar [Mn/Fe] values for MOA-2006-BLG-99S and OGLE-99 are
also in support of this idea, because they were seen in an environment that had many Type
II SNe occur, unlike the Sagittarius stars.
5.9. Copper and Zinc
The percent of Cu and Zn production to be ascribed to different nucleosynthesis sites
(e.g., Type II SNe, AGB stars, Type Ia SNe) is uncertain. The observations show that
[Cu/Fe] ∼ −1 at the lowest metallicities and then rises to solar by [Fe/H]∼ −0.8 (e.g.
Mishenina et al. 2002). [Zn/Fe], on the other hand, has supersolar values at the lowest metal-
licities and then decreases to closer to solar (e.g. Mishenina et al. 2002). Matteucci et al.
(1993) used new weak s-process calculations and available SN models to argue that approx-
imately two-thirds of the Zn and Cu production in the Universe is due to Type Ia SNe.
The rest of the Zn is from a primary process in massive stars, while the Cu comes from a
secondary (=metallicity-dependent) process in massive stars. Using this model and a chem-
ical evolution model for the Bulge, Matteucci et al. (1999) predicted that both [Cu/Fe] and
[Zn/Fe] would be ∼ 0.2 at [Fe/H]=0.3. The SN models available to Matteucci et al. (1993)
did not include important effects, such as detailed calculation of neutron-capture elements
beyond Fe. Bisterzo et al. (2005) used updated results and considered Zn and Cu production
by neutron-capture in the O-rich parts of Type II SNe ("weak sr-process"). In their analy-
sis, Cu is mostly produced in this weak sr-process, a secondary process. A small amount of
primary Cu is made as radioactive Zn in the inner regions of Type II SNe. Zn production
is also due to massive star nucleosynthesis, but here there is a large primary production
in the α-rich freeze out in Type II SNe, which is supplemented at higher metallicities by
-- 27 --
Fig. 12. -- [Sc/Fe], [Ti/Fe], [Mn/Fe], and [Ni/Fe] for MOA-2006-BLG-99S and OGLE-2006-
BLG-265S (filled black squares) compared to Bulge giants (filled blue circles) and field stars
(open red triangles).
The [Ti/Fe] value in MOA-2006-BLG-99S agrees well with [Ti/Fe] ratios measured in bulge
giants of similar metallicity. Within the error bars, the other [iron-peak/Fe] ratios follow the
trends seen the disk stars.
-- 28 --
Fig. 13. -- [Cu/Fe] and [Zn/Fe] for MOA-2006-BLG-99S and OGLE-2006-BLG-265S (filled
black squares) compared to field stars (open red triangles). The Matteucci et al. (1999)
models predict [Zn/Fe]≈ 0.2 for high-metallicity stars in the Bulge. Therefore, the low
[Zn/Fe] disagrees with the theory that Zn is produced in large amounts of Type Ia SNe.
The solar and supersolar [Cu/Fe] values in the Bulge are consistent with either Type Ia SN
production or metal-rich Type II SN production.
-- 29 --
a secondary contribution from the weak-sr process. Bisterzo et al. (2005) find no need for
contributions to Cu and Zn from Type Ia SNe or AGB stars.
The measurements of [Cu/Fe] and [Zn/Fe] in MOA-2006-BLG-99S support the Bisterzo et al.
(2005) model and argue against the production of large amounts of Zn in Type Ia SNe.
Because the abundance of Cu is dominated by a secondary process, the solar [Cu/Fe] at
[Fe/H]=0.36 and the supersolar value at [Fe/H]=0.56 are the result of copious Cu produc-
tion in metal-rich Type II SNe. However, the primary production of Zn and the smaller
secondary contribution is not sufficient to keep up with the Fe from both Type II SNe and
Type Ia SNe. The observations also show again that the chemical evolution of Zn is separate
from Fe.
5.10. Barium
The [Ba/Fe] for MOA-2006-BLG-99S falls considerably below the solar value ([Ba/Fe]=−0.61
(Fig. 14). It falls in the range not seen in other parts of Galaxy, except for the metal-poor
halo. We note that we could only measure 1 line of Ba in MOA-2006-BLG-99S, and the solar
value we measure is the most discrepant from the solar value in Grevesse & Sauval (1998).
However, even if the Grevesse & Sauval value is used, the [Ba/Fe] for MOA-2006-BLG-99S
is still subsolar ([Ba/Fe]=−0.24). If we use the solar value of Ba (log(ǫ(Ba)=2.39) reported
in Johnson et al. (2007), OGLE-2006-BLG-265S has a [Ba/Fe]= −0.28.
The paucity of Ba in the halo stars is explained by the fact that the r-process is the only
available channel for producing the heavy elements in the early Univere, and the r-process
is not an efficient producer of Ba.
It is tempting to ascribe the low Ba in MOA-2006-
BLG-99S to the same cause, especially in light of the high [Eu/Fe] measurements in Bulge
giants by McWilliam & Rich (1994). Stellar populations dominated by Type II SNe and r-
process production should have high [Eu/Fe] and [Eu/Ba] before Type Ia SNe and AGB stars
eventually add Fe and Ba to the ISM. Whether the Ba deficiency in MOA-2006-BLG-99S can
be explained by a lack of contributions from AGB stars depends on the relative timescales
of Type Ia SN pollution and AGB pollution and whether the low [α/Fe] values in MOA-
2006-BLG-99S are the result of Type Ia pollution.
If we assume that Type Ia SNe have
contributed significantly to the abundances of MOA-2006-BLG-99S, which is reasonable,
then AGB pollution must trail Type Ia SN production. The evidence on this point is mixed.
Simmerer et al. (2004) saw, in addition to a wide range at any given metallicity, a rise in
the [La/Fe] ratios at [Fe/H] > −2. Because La, like Ba, is mostly due to the s-process,
this would indicates the s-process from AGB stars is added before Fe from Type Ia SNe
causes the [α/Fe] ratios to turn over. On the other hand, Mel´endez & Cohen (2007) argued
-- 30 --
Fig. 14. -- [Ba/Fe] for MOA-2006-BLG-99S and OGLE-2006-BLG-265S (filled black squares)
compared to field stars (open red triangles).
[Ba/Fe] shows the largest deviation from the
trends seen in the disk stars of any element studied in this paper. This low Ba value is
consistent with the idea that the r-process is the dominate producer of heavy elements in
the Bulge. It also puts constraints on s-process contributions to Ba (and the accompanying
C and N contributions) from AGB stars.
-- 31 --
based on the isotope ratios of Mg that 3-6 M⊙ stars did not start contributing to the halo
until [Fe/H]≥ 1.5, at the same time or later than the Type Ia SNe. The [Ba/Fe] measured in
OGLE-2006-BLG-265S is closer to the solar value, and suggests that by [Fe/H]=0.5 the Bulge
had reached the same point in chemical evolution as the solar neighborhood, with substantial
amounts of Ba supplied by the s-process in AGB stars balancing the iron supplied by Type
Ia SNe. Ba abundances for Bulge stars with a wide range of metallicity would help clarify
the origin of the Ba abundance in the Bulge.
6. Conclusions
The results for MOA-2006-BLG-99S demonstrate the unique information that can be ob-
tained from high-resolution spectra of microlensed dwarfs in the Bulge. Because microlensing
is not biased in metallicity, we can measure the MDF of the Bulge main-sequence stars when
sufficient number of highly magnified Bulge dwarfs have been observed with high-dispersion
spectrographs. However, the two dwarfs we have studied so far both have [Fe/H]> 0.30,
which makes them more metal-rich than any of the M giants observed at high-resolution.
If the temperatures and metallicities measured for the two dwarfs are accurate, then these
stars are younger than the bulk of the Bulge population.
For many elements, the abundance ratios we measure in MOA-2006-BLG-99S are not
surprising. The iron-peak and α-elements (O, Mg, Si, Ca, and Ti) fall on the trend observed
in the Bulge giants. The solar values for the [α/Fe] ratios suggest that Type Ia SN ejecta are
responsible for some of the Fe, which would be expected if these metal-rich stars were part
of a younger population in the Bulge. The [C/Fe] and [N/Fe] ratios are also ∼ 0, suggesting
that AGB stars have also begun to contribute these light elements to offset the contribution
of Fe from Type Ia SNe. It is therefore unexpected to find [Ba/Fe]∼-0.5. Such low [Ba/Fe]
values, when found in the halo, imply that only Type II SNe and the r-process have polluted
the gas, and not the s-process from AGB stars.
The C and N abundances we measure in MOA-2006-BLG-99S are unaffected by any
mixing or dilution and therefore are the result of the chemical evolution of the Bulge. Com-
paring the total amount of C+N to that measured in Bulge giants, we conclude that only
mild mixing, as expected in theoretical models, has occurred in the Bulge. This agrees with
the assertions by Lecureur et al. (2007), Zoccali et al. (2006), and McWilliam et al. (2007)
that the O, Na and Al abundances in Bulge giants should not have been affected by mixing.
The [Na/Fe] and [Al/Fe] ratios for MOA-2006-BLG-99S fall on the lower end of the values
seen in Bulge giants of this metallicity. The [Na/Fe] and [Al/Fe] ratios in giants are marked
above all else by large scatter and more abundance ratios in dwarfs are needed before a
-- 32 --
comparison of scatter can be made.
We can use our measurements of the abundances of K, Cu and Zn, rare for Bulge stars,
to exploit the different star formation history of the Bulge to study the nucleosynthesis sites
of these elements. The supersolar [Cu/Fe] and subsolar [Zn/Fe] values agree with the model
of Bisterzo et al. (2005) and the production of these two elements predominantly in Type II
SNe. The chemical evolution of K is more difficult to understand, because the [K/Mg] value
in the Bulge is similar to that in the metal-poor halo, although the production of K should
be enhanced in metal-rich SNe.
Finally, we note that the standard microlensing technique of estimating the intrinsic
color and magnitude of the source star by using the offset from the position of the red
clump can be tested by deriving a temperature and gravity from the spectrum. For MOA-
2006-BLG-99S, the photometric and spectroscopic temperature agree very well, although the
agreement was considerably less good for OGLE-2006-BLG-265S.
In summary, the abundances in MOA-2006-BLG-99S provide unique information about
the formation and evolution of the Bulge. Larger samples of dwarfs observed in this way
would allow the derivation of the dwarf MDF and the trends in abundance ratios over a
wide range of metallicity. By taking advantage of microlensing events in the Bulge, we can
achieve this goal with a modest amount of telescope time.
Our thanks to Ian Thompson for help with the echelle observations at Las Campanas.
Thomas Masseron and Bertrand Plez provided invaluable support in installing and running
TurboSpectrum. We would also like to thank Daniel Kelson for codes and support for reduc-
ing the MIKE data. Jon Fulbright, Manuela Zoccali, and Solange Rami´rez kindly provided
metallicity data for the bulge giant metallicity distribution functions. We acknowledge sup-
port from: NSF AST-042758 and NASA NNG04GL51G (AG). Work by B.S.G. was partially
supported by a Menzel Fellowship from the Harvard College Observatory.
-- 33 --
Table 1. Line Parameters and Equivalent Widths
Ion Wavelength
E.P.
log gf
EWstar
EWsun
Source
O I
O I
O I
Na I
Na I
Na I
Na I
Mg I
Mg I
Al I
Si I
Si I
Si I
Si I
Si I
Si I
Si I
Si I
Si I
Si I
Si I
Si I
K I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
7771.941
7774.161
7775.390
5682.633
5688.194
6154.226
6160.747
5711.088
7387.689
7836.134
5665.555
5690.425
5772.146
6125.021
6142.483
6155.134
6237.319
6244.466
7415.948
7918.384
7932.348
7944.001
7698.974
5581.965
5590.114
6161.297
6166.439
6169.042
6169.563
6455.598
6471.662
0.370
9.15
0.220
9.15
0.000
9.15
2.10 −0.710
2.10 −0.400
2.10 −1.570
2.10 −1.270
4.33 −1.870
5.75 −1.270
4.02 −0.650
4.93 −1.940
4.93 −1.770
5.08 −1.650
5.61 −1.520
5.62 −1.500
5.62 −0.720
5.61 −1.050
5.62 −1.320
5.61 −0.850
5.95 −0.510
5.96 −0.370
5.98 −0.210
0.00 −0.170
2.52 −0.530
2.52 −0.571
2.52 −1.266
2.52 −1.142
2.52 −0.797
2.52 −0.478
2.52 −1.290
2.52 −0.686
101.5
86.7
71.1
148.1
150.6
60.9
82.3
140.6
141.2
93.1
89.2
57.4
69.9
57.7
61.8
124.5
90.9
77.6
115.0
157.5
131.8
140.8
syn
111.8
148.4
77.3
83.2
110.2
147.8
59.6
116.1
70.2
60.1
47.4
96.9
118.7
35.9
55.8
112.7
70.6
54.2
40.2
49.7
51.9
30.4
32.3
84.5
63.9
47.3
85.5
84.5
92.3
107.0
syn
94.7
91.0
59.2
70.2
88.7
105.2
56.9
90.2
1
1
1
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
1
3
3
3
3
3
3
3
3
-- 34 --
Table 1 -- Continued
Ion Wavelength
E.P.
log gf
EWstar
EWsun
Source
Ca I
Sc II
Sc II
Sc II
Ti I
Ti I
Ti I
Ti I
Ti II
Mn I
Mn I
Mn I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
6499.650
5526.820
5657.880
6245.620
5866.451
6126.216
6258.102
6261.098
5381.015
5470.270
6013.520
6021.710
5307.361
5321.108
5322.041
5329.989
5379.574
5501.465
5506.779
5701.545
5705.465
6027.051
6082.711
6165.360
6173.336
6219.281
6252.555
6265.134
6297.793
6301.501
6311.500
2.52 −0.818
1.77
0.02
1.51 −0.600
1.51 −1.070
1.07 −0.780
1.07 −1.370
1.44 −0.300
1.43 −0.420
1.57 −1.920
2.14 −1.460
3.07 −0.250
3.07
0.03
1.61 −2.987
4.43 −0.951
2.28 −2.803
4.08 −1.189
3.70 −1.510
0.96 −3.047
0.99 −2.797
2.56 −2.120
4.30 −1.500
4.08 −1.090
2.22 −3.573
4.14 −1.470
2.22 −2.880
2.20 −2.420
2.40 −1.690
2.18 −2.550
2.22 −2.740
3.65 −0.718
2.83 −3.141
97.5
syn
syn
syn
60.6
59.7
72.9
61.1
80.4
syn
syn
syn
105.7
65.1
105.4
128.7
74.2
127.1
149.4
114.7
54.3
78.4
63.5
72.3
119.9
114.4
148.8
107.8
94.6
141.5
49.0
84.1
syn
syn
syn
48.1
23.4
49.8
50.1
59.3
syn
syn
syn
84.5
40.8
59.8
55.7
60.5
115.4
115.1
81.9
38.7
65.1
34.0
43.7
68.7
86.6
115.3
87.2
72.8
121.6
27.1
3
5
5
5
2
2
2
2
4
5
5
5
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
-- 35 --
Table 1 -- Continued
Ion Wavelength
E.P.
log gf
EWstar
EWsun
Source
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe II
Fe II
Fe II
Fe II
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
6322.685
6335.331
6344.149
6380.743
6408.018
6411.649
6498.939
6592.914
6593.870
6804.001
6810.263
6855.162
6858.150
7401.685
7710.364
7941.089
6247.557
6416.919
6432.680
6456.383
5754.655
5805.213
6108.107
6111.066
6128.963
6175.360
6176.807
6314.653
6378.247
6482.796
6598.593
2.59 −2.426
2.20 −2.180
2.43 −2.920
4.19 −1.380
3.69 −1.018
3.64 −0.720
0.96 −4.700
2.73 −1.470
2.43 −2.420
4.65 −1.496
4.61 −0.986
4.56 −0.740
4.61 −0.930
4.19 −1.350
4.22 −1.110
3.27 −2.580
3.89 −2.430
3.89 −2.880
2.89 −3.690
3.90 −2.190
1.93 −1.850
4.17 −0.620
1.68 −2.430
4.09 −0.820
1.68 −3.360
4.09 −0.500
4.09 −0.260
1.94 −2.000
4.15 −0.810
1.93 −2.760
4.24 −0.910
100.1
127.5
91.7
82.6
131.7
139.5
49.8
129.6
117.3
67.2
71.4
102.2
79.5
58.7
103.9
59.8
86.3
65.0
61.3
71.7
133.5
60.5
68.5
70.0
46.8
63.5
95.1
92.1
59.5
54.8
42.0
79.0
95.8
58.4
51.6
94.7
128.7
45.0
107.3
81.7
21.1
48.9
69.0
50.8
39.7
63.5
42.0
51.8
41.0
40.0
61.2
75.4
42.1
63.3
32.7
24.8
46.9
63.4
72.4
29.9
40.7
24.8
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
-- 36 --
Table 1 -- Continued
Ion Wavelength
E.P.
log gf
EWstar
EWsun
Source
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Ni I
Cu I
Zn I
Zn I
Ba II
6643.629
6767.768
6772.313
7110.892
7715.583
7748.891
7788.936
7797.586
5782.050
4722.153
4810.528
6496.410
1.68 −1.910
1.83 −2.100
3.66 −0.940
1.94 −2.880
3.70 −1.010
3.71 −0.330
1.95 −1.750
3.90 −0.320
1.64 −2.920
4.03 −0.338
4.08 −0.137
0.60 −0.380
122.2
106.2
83.8
57.8
105.7
164.4
123.4
103.6
syn
syn
syn
syn
95.0
80.5
49.8
36.0
49.7
84.0
92.2
79.2
syn
syn
syn
syn
2
2
2
2
2
2
2
2
5
1
1
5
References. -- (1) VALD database, Piskunov et al. (1995), (2)
Bensby et al. (2003), (3) Smith & Raggett (1981), (4) Pickering et al.
(2001), (5)Johnson et al. (2006) and references therein
-- 37 --
Table 2. Abundances
Ion
logǫ
σǫ
[X/FeI]a σ[X/F eI]
σ
Nlines
Solar
meas. GS98
C (CH)
N (CN)
O I
Na I
Mg I
Al I
Si I
K I
Ca I
Sc II
Ti I
Ti II
Mn I
Fe I
Fe II
Ni I
Cu I
Zn I
Ba II
8.8
8.2
9.09
6.69
8.18
6.98
7.98
5.60
6.58
3.25
5.25
5.18
5.62
7.81
7.81
6.61
4.40
4.63
2.25
0.26
0.56
0.11
0.13
0.21
0.16
0.10
0.28
0.29
0.31
0.23
0.36
0.13
0.18
0.28
0.10
0.23
0.23
0.43
0.04
−0.06
−0.16
0.09
0.16
0.11
0.06
0.14
−0.09
−0.30
−0.05
−0.22
−0.09
0.36
−0.18
0.02
−0.01
−0.28
−0.61
0.22
0.43
0.22
0.16
0.25
0.22
0.13
0.29
0.16
0.23
0.17
0.28
0.12
. . .
0.19
0.09
0.20
0.21
0.32
0.20
0.20
0.33
0.12
0.17
0.10
0.10
0.20
0.23
0.13
0.27
0.20
0.08
0.25
0.25
0.26
0.10
0.10
0.20
. . .
. . .
3
4
2
1
12
1
9
3
4
1
3
35
4
19
1
2
1
8.4
7.9
8.89
6.24
7.66
6.51
7.56
5.10
6.31
3.19
4.94
5.04
5.34
7.45
7.63
6.23
4.05
4.55
2.50
8.52
7.92
8.83
6.33
7.58
6.47
7.55
5.12
6.36
3.17
5.02
5.02
5.39
7.50
7.50
6.25
4.21
4.60
2.13
a[X/Fe I] given for all species except Fe I, where [Fe I/H] is given.
Table 3. Literature Sources
Source
C
N O Na Mg
Al
Si K Ca
Sc
Ti Mn
Ni
Cu
Zn
Ba
Fulbright et al. (2007)
Rich & Origlia (2005)
Rich et al. (2007)
Lecureur et al. (2007)
Cunha & Smith (2006)
Reddy et al. (2006)
Reddy et al. (2003)
Feltzing et al. (2007)
Bensby & Feltzing (2006)
Bensby et al. (2005)
Bensby et al. (2004)
Bensby et al. (2003)
Chen et al. (2004)
Carretta et al. (2000)
Zhang et al. (2006)
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
Bulge Stars
x
x
x
x
x
Halo/Disk Data
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
--
3
8
--
-- 39 --
REFERENCES
Alvarez, R., & Plez, B. 1998, A&A, 330, 1109
Balachandran, S. 1995, ApJ, 446, 203
Alonso, A., Arribas, S., & Martinez-Roger, C. 1996, A&A, 313, 873
Baade, W. 1946, PASP, 58, 249
Barklem, P. S., Piskunov, N., & O'Mara, B. J. 2000, A&AS, 142, 467
Barklem, P. S., Stempels, H. C., Allende Prieto, C., Kochukhov, O. P., Piskunov, N., &
O'Mara, B. J. 2002, A&A, 385, 951
Bensby, T., Feltzing, S., & Lundstrom, I. 2003, A&A, 410, 527
Bensby, T., Feltzing, S., & Lundstrom, I. 2004, A&A, 415, 155
Bensby, T., & Feltzing, S. 2006, MNRAS, 367, 1181
Bensby, T., Feltzing, S., Lundstrom, I., & Ilyin, I. 2005, A&A, 433, 185
Bernstein, R., Shectman, S. A., Gunnels, S. M., Mochnacki, S., & Athey, A. E. 2003, Pro-
ceedings of the SPIE, 4841, 1694
Bisterzo, S., Pompeia, L., Gallino, R., Pignatari, M., Cunha, K., Heger, A., & Smith, V.
2005, Nuclear Physics A, 758, 284
Carretta, E., Gratton, R. G., & Sneden, C. 2000, A&A, 356, 238
Castelli, F., & Kurucz, R. L. 2003, Modelling of Stellar Atmospheres, 210, 20
Cayrel, R., et al. 2004, A&A, 416, 1117
Chen, Y. Q., Nissen, P. E., Zhao, G., Zhang, H. W., & Benoni, T. 2000, A&AS, 141, 491
Chen, Y. Q., Zhao, G., Nissen, P. E., Bai, G. S., & Qiu, H. M. 2003, ApJ, 591, 925
Chen, Y. Q., Nissen, P. E., & Zhao, G. 2004, A&A, 425, 697
Cunha, K., & Smith, V. V. 2006, ApJ, 651, 491
Cunha, K., Sellgren, K., Smith, V. V., Ramirez, S. V., Blum, R. D., & Terndrup, D. M.
2007, ArXiv e-prints, 707, arXiv:0707.2610
-- 40 --
Demarque, P., Woo, J.-H., Kim, Y.-C., & Yi, S. K. 2004, ApJS, 155, 667
Feltzing, S., Fohlman, M., & Bensby, T. 2007, A&A, 467, 665
Feltzing, S., & Gilmore, G. 2000, A&A, 355, 949
Fulbright, J. P., McWilliam, A., & Rich, R. M. 2006, ApJ, 636, 821
Fulbright, J. P., McWilliam, A., & Rich, R. M. 2007, ApJ, 661, 1152
Gould, A., et al. 2006, ApJ, 644, L37
Gratton, R. G., & Sneden, C. 1987, A&A, 178, 179
Grevesse, N., & Sauval, A. J. 1998, Space Science Reviews, 85, 161
Griest, K. & Safizadeh, N. 1998, ApJ, 500, 37
Gustafsson, B., Karlsson, T., Olsson, E., Edvardsson, B., & Ryde, N. 1999, A&A, 342, 426
Holtzman, J.A., Watson, A.M., Baum, W.A., Grillmair, C.J., Groth, E.J., Light, R.M.,
Lynds, R., & O'Neil, E.J. 1998, AJ, 115, 1946
Johnson, J. A., Gal-Yam, A., Leonard, D. C., Simon, J. D., Udalski, A., & Gould, A. 2007,
ApJ, 655, L33
Johnson, J. A., Ivans, I. I., & Stetson, P. B. 2006, ApJ, 640, 801
Kurucz, R. L., Furenlid, I., Brault, J., & Testerman, L. 1984, National Solar Observatory
Atlas, Sunspot, New Mexico: National Solar Observatory, 1984,
Lambert, D. L., & Reddy, B. E. 2004, MNRAS, 349, 757
Lecureur, A., Hill, V., Zoccali, M., Barbuy, B., G´omez, A., Minniti, D., Ortolani, S., &
Renzini, A. 2007, A&A, 465, 799
Lecureur, A. et al. 2008, in prep
McWilliam, A., & Rich, R. M. 1994, ApJS, 91, 749
McWilliam, A., Preston, G. W., Sneden, C., & Searle, L. 1995, AJ, 109, 2757
McWilliam, A., Rich, R. M., & Smecker-Hane, T. A. 2003, ApJ, 592, L21
McWilliam, A., Matteucci, F., Ballero, S., Rich, R. M., Fulbright, J. P., & Cescutti, G. 2007,
ArXiv e-prints, 708, arXiv:0708.4026
-- 41 --
Mao, S. & Paczy´nski, B. 1991, ApJ, 374, L37
Mashonkina, L., Gehren, T., & Bikmaev, I. 1999, A&A, 343, 519
Matteucci, F., Raiteri, C. M., Busson, M., Gallino, R., & Gratton, R. 1993, A&A, 272, 421
Matteucci, F., Romano, D., & Molaro, P. 1999, A&A, 341, 458
Mel´endez, J., & Cohen, J. G. 2007, ApJ, 659, L25
Mishenina, T. V., Kovtyukh, V. V., Soubiran, C., Travaglio, C., & Busso, M. 2002, A&A,
396, 189
Moore, C. E., Minnaert, M. G. J. & Houtgast, J. 1966. The Solar Spectrum 2935A- 8770 A,
US Govt. Printing Office, Washington, D. C.
Ortolani, S., Renzini, A., Gilmozzi, R., Marconi, G., Barbuy, B., Bica, E., & Rich, R. M.
1995, Nature, 377, 701
Pickering, J. C., Thorne, A. P., & Perez, R. 2001, ApJS, 132, 403
Piskunov, N. E., Kupka, F., Ryabchikova, T. A., Weiss, W. W., & Jeffery, C. S. 1995, A&AS,
112, 525
Ram´ırez, S. V., Stephens, A. W., Frogel, J. A., & DePoy, D. L. 2000, AJ, 120, 833
Rich, R. M., & Origlia, L. 2005, ApJ, 634, 1293
Rich, R. M., Origlia, L., & Valenti, E. 2007, ApJ, 665, L119
Ram´ırez, I., & Mel´endez, J. 2005, ApJ, 626, 465
Reddy, B. E., Lambert, D. L., & Allende Prieto, C. 2006, MNRAS, 367, 1329
Reddy, B. E., Tomkin, J., Lambert, D. L., & Allende Prieto, C. 2003, MNRAS, 340, 304
Rich, R. M. 1988, AJ, 95, 828
Sadler, E. M., Rich, R. M., & Terndrup, D. M. 1996, AJ, 112, 171
Shetrone, M. D. 1996, AJ, 112, 1517
Simmerer, J., Sneden, C., Cowan, J. J., Collier, J., Woolf, V. M., & Lawler, J. E. 2004, ApJ,
617, 1091
-- 42 --
Smith, G., & Raggett, D. S. J. 1981, Journal of Physics B Atomic Molecular Physics, 14,
4015
Terndrup, D. M. 1988, AJ, 96, 884
Udalski, A., et al. 2005, ApJ, 628, L109
Whitford, A. E., & Rich, R. M. 1983, ApJ, 274, 723
Yi, S., Demarque, P., Kim, Y.-C., Lee, Y.-W., Ree, C. H., Lejeune, T., & Barnes, S. 2001,
ApJS, 136, 417
Yoo, J., et al. 2004a, ApJ, 603, 139
Zhang, H. W., Gehren, T., Butler, K., Shi, J. R., & Zhao, G. 2006, A&A, 457, 645
Zoccali, M., et al. 2003, A&A, 399, 931
Zoccali, M., et al. 2006, A&A, 457, L1
Zoccali, M., et al. 2008, in prep
This preprint was prepared with the AAS LATEX macros v5.2.
|
astro-ph/9803132 | 1 | 9803 | 1998-03-11T17:53:35 | The poor cluster of galaxies S639 | [
"astro-ph"
] | We have studied the poor southern cluster of galaxies S639. Based on new Stromgren photometry of stars in the direction of the cluster we confirm that the galactic extinction affecting the cluster is large. We find the extinction in Johnson B to be A_B = 0.75+-0.03. We have obtained new photometry in Gunn r for E and S0 galaxies in the cluster. If the Fundamental Plane is used for determination of the relative distance and the peculiar velocity of the cluster we find a distance, in velocity units, of 5706+-350 km/s, and a substantial peculiar velocity, 839+-350 km/s. However, the colors and the absorption line indices of the E and S0 galaxies indicate that the stellar populations in these galaxies are different from those in similar galaxies in the two rich clusters Coma and HydraI. This difference may severely affect the distance determination and the derived peculiar velocity. The data are consistent with a non-significant peculiar velocity for S639 and the galaxies in the cluster being on average 0.2 dex younger than similar galaxies in Coma and HydraI. The results for S639 caution that some large peculiar velocities may be spurious and caused by unusual stellar populations. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 29 December 2017
(MN LATEX style file v1.4)
The poor cluster of galaxies S639
Inger Jørgensen1,⋆ † and Helge Jønch-Sørensen2,⋆
1McDonald Observatory, The University of Texas at Austin, RLM 15.308, Austin, TX 78712, USA
2Copenhagen University Observatory, Juliane Maries Vej 30, DK-2100 Copenhagen Ø, Denmark
March 11, 1998, accepted for publication in MNRAS
8
9
9
1
r
a
M
1
1
1
v
2
3
1
3
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
ABSTRACT
We have studied the poor southern cluster of galaxies S639. Based on new Stromgren
photometry of stars in the direction of the cluster we confirm that the galactic ex-
tinction affecting the cluster is large. We find the extinction in Johnson B to be
AB = 0.75 ± 0.03. We have obtained new photometry in Gunn r for E and S0 galaxies
in the cluster. If the Fundamental Plane is used for determination of the relative dis-
tance and the peculiar velocity of the cluster we find a distance, in velocity units, of
(5706 ± 350)km s−1, and a substantial peculiar velocity, (839 ± 350)km s−1. However,
the colors and the absorption line indices of the E and S0 galaxies indicate that the
stellar populations in these galaxies are different from those in similar galaxies in the
two rich clusters Coma and HydraI. This difference may severely affect the distance
determination and the derived peculiar velocity. The data are consistent with a non-
significant peculiar velocity for S639 and the galaxies in the cluster being on average
0.2 dex younger than similar galaxies in Coma and HydraI. The results for S639 cau-
tion that some large peculiar velocities may be spurious and caused by unusual stellar
populations.
Key words: galaxies: elliptical and lenticular, cD -- galaxies: distances -- galaxies:
fundamental parameters -- galaxies: stellar content -- dust, extinction
1
INTRODUCTION
The relation known as the Fundamental Plane (FP) may
be used for determination of relative distances to E and
S0 galaxies (e.g., Dressler et al. 1987; Jørgensen, Franx &
Kjaergaard 1996, hereafter JFK96; Baggley 1996; Hudson et
al. 1997). The FP relates the effective radius, re, the mean
surface brightness within this radius, <I>e and the (cen-
tral) velocity dispersion σ, in a relation, which is linear in
logarithmic space (Djorgovski & Davis 1987; Dressler et al.
1987).
The FP has a low scatter (15 − 20% in re) and is there-
fore a valuable tool for studies of peculiar velocities of galax-
ies and clusters (e.g., Baggley 1996; Hudson et al. 1997). The
use of the FP for determination of distances and peculiar ve-
locities relies on the assumption that the FP is universally
valid. Several authors have investigated possible differences
in the FP related to the cluster environment (e.g., Burstein,
Faber & Dressler 1990; Lucey et al. 1991; de Carvalho &
Djorgovski 1992; JFK96; Baggley 1996). Only de Carvalho &
Djorgovski find that the environment has significant effects
on the FP. These authors find field galaxies to be brighter
⋆ E-mail: [email protected], [email protected]
† Hubble Fellow.
than cluster galaxies of similar effective radii and velocity
dispersions. de Carvalho & Djorgovski also find field galax-
ies to be bluer and have weaker Mg2 line indices than cluster
galaxies with similar velocity dispersions. This is in general
agreement with studies that show that E and S0 galaxies
in the outer parts of clusters have weaker Mg2 and <Fe>
indices than those in the central parts of clusters (Guzm´an
et al. 1992; JFK96; Jørgensen 1997).
In this paper we study the poor cluster of galaxies S639,
previously studied by JFK96. The cluster identification is
from Abell, Corwin & Olowin (1989). The cluster has a ra-
dial velocity in the Cosmic Microwave Background (CMB)
frame of czCMB = 6545km s−1 and is located ≈ 28◦ from
the direction to the large mass-concentration known as the
"Great Attractor" (Faber & Burstein 1988). The velocity
dispersion of the cluster is 456+83
−74km s−1 (JFK96). Its rich-
ness is 14 measured as the number of galaxies with magni-
tudes between m3 and m3 + 2 (Abell et al. 1989). m3 is the
magnitude of the third ranked galaxy. S639 has a smaller ve-
locity dispersion and is poorer than clusters like the Coma
cluster and the HydraI cluster. Coma and HydraI have ve-
locity dispersions of 1010+51
−39km s−1, re-
spectively (Zabludoff, Huchra & Geller 1990). The richnesses
given by Abell et al. (1989) is 106 for Coma and 39 for
HydraI.
−44km s−1 and 608+58
c(cid:13) 0000 RAS
2
I. Jørgensen, H. Jønch-Sørensen
Using the FP for 10 E and S0 galaxies in S639 JFK96
found a large peculiar velocity of the cluster, vpec = (1295 ±
359)km s−1 relative to the CMB frame. Further, JFK96
found that the galaxies in the cluster follow a Mg2-σ relation
offset from the relation established for their full sample of 11
clusters. The galaxies in S639 had on average weaker Mg2
indices, see also Jørgensen (1997). JFK96 tried to correct
the derived peculiar velocity of the cluster for the offset in
the Mg2 indices by including a Mg2 term in the FP. The re-
sult was vpec = (879 ± 392)km s−1. However, the coefficient
for the Mg2 term is not well determined, cf. JFK96.
S639 is located at
low galactic latitude,
(l,b) =
(280◦,11◦). Thus, the galactic extinction is large and un-
certainties in the adopted value may severely affect the pre-
cision of the derived distance and peculiar velocity for the
cluster.
The main issue discussed in this paper is whether the
large peculiar velocity of S639 found by JFK96 is real or
the result was caused either by incorrect correction for the
(large) galactic extinction, by selection effects, or by unusual
stellar populations. In order to reach conclusions about these
issues we have obtained additional photometry of galaxies in
S639, giving a sample of 21 E and S0 galaxies with available
photometric and spectroscopic parameters. We have also ob-
tained Stromgren uvby-β photometry for stars in the direc-
tion of the cluster. This photometry is used to determine
the galactic extinction affecting the cluster.
The sample selection for the E and S0 galaxies and the
available data are briefly described in Sect. 2. The determi-
nation of the galactic extinction is covered in Sect. 3. In Sect.
4 the FP is discussed and used for determining the distance
to the cluster. The importance of the stellar populations is
investigated in Sect. 5. The conclusions are summarized in
Sect. 6.
The relations between the parameters for the galaxies
established in this paper are determined by minimization
of the sum of the absolute residuals perpendicular to the
relations. This fitting technique has the advantage that it
is rather insensitive to a few outliers, and that it treats the
coordinates in a symmetric way. The uncertainties of the co-
efficients are derived by a bootstrap method. See also JFK96
for a discussion of this fitting technique.
2 SAMPLE SELECTION AND DATA
The sample of E and S0 galaxies in the poor cluster S639
used in the analysis by JFK96 was selected based on sky
survey images. This sample was by no means complete. We
have obtained additional photometry in Gunn r of galaxies
within the central 25′ × 35′ of the cluster. We aimed at con-
structing a magnitude limited sample of E and S0 galaxies
from the photometry from Jørgensen, Franx & Kjaergaard
(1995a, hereafter JFK95a), our new photometry in Gunn r
and the spectroscopic data from Jørgensen, Franx & Kjaer-
gaard (1995b, hereafter JFK95b) and Jørgensen (1997). We
classified the galaxies based on our CCD images. For galax-
ies outside the central area the classifications from JFK95a
were adopted. In order to further check our classifications
we subtracted models of the best fitting elliptical isophotes
from the images of the galaxies. None of the galaxies in the
sample show residual spiral arms after the model subtrac-
tion. Thus, we find it unlikely that our sample is conta-
minated by early-type spirals. The accidential inclusion of a
few early-type spirals in the fainter part of the sample, for
which the residual spiral arms may be difficult to detect, will
not affect our results significantly, since none of the results
depends critically on the faintest third of the sample.
The final sample contains 21 E and S0 galaxies for which
both photometric and spectroscopic parameters are avail-
able. Within the central 25′ × 35′ the sample is 90% com-
plete to a total magnitude of 15.m5 in Gunn r (magnitudes
corrected for galactic extinction, k-corrected and corrected
for cosmological dimming). Further, we have data for two
fainter galaxies, and two galaxies outside the central area.
2.1 Available data for the galaxies in S639
The new photometry of S639 was obtained with the Danish
1.5-m telescope, LaSilla, March 22 -- 28, 1997. We used the
DFOSC (Danish Faint Object Spectrograph and Camera)
equipped with a 2k×2k thinned Loral CCD. The data were
reduced with standard methods. Two-dimensional surface
photometry of the galaxies and the effective parameters were
derived in the same way as done by JFK95a. The magnitudes
of the new data were standard calibrated by means of zero
point offsets relative to the photometry from JFK95a. The
comparison of the aperture magnitudes with the data from
JFK95a has an rms scatter of 0.m01. More details of the data
reduction are given in Appendix A.
JFK95a derived the (g − r) colors for 10 of the E and
S0 galaxies in S639. The colors were measured within the
effective radii in Gunn r. We have transformed the (g − r)
colors to (B−r) using the relation (B−r) = 0.88(g−r)+0.73.
We established the relation from color data for a total of 41
E and S0 galaxies in the clusters Abell 194, DC2345-28, and
HydraI (JFK95a; Milvang-Jensen & Jørgensen 1998). The
rms scatter of the relation is 0.025.
We use spectroscopic data from JFK95b and Jørgensen
(1997). Velocity dispersions are available for 21 E and S0
galaxies. The absorption line indices Mg2 and <Fe> have
been measured for 20 of those galaxies, and four galaxies
have measured HβG. The Mg2 and <Fe> line indices are
on the Lick/IDS system (named after the Lick Image Dis-
sector Scanner; Faber et al. 1985; Worthey et al. 1994).
The HβG index is related to the Lick/IDS Hβ index as
HβG = 0.866Hβ + 0.485 (Jørgensen 1997). All the spectro-
scopic parameters are centrally measured values corrected to
a circular aperture with a diameter of 1.19 h−1 kpc (JFK95b;
Jørgensen 1997), H0 = 100 h km s−1 Mpc−1. The line indices
are corrected for the effect of the velocity dispersion.
The photometric and spectroscopic parameters for the
galaxies in S639 are given in the Appendix A.
2.2 Reference samples in other nearby clusters
We use the Coma cluster and the HydraI cluster as refer-
ence clusters. The photometric data for the Coma cluster
are from JFK95a. The spectroscopic data are from the lit-
erature (Davies et al. 1987; Dressler 1987; Lucey et al. 1991;
Guzm´an et al. 1992) and have been calibrated to a consistent
system by JFK95b. We also use new spectroscopic data for
additional galaxies in the Coma cluster (Jørgensen 1998).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The data for the HydraI cluster are from JFK95a, JFK95b
and Milvang-Jensen & Jørgensen (1998).
The sample of E and S0 galaxies in Coma covers the
central 64′ × 70′ of the cluster and is 93% complete to a
magnitude limit of 15.m05 in Gunn r. This sample contains
116 galaxies. The sample of E and S0 galaxies in HydraI
covers the central 46′ × 76′ of the cluster and is 80% com-
plete to a limit of 14.m5 in Gunn r. This sample contains
45 galaxies. Both samples include a few galaxies below their
magnitude limit. The line indices <Fe> and HβG are avail-
able for sub-samples of the samples in Coma and HydraI.
There are three reasons to use these large samples of
galaxies in Coma and HydraI as the reference samples,
rather than making use of all the clusters studied by JFK96.
First, our S639 sample is magnitude limited and we there-
fore expect more accurate results by comparing this sample
to other magnitude limited sampes, rather than using the
incomplete samples from JFK96. Second, the JFK96 sam-
ple contains two clusters with smaller velocity dispersions
than S639 and comparable offsets in the Mg2 index relative
to the Coma cluster. Thus, including these clusters in the
reference sample would make that sample inhomogeneous.
Third, our goal is to discuss the reality of the large peculiar
velocity of S639 found by JFK96. JFK96 derived peculiar
velocities under the assumption that Coma is at rest rela-
tive to the CMB frame. In this paper we make the same
assumption in order to facilitate the comparison with the
results from JFK96.
3 THE GALACTIC EXTINCTION IN THE
DIRECTION OF S639
The galactic extinction affecting observations of extragalac-
tic objects is traditionally determined from the HI-mapping
done by Burstein & Heiles (1982). From this mapping we
obtain a galactic extinction in the direction of S639 of
AB = 0.80. However, in some areas especially near the
galactic plane the extinction has been found to be differ-
ent from estimates based on the Burstein & Heiles maps
(e.g., Burstein et al. 1987; Jønch-Sørensen 1994b). In the
following we therefore use two other methods to estimate
the galactic extinction in the direction of S639: Stromgren
photometry of stars in the direction of the cluster, and the
color-Mg2 relation for galaxies in the cluster.
3.1 The Stromgren photometry
S639 was observed as part of a program to obtain Stromgren
uvby-β photometry of stars in the directions of several
galaxy clusters at low galactic latitude. The observations
were obtained with the Danish 1.5-m telescope, La Silla,
March 13 -- 29, 1996. The telescope was equipped with the
DFOSC. The reduction and calibration of the photometry
are described in detail in Appendix B. Here we concentrate
on the determination of the galactic extinction in the direc-
tion of S639.
Intrinsic colors can be estimated in the uvby-β system
for individual stars of spectral types ranging from B- to G-
type. This has been used extensively to map the reddening
in the solar neighborhood (Hilditch, Hill & Barnes 1983;
Franco 1989). Recently deep CCD-photometry surveys have
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The poor cluster of galaxies S639
3
Figure 1. The galactic reddening E(b − y) in the direction of
S639. The 22 stars with 2.58 ≤ β ≤ 2.72 and computable E(b − y)
and MV are shown. The solid curve is a model of the reddening
caused by the diffuse interstellar gas. The dashed curve marks the
maximum observable reddening at a limiting magnitude of V =
17m and MV = 4.m1, corresponding to the limiting magnitude
and mean MV of the observed sample of stars.
made it possible to trace reddening to large distances, hence
all the way through the galactic dust disk using individ-
ual stars (Jønch-Sørensen 1994b; Jønch-Sørensen & Knude
1994). Nissen (1994) reviewed the calibrations and the ac-
curacy of the derived parameters. He showed that the rms
scatter of the residuals in the calibrations sets the limit of the
precision with which E(b − y) can be derived to 0.009. This
corresponds to an uncertainty of 0.012 in E(B−V ) and 0.050
in AB. We use E(B − V ) = 1.35E(b − y), AV = 4.2E(b − y),
and AB = 5.57E(b − y) to convert between reddening and
galactic extinction.
In Fig. 1 distance versus E(b − y) is shown for the
stars in the direction of S639. The typical uncertainty on
E(b − y) is ≈ 0.04. The relative distance uncertainty is
<∼30% for V <∼ 16m and increasing rapidly for fainter stars,
see Appendix B. The mean reddening E(b − y) for stars
at distances larger than 750 pc is 0.106 with an rms scat-
ter of 0.031. The solid curve in Fig. 1 represents the pre-
dicted E(b − y)-distance relation for a model from Jønch-
Sørensen (1994b) with a vertical scale height of the (dif-
fuse) absorbing material of 140 pc and a local normalization
of 0.42 atoms cm−3. For other parameters of the model see
Jønch-Sørensen (1994b). The model indicates a total redden-
ing of E(b − y) = 0.14, equivalent to AB = 0.78, for objects
situated outside the disk. From this we find the visual ab-
sorption acquired through half the diffuse disk in this direc-
tion of the Galaxy to be AV sin b = 0.11 (b in this expression
is the galactic latitude). This value can be compared with
the results for the seven fields surveyed by Jønch-Sørensen
(1994b) and Jønch-Sørensen & Knude (1994). The result for
the S639 direction is comparable with AV sin b = 0.13 found
for the low latitude field (l,b)=(262◦,+4◦), but lower than
the 0.21 found for (l,b)=(270◦,+35◦).
The reddening caused by the diffuse interstellar medium
may be understood as contributions from both diffuse in-
terstellar clouds and an intercloud medium (e.g., Knude
1979b). The typical reddening due to a diffuse cloud is
E(b − y) ≈ 0.03 and the average number of clouds in-
4
I. Jørgensen, H. Jønch-Sørensen
Figure 2. The (B − r)-Mg2 relation. Large boxes -- S639; small
triangles -- Coma and HydraI. Typical error bars are given on
the figure. The solid line is the relation derived from Coma and
HydraI, (B − r) = (1.36 ± 0.22)Mg2 + 0.77. The rms scatter in
(B − r) for the galaxies in S639 is of 0.04. The galactic extinction
for S639 was assumed to be AB = 0.72. This gives consistency
with the (B − r)-Mg2 relation for the Coma and HydraI galaxies.
tercepted is 4.3 per kpc (Knude 1979a, 1981). The inter-
cloud medium contributes with only E(b − y) ≈ 0.001 per
100 pc (Knude 1979b). In total this yields a reddening of
E(b −y) = 0.14 kpc−1 along a line of sight in the disk. In the
direction of S639 a distance of 1 kpc corresponds to a height
above the plane of 190 pc, or about 1.5 times the scale height
of the diffuse medium. Thus, based on the results from local
data we would expect the reddening towards S639 due to the
diffuse interstellar medium to be close to E(b − y) = 0.14, as
indicated by Fig. 1. Two distant stars in the observed field
show low reddening. This may be a result of the 'stochas-
tic' nature of the distribution of the clouds, or simply due
to the rather large measurement errors on E(b − y) for the
individual stars. The two stars have reddenings of approxi-
mately 0.03 and 0.07, and may represent lines of sight that
intercept one and two clouds, respectively.
The sampling and accuracy of the present results do not
allow for a specific mapping of clouds or tracing of reddening
variations across the field. The best estimate of the typical
reddening affecting the observations of the galaxies in S639
is E(b − y) = 0.14, corresponding to E(B − V ) = 0.19,
AV = 0.59 and AB = 0.78.
3.2 The (B − r) - Mg2 relation
E and S0 galaxies in nearby clusters are known to follow a
relation between the color and the Mg2 index (Burstein et al.
1988; Bender, Burstein & Faber 1993). If we assume that the
relation is universal, then it may be used as an independent
way of estimating the galactic extinction. The assumption
means that we expect it to be a universal property of these
galaxies that the global stellar populations match the central
stellar populations (cf., Bender et al. 1993), since the col-
ors are measured within large apertures (the effective radii)
while the Mg2 indices are measured within much smaller
apertures. The color-Mg2 relation was used by Burstein et
al. (1987) to estimate the galactic extinction affecting galax-
ies at some directions close to the galactic plane. The rela-
tion has also been used recently by Schlegel, Finkbeiner &
Davis (1998) to calibrate the DIRBE/IRAS dust maps to
galactic reddening (DIRBE: Diffuse Infra Red Background
Experiment on board the COBE satellite; IRAS: Infrared
Astronomy Satellite). Burstein et al. (1988) found the scat-
ter in the color-Mg2 relation to be compatible with the mea-
surement errors. Therefore, we do not expect the stellar pop-
ulation differences discussed in Sect. 5 to have significant
effect on the galactic extinction derived from the color-Mg2
relation.
We assume that E and S0 galaxies in S639 follow the
same (B −r)-Mg2 relation as E and S0 galaxies in Coma and
HydraI. The (B − r)-Mg2 relation for Coma and HydraI is
shown in Fig. 2. The data for the E and S0 galaxies in S639
are overplotted. With a galactic extinction of AB = 0.72
for S639 the data for the cluster are consistent with the
Coma and HydraI (B − r)-Mg2 relation. The rms scatter
the (B − r)-Mg2 relation for S639 is 0.04. This gives an
uncertainty on AB of ±0.03, equivalent to an uncertainty on
the galactic extinction in Gunn r, Ar, of ±0.02.
In the following we use the average of AB derived from
the Stromgren photometry and the AB determination based
on the (B − r)-Mg2 relation. We adopt AB = 0.75 ± 0.03
(E(B − V ) = 0.18 ± 0.01) for all the galaxies in S639,
where the uncertainty is half the difference between our
two determinations. The corresponding galactic extinction
in Gunn r is Ar = 0.47 ± 0.02. We note that our determina-
tion of E(B − V ) for S639 agrees with the dust maps from
Schlegel et al. (1998). These maps imply a mean value of
E(B − V ) = 0.152 ± 0.024 within the central 25′ × 35′ of the
cluster.
4 THE FUNDAMENTAL PLANE FOR S639
In Fig. 3a we show the FP for S639 edge-on. The effective
radii are in arcsec and the mean surface brightnesses are
given as log <I>e = −0.4(<µ>e − 26.4). <I>e has units of
L⊙ pc−2. The solid line on the figure is the FP for nearby
clusters found by JFK96, log re = (1.24±0.07) log σ−(0.82±
0.02) log <I>e + γ. The zero point is defined by the present
sample of galaxies in S639. The rms scatter for S639 is 0.122
in log re.
If we fit the FP to the S639 data we find
log re =
1.32 log σ − 0.78 log <I>e + 0.007
±0.30
± 0.09
(1)
with an rms scatter 0.136 in log re. Thus, the FP for S639 is
not significantly different from the FP found for nearby clus-
ters (JFK96), or the FP fitted to the Coma and HydraI sam-
ples (Jørgensen 1998; Milvang-Jensen, Jørgensen & Jønch-
Sørensen 1998). In the following we use the FP established
by JFK96.
Fig. 3b shows the FP edge-on with the effective radii
given in kpc (H0 = 50 km s−1 Mpc−1). The data for S639
have been overplotted upon the data for the clusters Coma
and HydraI. We have assumed that the peculiar velocity of
S639 is zero. The solid line on this figure is the FP with the
zero point defined from Coma. The offset between Coma
and S639 is 0.100 ± 0.027 in log re. This may be interpreted
as due to a non-zero peculiar velocity of S639. We assume
that the Coma cluster is at rest relative to the CMB frame
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The poor cluster of galaxies S639
5
Figure 3. The FP for S639 seen edge-on. The photometry is in Gunn r, AB = 0.75 was used. (a) The effective radii in arc seconds.
The solid line is the FP with the zero point derived from the S639 galaxies. (b) The effective radii in kpc, under the assumption of zero
peculiar velocity. Large boxes -- S639; small triangles -- Coma and HydraI. The solid line is the FP with the zero point defined by the E
and S0 galaxies in Coma. Typical error bars are given on the panels.
and has czCMB = 7200km s−1 (JFK96). We then find the
distance of S639, expressed as the expected radial velocity,
to be (5706 ± 350)km s−1. This implies a peculiar velocity
of (839 ± 350)km s−1. If we include a correction for the clus-
ter's offset in the Mg2-σ relation, see Sect. 5, as also tried
by JFK96, we find a distance of (5989 ± 378)km s−1 and
vpec = (556 ± 378)km s−1. The results were not corrected
for Malmquist bias. Based on the rms scatter of the FP we
expect the Malmquist bias from a homogeneous density field
to be ≈ 1.3%, cf. Lynden-Bell et al. (1988).
For both methods the peculiar velocities of S639 found
here are smaller than those found by JFK96. The differences
are partly due to the larger sample of galaxies included in
the present study, and partly due to the slightly lower value
assumed of the galactic extinction. Our result confirms, how-
ever, that if the FP without any Mg2 term is used for dis-
tance determination for S639 then a large positive peculiar
velocity is found.
The rms scatter of the FP for S639 is slightly larger
than found for the Coma and the HydraI clusters (JFK96;
Jørgensen 1998; Milvang-Jensen et al. 1998). Part of this
may be caused by variation of the galactic extinction over
the field of S639. Our Stromgren photometry for the stars in
the field is not accurate enough to investigate this. However,
for other fields within 30 degrees of the galactic plane values
of the rms scatter in E(b − y) of 0.035-0.10 have been found
(Jønch-Sørensen 1994b; Jønch-Sørensen & Knude 1994). If
the field of S639 has similar variations in E(b − y) it would
result in a contribution of ≈ 0.04 to the rms scatter of the
FP (measured in the direction of log re). This explains only
a small part of the scatter in the FP for S639, but it does
bring the unexplained part of the scatter, ≈ 0.10 in log re, in
better agreement with recent results for Coma and HydraI
(Jørgensen 1998; Milvang-Jensen et al. 1998).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Table 1. Model predictions from Vazdekis et al. (1996)
≈ 0.12 log age + 0.19[M/H] + 0.14
Mg2
log <Fe> ≈ 0.12 log age + 0.25[M/H] + 0.34
log HβG ≈ -- 0.27 log age -- 0.135[M/H] + 0.51
log M/Lr ≈ 0.63 log age + 0.26[M/H] -- 0.16
(B − r) ≈ 0.36 log age + 0.36[M/H] + 0.83
Note -- [M/H]≡ log Z/Z⊙ is the total metallicity relative to solar.
Table 2. Systematic offsets for S639 galaxies
Parameter
Offset
∆ log age
∆[M/H]
log M/Lr
(B − r)
Mg2
log <Fe>
log HβG
−0.12 ± 0.03
−0.19 ± 0.05 −0.46 ± 0.12
−0.045 ± 0.010 −0.13 ± 0.03 −0.13 ± 0.03
−0.021 ± 0.006 −0.18 ± 0.05 −0.11 ± 0.03
−0.034 ± 0.019 −0.28 ± 0.06 −0.14 ± 0.08
0.078 ± 0.015 −0.29 ± 0.06 −0.58 ± 0.11
Notes -- offsets calculated as the mean difference between the
parameters for galaxies in S639 and the established relations.
5 THE STELLAR POPULATIONS OF THE
GALAXIES
In this section we investigate whether the offset between the
FP for S639 and the FP for Coma and HydraI may be due
to differences in the stellar populations in the galaxies.
E and S0 galaxies show a number of relations between
the velocity dispersions, the absorption line indices and the
colors. Differences in the stellar populations of galaxies in
different clusters are expected to show up as differences in
these relations. We establish the differences as zero point
offsets for the S639 sample relative to the relations for the
Coma and HydraI samples. In order to interpret the offsets
in terms of differences in the mean age and/or metallicity
of the stellar populations we use the single stellar popula-
tion models from Vazdekis et al. (1996). Table 1 gives the
expected changes in the line indices, the M/L ratio and the
color due to changes in the age and/or metallicity. These re-
lations are valid for models with ages larger than 5 Gyr and
6
I. Jørgensen, H. Jønch-Sørensen
Figure 4. The color (B − r) and the line indices Mg2, <Fe>
and HβG versus the velocity dispersion. Large boxes -- S639;
small triangles -- Coma and HydraI. Typical error bars are
given on the panels. The relation on panel (a) has been es-
tablished from the Coma and HydraI samples. We find (B − r)
= (0.188 ± 0.044) log σ − 0.740 with an rms scatter of 0.037. The
relations on panels (b), (c) and (d) are adopted from Jørgensen
(1997). The galaxies in S639 are systematically bluer, have smaller
Mg2 indices, and larger HβG indices than the galaxies in Coma
and HydraI. A few of the galaxies in S639 also have substantially
smaller <Fe> indices than the bulk of the galaxies in Coma and
HydraI.
a bi-modal initial mass function (IMF) with a high mass
slope identical to a Salpeter (1955) IMF (µ=1.35).
The relations between the velocity dispersions and the
indices Mg2, <Fe> and HβG as well as between the veloc-
ity dispersion and the color (B − r) are shown in Fig. 4.
The (B − r)-σ relation on Fig. 4a was established from the
Coma and HydraI galaxies, while the relations between the
line indices and the velocity dispersion are adopted from
Jørgensen (1997). The galaxies in Coma and HydraI follow
these relations.
The galaxies in S639 are offset relative to the (B − r)-σ
relation, the Mg2-σ relation and the HβG-σ relation. The
galaxies are on average bluer, have smaller Mg2 indices and
larger HβG indices than the galaxies in Coma and HydraI.
Table 2 summarizes the offsets for all the relations. The table
also gives the offset for the FP in terms of an offset in the
M/L ratio, assuming that the peculiar velocity of S639 is
zero.
We use the stellar population models to estimate the
change in age or metallicity that would create the measured
offsets in the various parameters, see Table 2. We assume
that the offsets are either due to age variations, only, or
metallicity variations only. The offsets in Mg2, HβG, (B − r)
and the M/L ratio are all consistent with a difference be-
tween the mean ages of the galaxies in S639 and the galaxies
in Coma and HydraI of -- 0.2 dex. Though the <Fe> indices
are consistent with this it is obvious from Fig. 4c that the
offset in log <Fe> is caused by four galaxies with very small
<Fe> indices. For the rest of the galaxies there is no sig-
nificant offset in <Fe>, while these galaxies show the same
offsets in Mg2 and log M/Lr as the full sample. A possi-
ble reason may be that the galaxies in S639 have a lower
abundance ratio [Mg/Fe] than the bulk of E and S0 galax-
ies in Coma and HydraI. It requires better measurements of
<Fe> to further investigate this potential problem with our
interpretation.
If we interpret the offsets as due to a systematic offset
in the metallicity, while there is no difference in the mean
ages, we find a mean difference of -- 0.12 dex, based on the
offsets in (B − r) and Mg2. However, a difference in the
metallicity cannot fully explain the offset in the M/L ratio.
In this case S639 may have a non-zero peculiar velocity of
≈ 490km s−1. We also note that the offset in HβG cannot be
fully explained by a metallicity difference.
6 CONCLUSIONS
The galactic extinction in the direction of the poor cluster of
galaxies S639 has been determined from Stromgren uvby-β
photometry for stars in the direction of the cluster. Further,
we have tested the consistency of the derived galactic extinc-
tion by using the (B − r)-Mg2 relation for E and S0 galaxies
in the cluster. Our best estimate of the galactic extinction
in the direction of the cluster is AB = 0.75 ± 0.03.
The FP for S639 has been established based on a sample
of 21 E and S0 galaxies. The coefficients for the FP for this
cluster are not significantly different from those of the FP for
the Coma and the HydraI clusters, and are also in agreement
with previous results for other nearby clusters (e.g., JFK96).
Under the assumption that the FP (coefficients and zero
point) is universal we find a distance, in velocity units, to
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
S639 of (5706 ± 350)km s−1. This implies a peculiar velocity
for the cluster of (839 ± 350)km s−1.
The E and S0 galaxies in S639 have significantly smaller
Mg2 indices, larger HβG indices and are bluer than E and S0
galaxies of similar velocity dispersions in Coma and HydraI.
The offset in the FP for S639 relative to Coma and HydraI
may be due to a difference in the stellar populations, rather
than a large peculiar velocity for S639. The data are con-
sistent with a zero peculiar velocity and mean ages of the
S639 galaxies 0.2 dex younger than the mean ages of sim-
ilar galaxies in Coma and HydraI. Alternatively, the Mg2
indices and the (B − r) colors are consistent with a metal-
licity difference of 0.1 dex, with galaxies in S639 having a
lower metallicity than those in Coma and HydraI. In this
case the peculiar velocity of S639 is ≈ 490km s−1. However,
this interpretation is not consistent with the strong HβG
indices measured for the four galaxies, for which we have
measurements of this index.
We conclude that the peculiar velocity of S639 is most
likely overestimated if the FP is used as a distance determi-
nator for this cluster. Even though many studies have shown
that the FP (and the Dn-σ relation which is a projection of
the FP) to a large degree is universally valid (e.g., Burstein,
Faber & Dressler 1990; JFK96; Baggley 1996), our results for
S639 caution that there may be exceptions (see also Gregg
1992). When distance determinations are attempted the best
approach will be to obtain colors and line indices together
with the other required data. This will give the possibility
of identifying clusters (and galaxies), which deviate strongly
from the mean relations between the various global param-
eters. These clusters can also be expected to deviate from
the FP otherwise valid for the bulk of the E and S0 galaxies.
One may attempt to include a Mg2 term in the FP in order
to correct for the effects caused by differences in the stel-
lar populations. However, the coefficient for such a term is
not well determined, cf. JFK96, and the peculiar velocities
derived with this method may not be accurate enough for
investigations of large-scale flows.
Acknowledgements: Lars Freyhammer is thanked for
obtaining part of the observations used for this research.
The Danish Board for Astronomical Research and the Eu-
ropean Southern Observatory are acknowledged for assign-
ing observing time for this project and for financial support.
Support for this work was provided by NASA through grant
number HF-01073.01.94A to IJ from the Space Telescope
Science Institute, which is operated by the Association of
Universities for Research in Astronomy, Inc., under NASA
contract NAS5-26555. HJS acknowledges financial support
from the Carlsberg Foundation, Denmark.
REFERENCES
Abell G. O., Corwin H. G., Jr., Olowin, R. P., 1989, ApJS, 70, 1
Baggley G., 1996, PhD thesis, Oxford University
Bender R., Burstein D., Faber S. M., 1993, ApJ, 411, 153
Burstein D., Heiles C., 1982, AJ, 87, 1165
Burstein D., Davies R. L., Dressler A., Faber S. M., Stone R. P.
S., Lynden-Bell D., Terlevich R. J., Wegner G., 1987, ApJS,
64, 601
Burstein D., Davies R. L., Dressler A., Faber S. M., Lynden-Bell
D., Terlevich R. J., Wegner G., 1988, in Kron R. G., Renzini
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The poor cluster of galaxies S639
7
A., eds., Towards Understanding Galaxies at Large Redshifts,
Kluwer Academic Publishers, Dordrecht, p. 17
Burstein D., Faber S. M., Dressler A., 1990, ApJ, 354, 18
Crawford D. L., 1975, AJ, 80, 955
Davies R. L., Burstein D., Dressler A., Faber S. M., Lynden-Bell
D., Terlevich R. J., Wegner G., 1987, ApJS, 64, 581
de Carvalho R. R., Djorgovski S., 1992, ApJ, 389, L49
Djorgovski S., Davis M., 1987, ApJ, 313, 59
Dressler A., 1987, ApJ, 317, 1
Dressler A., Lynden-Bell D., Burstein D., Davies R. L., Faber S.
M., Terlevich R. J., Wegner G., 1987, ApJ, 313, 42
Faber S. M., Burstein D., 1988, in Rubin V. C., Coyne G. V.,
eds., Large-Scale Motions in the Universe. A Vatican Study
Week, Princeton University Press, Princeton, p. 115
Faber S. M., Friel E. D., Burstein D., Gaskell C. M., 1985, ApJS,
57, 711
Franco G. A. P., 1989, A&A, 227, 499
Franx M., Illingworth G., Heckman T., 1989, AJ, 98, 538
Gregg M. D., 1992, ApJ, 384, 43
Guzm´an R., Lucey J. R., Carter D., Terlevich R. J., 1992, MN-
RAS, 257, 187
Hilditch R. W., Hill G., Barnes J., 1983, MNRAS, 204, 41
Hudson M. J., Lucey J. R., Smith R. J., Steel J., 1997, MNRAS,
291, 488
Jønch-Sørensen H., 1993, A&AS, 102, 637
Jønch-Sørensen H., 1994a, A&AS, 108, 403
Jønch-Sørensen H., 1994b, A&A, 292, 92
Jønch-Sørensen H., Knude J., 1994, A&A, 288, 139
Jørgensen I., 1997, MNRAS, 288, 161
Jørgensen I., 1998, in preparation
Jørgensen I., Franx M., Kjaergaard P., 1992, A&AS, 95, 489
Jørgensen I., Franx M., Kjaergaard P., 1995a, MNRAS, 273, 1097
(JFK95a)
Jørgensen I., Franx M., Kjaergaard P., 1995b, MNRAS, 276, 1341
(JFK95b)
Jørgensen I., Franx M., Kjaergaard P., 1996, MNRAS, 280, 167
(JFK96)
Knude J., 1979a, A&A, 71, 344
Knude J., 1979b, A&A, 77, 198
Knude J., 1981, A&A, 97, 380
Lucey J. R., Guzm´an R., Carter D., Terlevich, R. J., 1991, MN-
RAS, 253, 584
Lynden-Bell D., Faber S. M., Burstein D., Davies R. L., Dressler
A., Terlevich R. J., Wegner G., 1988, ApJ, 326, 19
Milvang-Jensen B., Jørgensen I., 1998, in preparation
Milvang-Jensen B., Jørgensen I., Jønch-Sørensen H., 1998,
in
preparation
Nissen P. E., 1994, Rev. Mex. Astron. Astrofis., 29, 129
Olsen E. H., 1988, A&A, 189, 173
Salpeter E. E., 1955, ApJ, 121, 161
Schlegel D. J., Finkbeiner D. P., Davis M., 1998, ApJ, in press
(astro-ph/9710327)
Vazdekis A., Casuso E., Peletier R. F. Beckman, J. E., 1996,
ApJS, 106, 307
Worthey G., Faber S. M., Gonz´ales J. J., Burstein D., 1994, ApJS,
94, 687
Zabludoff A., Huchra J. P., Geller M. J., 1990, ApJS, 74, 1
APPENDIX A: PHOTOMETRY FOR THE
GALAXIES
Observations were obtained for five CCD fields, covering the
central 25′ × 35′ of S639. The exposure time for each field
was two times 5 minutes in Gunn r. The CCD images were
reduced in a standard way. This includes correction for bias,
and flat field correction with dome flat fields with an addi-
8
I. Jørgensen, H. Jønch-Sørensen
tional correction for the low frequency variation derived from
flat fields obtained at twi-light. The pixel-to-pixel accuracy
of the flat field correction is 0.6%, while the accuracy of the
correction for the low frequency variation is better than 1%.
The two images of each field were registered and added
before determination of the photometric parameters of the
galaxies. The determination of the photometric parameters
were done in the same way as in JFK95a. Two-dimensional
surface photometry for the galaxies were derived with the
program package GALPHOT (Franx, Illingworth & Heck-
man 1989; Jørgensen, Franx & Kjaergaard 1992). This pack-
age fits a full 2-dimensional model to the image of a galaxy.
The results are radial profiles of the surface brightness,
the ellipticity, the position angle and the deviations of the
isophotes from ellipses, as well as the growth curves of the
galaxies.
The magnitudes were standard calibrated by means of
offsets relative to the photometry presented in JFK95a. Fig-
ure A1 shows the comparison of aperture magnitudes for the
two sets of data, after the offset has been applied. The rms
scatter of the comparison is 0.m01.
Effective radii and surface brightnesses were derived by
fitting the growth curves of the galaxies with growth curves
for r1/4 profiles. The seeing was taken into account. We refer
to JFK95a for further details.
We have compared the effective parameters derived
from the new data to those given in JFK95a. There are 11
galaxies in common. For differences derived as "JFK95a" --
"this paper" we find mean offsets of ∆ log re = −0.006 ±
0.019 with an rms scatter of 0.063; ∆<µ>e = −0.02 ± 0.06
with an rms scatter of 0.22; ∆mT = 0.01 ± 0.03 with an rms
scatter of 0.10. The combination log re − 0.35<µ>e, which
enters the FP has also been compared. We find the differ-
ence to be 0.001±0.003 with an rms scatter of 0.011. For the
galaxies in common we use the average of the photometric
parameters given in JFK95a and the new parameters derived
here.
Table A1 lists the average photometric parameters for
all the observed galaxies. The table also gives the spectro-
scopic parameters used in the analysis. These data are from
JFK95b and Jørgensen (1997). Further, the (B − r) colors
derived from the measured (g −r) colors (JFK95a) are listed.
APPENDIX B: STR OMGREN PHOTOMETRY
FOR THE STARS
Stromgren photometry of S639 was obtained with the Dan-
ish 1.5m telescope, La Silla, using the DFOSC. We used a
2k×2k thinned Loral chip (W11-4). Due to problems with
the stability of the flat fields the chip could not be used
in the UV-flooded mode. The stability problem of the flat
fields was avoided by using the unflooded mode, but the re-
sulting response in the u-band was <∼10%. Furthermore, in
unflooded mode the chip had a very low quantum efficiency
(QE) near the edge of the frame. As a consequence only
the central 1450×1450 pixels were used for the Stromgren
photometry. This area corresponds to 9.′5 × 9.′5 on the sky.
The exposure times were 2 × 45 minutes i u, 20 minutes
in v, 10 minutes in b and y, 12 minutes in βwide and 40
minutes in βnarrow. Further, sequences of observations with
Figure A1. Comparison of aperture magnitudes in Gunn r for
galaxies in S639. Filled boxes -- aperture radius 6′′; open boxes --
aperture radius 10′′. The differences are calculated as "JFK95a"-
"this paper". The photometry presented in this paper have been
offset to the mean zero point defined by the JFK95a photometry.
The rms scatter of the comparison is 0.m01.
shorter exposure times were obtained in order to tie-in the
bright and faint stars in the field.
The CCD images were reduced in a standard way, cor-
recting for bias and flat field corrected. Twi-light sky flat
fields were used for all six filters. The accuracy of the flat
field correction is ≈ 1% in all filters except in u where the
effect of the low edge-QE is noticeable as an increase in the
noise.
B1 Transformation to the standard system
During the first seven nights of the observing period simul-
taneous observations were performed using the Danish 0.5m
telescope (the Stromgren Automatic Telescope, SAT), La
Silla. The SAT was used for photoelectric observations of the
uvby-β primary standard stars, some of the secondary stan-
dard stars used for the CCD-photometry, and some bright
stars in the CCD program fields (though none in S639). The
SAT also supplied nightly extinction coefficients. For the rest
of the period where no SAT observations were available we
used the mean extinction coefficients for the first period. One
night was used for CCD observations of E-region standard
stars (Jønch-Sørensen 1993). A total of 17 standard stars
were observed. Transformations from instrumental magni-
tudes to V, (b−y), m1 ≡ (v−b)−(b−y), c1 ≡ (u−v)−(v−b)
and β were derived. The rms scatter of the transforma-
tions were 0.011, 0.010, 0.011, 0.057 and 0.013, respectively.
The transformations will be described in more detail
in
Milvang-Jensen et al. (1998). The transformations are valid
for stars from A-type to early G-type. These stars have
2.58 <∼ β <∼ 2.90, corresponding to 0.43 >∼ (b − y) >∼ 0.0
and luminosity classes V-III.
B2 The limiting magnitude and the accuracy of
the magnitudes
In the observed field 110 objects have complete uvby-β data.
The final sample used for the E(b − y) analysis was limited
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The poor cluster of galaxies S639
9
Table A1. Photometric and spectroscopic parameters for S639 galaxies
Galaxy
log re <µ>e
(B − r)
log σ Mg2
log <Fe> log HβG
E264G23
E264G24
E264G26a
E264G28a
E264G300
E264G301
E264G302
E264G31
J06a
J09
J10
J13
J14
J15
J16
J18
J19a
J20
J23
J26
J31
J32
J101
J104
J109
1.24
0.97
1.18
1.46
1.18
0.66
0.60
1.28
0.99
0.59
0.85
0.84
0.89
0.49
0.61
0.39
0.96
0.52
0.38
0.49
0.81
0.72
0.73
1.26
1.15
20.63
18.92
20.74
21.44
20.01
18.91
18.83
20.35
20.41
20.43
21.50
19.08
19.76
18.73
19.60
19.91
19.64
20.38
18.64
18.31
20.98
19.25
20.76
21.49
21.00
1.14
1.06
...
...
1.15
1.11
1.17
1.11
1.12
...
...
1.13
1.09
1.09
1.11
...
1.10
...
1.13
...
...
...
...
...
...
2.342
2.349
2.115
2.191
2.449
2.265
2.276
2.392
2.041
1.852
1.856
2.298
2.135
2.105
2.094
1.935
...
1.633
2.140
2.496
1.715
2.187
1.960
2.018
2.154
0.291
0.273
0.243
0.242
0.315
0.280
0.262
0.283
0.201
0.210
0.192
0.263
0.244
0.207
0.254
0.223
...
0.182
0.249
0.303
...
0.267
0.210
0.137
0.252
0.429
0.502
0.456
0.490
0.462
0.497
0.424
0.509
0.403
0.260
0.463
0.471
0.442
0.422
0.455
0.294
...
0.206
0.402
0.484
...
0.510
0.421
0.207
0.501
...
0.392
...
...
0.304
0.400
...
...
...
...
...
0.352
...
...
...
...
...
...
...
...
...
...
...
...
...
Notes -- a spiral galaxy, not included in the analysis. re is in arcsec. <µ>e and (B − r)
have been corrected for the galactic extinction (AB=0.75; Ar=0.47), k-corrected, and
corrected for the cosmological dimming. The total correction for <µ>e is 0.m586 which
has been subtracted from the standard calibrated magnitudes. The total magnitudes
can be calculated as mT = <µ>e −5 log re −2.5 log 2π. The typical internal uncertain-
ties on log re and <µ>e are 0.01 and 0.04, respectively. The (B − r) colors are derived
from (g − r) given by JFK95a using the relation (B − r) = 0.88(g − r) + 0.73. The
spectroscopic parameters are from JFK95b and Jørgensen (1997), see these references
for uncertainties on the parameters.
to 22 stars meeting the restrictions implied by the applied
calibrations (see Sect. B3). The low sensitivity of the CCD at
short wavelengths limits the sample of stars with complete
uvby-β information to stars brighter than V = 17m. The
limiting magnitude in the other filters is V ≈ 20m. For stars
brighter than V = 17m the internal uncertainties are σV ≤
0.m008, σb−y ≤ 0.012, σm1 ≤ 0.018, σc1 ≤ 0.065 and σβ ≤ 0.008.
The poor UV-response is manifested in the large uncertainty
of the c1 index.
B3 Determination of E(b − y) and MV
Intrinsic colors can be estimated in the uvby-β system for
individual stars of spectral types ranging from B- to G-type.
The β index (≡ βnarrow −βwide) is not affected by reddening
and is an effective temperature indicator for the stars in
question. For definitions and properties of the uvby-β system
see Crawford (1975).
In this program we concentrate on F- and early G-
type stars since these stars are frequent and the calibra-
tion of uvby-β indices in terms of both intrinsic color and
astrophysical parameters such as MV, Teff and [Fe/H] are
well-established. We follow closely the procedure outlined in
Jønch-Sørensen (1994ab).
We use the intrinsic color calibration of (b−y)0 by Olsen
(1988). (b − y)0 is calibrated using β as indicator of the
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
effective temperature with corrections of the luminosity and
the metallicity via terms involving c1 and m1, respectively
(see Olsen 1988 for details). The rather complex relation
between the three indices means that the uncertainty on the
derived E(b − y) depends upon position in the (β, m1, c1)
space. There is no tight relation between V and and the
uncertainty on E(b − y) as one might have expected from
the relation between the magnitude uncertainties and V .
However, there is an effect of the spectral type in the sense
that the uncertainty on E(b − y) increases near the cool end
of the sample. For (b − y)0 ≤ 0.39 the average uncertainty is
<σE(b−y)> = 0.022 while for (b−y)0 ≥ 0.4 the uncertainty is
<σE(b−y)> = 0.070. The average uncertainty for our sample
is <σE(b−y)> = 0.039.
The distances are estimated from MV, which is derived
using the calibration by Crawford (1975). Again β is the
main parameter with corrections of the luminosity from c1.
The uncertainty of the derived MV is closely correlated with
V , increasing from <σMV> = 0.m17 for V ≤ 14m to 0.m93 for
V ≥ 16m. This includes contributions from the uncertainties
on E(b − y) and V . The uncertainties on MV correspond
to (formal) relative distance uncertainties between ≈10%
and ≈45%. Note, that the rms scatter of the calibration of
photometric distances is ≈15%, cf., Nissen (1994).
|
astro-ph/0201425 | 1 | 0201 | 2002-01-25T19:37:09 | Evidence for a developing gap in a 10 Myr old protoplanetary disk | [
"astro-ph"
] | We have developed a self-consistent model of the disk around the nearby 10 Myr old star TW Hya which matches the observed spectral energy distribution and 7mm images of the disk. The model requires a significant dust size evolution and a partially-evacuated inner disk region, as predicted by theories of planet formation. The outer disk, which extends to at least 140 AU in radius, is very optically thick at IR wavelengths and quite massive ~0.06 Msun for the relatively advanced age of this T Tauri star. This implies long viscous and dust evolution timescales, although dust must have grown to sizes of order ~1cm to explain the sub-mm and mm spectral slopes. In contrast, the negligible near-infrared excess emission of this system requires that the disk be optically thin inside ~4 AU.This inner region cannot be completely evacuated; we need ~0.5 lunar mass of ~1 micron particles remaining to produce the observed 10 micron silicate emission. Our model requires a distinct transition in disk properties at ~4 AU, separating the inner and outer disk. The inner edge of the optically-thick outer disk must be heated almost frontally by the star to account for the excess flux at mid-IR wavelengths. We speculate that this truncation of the outer disk may be the signpost of a developing gap due to the effects of a growing protoplanet; the gap is still presumably evolving because material still resides in it, as indicated by the silicate emission, the molecular hydrogen emission, and by the continued accretion onto the central star (albeit at a much lower rate than typical of younger T Tauri stars). TW Hya thus may become the Rosetta stone for our understanding of the evolution and dissipation of protoplanetary disks. | astro-ph | astro-ph | Evidence for a developing gap in a 10 Myr old protoplanetary disk
Nuria Calvet 1, Paola D'Alessio 2,3, Lee Hartmann1, David Wilner1, Andrew Walsh4, and
Michael Sitko5
ABSTRACT
2
0
0
2
n
a
J
5
2
1
v
5
2
4
1
0
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
We have developed a physically self-consistent model of the disk around the
nearby 10 Myr old star TW Hya which matches the observed spectral energy
distribution and 7mm images of the disk. The model requires both significant
dust size evolution and a partially-evacuated inner disk region, as predicted
by theories of planet formation. The outer disk, which extends to at least 140
AU in radius, is very optically thick at infrared wavelengths and quite massive
(∼ 0.06M⊙) for the relatively advanced age of this T Tauri star. This implies
long viscous and dust evolution timescales, although dust must have grown
to sizes of order ∼ 1 cm to explain the sub-mm and mm spectral slopes. In
contrast, the negligible near-infrared excess emission of this system requires that
the disk be optically thin inside ∼< 4 AU. This inner region cannot be completely
evacuated; we need ∼ 0.5 lunar mass of ∼ 1 µm particles remaining to produce
the observed 10µm silicate emission. Our model requires a distinct transition in
disk properties at ∼ 4 AU, separating the inner and outer disk. The inner edge
of the optically-thick outer disk must be heated almost frontally by the star to
account for the excess flux at mid-infrared wavelengths. We speculate that this
truncation of the outer disk may be the signpost of a developing gap due to the
effects of a growing protoplanet; the gap is still presumably evolving because
material still resides in it, as indicated by the silicate emission, the molecular
hydrogen emission, and by the continued accretion onto the central star (albeit
at a much lower rate than typical of younger T Tauri stars). TW Hya thus may
become the Rosetta stone for our understanding of the evolution and dissipation
of protoplanetary disks.
1Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138, USA. Electronic
[email protected], [email protected],
mail: [email protected], [email protected],
[email protected], [email protected]
2Instituto de Astronomia, UNAM, Ap.P. 70-264, 04510 M´exico D.F., M´exico
3American Museum of National History, Central Park West at 79th Street, New York, NY10024-5192
4Max Planck Institut fur Radioastronomie, auf dem Hugel 69, Bonn, 53121, Germany
5Department of Physics, University of Cincinnati, Cincinnati, OH 45221-0011
– 2 –
Subject headings: Accretion, accretion disks, Stars: Circumstellar Matter, Stars:
Formation, Stars: Pre-Main Sequence
1.
Introduction
The discovery of extrasolar planets (Marcy & Butler 1998 and references therein) has
opened up a new era in the study of planetary systems. While many important clues to the
processes of planet formation can be obtained from studies of older systems, the best tests
of formation scenarios will require the direct detection of actively planet-forming systems.
It is thought that the formation of giant planets involves the sweeping up of material
in a wide annulus in the circumstellar disk, resulting in the development of a gap (Lin
& Papaloizou 1986, 1993; Bryden et al. 1999). Material inside the planet-driven gap can
continue to accrete onto the central star; if the planet can prevent material from accreting
across the gap into the inner disk, the eventual result would be the evacuation of the region
interior to the planet. In the case of the Solar System, the formation of Jupiter might have
prevented outer disk gas from reaching the inner solar system; the inner gas disk accreted
into the Sun, while solid planetesimals remaining behind eventually formed the terrestrial
planets.
The above scenario suggests that the signature of a forming giant planet would be
the presence of a gap that is not entirely evacuated. In this case dusty emission from the
inner disk might still be observable. In addition, giant planet formation requires the prior
consolidation of large solid bodies to serve as cores for subsequent gas accretion; while
such bodies would be invisible with current techniques, one would expect to see evidence
for substantial growth in dust particles. Finally, if the inner disk has not been completely
evacuated by the forming planet, one might expect to observe continued accretion onto the
central star, as all T Tauri systems with inner disks as detected from near-infrared disk
emission are also accreting (Hartigan et al. 1990).
In this article we propose that the relatively young low-mass star TW Hya has a
developing gap in its inner disk qualitatively similar to that expected from planet formation.
The evidence supporting this this proposal is: (1) reduced (optically-thin) emission from
the inner disk; (2) mm-wave spectra which seem to require grain growth; (3) extra emission
from the edge of the outer disk; and (4) continued accretion onto the central star, albeit at
a rate substantially lower than that observed from most T Tauri stars. TW Hya has an age
of 10 Myr and so according to current theories it is quite likely to be close to the epoch
– 3 –
of planet formation. It is also part of an association of young stars of similar age, but it
stands out in that it is the only system still accreting at a substantial rate (Muzerolle et al.
2000). The (outer) disk of TW Hya is quite massive, so that there is likely to be more than
enough material available to form giant planet(s). TW Hya is also the closest known such
system, and so it will be a prime target for following studies to confirm our model.
2. Observations
The observations used to constrain our disk model are taken from the literature. In
addition, we have obtained a narrow-band L (3.55-3.63 microns) measurement of TW Hya
on June 14th, 2000 with the Australian National University 2.3-m telescope at Siding
Spring, using the near infrared camera CASPIR (Cryogenic Array Spectrometer Imager;
McGregor et al. 1994). CASPIR contains a 256x256 InSb array. TW Hya was imaged at L
with a pixel scale of 0.25 arcseconds and an on-source integration time of 236 seconds. The
standard star BS4638 (magnitude 4.50 at L) was observed at a similar airmass to TW Hya
and was used to calibrate the magnitude of TW Hya. The L magnitude we obtained was
7.12, equivalent to 0.39Jy.
This measurement is important in constraining the amount of near-infrared excess in
the system. Using K=7.37 (Webb et al. 1999), K-L = 0.25. The K-L color for a K7V star
is 0.11 (Kenyon & Hartmann 1995), which would yield an infrared excess of 0.14. However,
such small infrared excess does not necessarily imply emission from disk material (Wolk
& Walter 1996). For example, the non-accreting stars in Taurus (spectral types ∼ K7 to
M2) have K-L between -0.05 and 0.25, while 95% of the accreting stars in Taurus have
K-L between 0.35 and 1.2 (Meyer, Calvet, & Hillenbrand 1997). This small value of K-L is
consistent with the simultaneous flux measurements of Sitko et al. (2000) indicating little if
any hot dust emission at wavelengths < 5µm, though it is difficult to rule out completely
any excess.
3. Model assumptions
We assume that the heating of the disk is due to stellar irradiation and viscous
dissipation, and calculate self-consistently the disk heights and temperatures following the
methods we have applied to interpret young ∼ 1 Myr old disks (D'Alessio et al. 1998 =
Paper I, 1999 = Paper II, 2000 = Paper III.) We adopt M∗ = 0.6 M⊙, R∗ = 1 R⊙ and
T∗ = 4000 K for the stellar mass, radius, and effective temperature (Webb et al. 1999),
– 4 –
M = 5 × 10−10 M⊙ yr−1 as derived by Muzerolle et al. (2000),
and a mass accretion rate of
which we take as constant through the whole disk. The inclination of the disk axis to the
line of sight is i ∼ 0◦, in agreement with the nearly symmetric HST images (Krist et al.
2000), and we adopt the Hipparcos distance of 55 pc (Wichmann et al. 1998).
With the assumption of steady accretion, the disk surface density scales as
Σ ∝ M α−1T −1, with T is the midplane temperature (see Paper II). The temperature
does not vary much between acceptable models, partly because it is mostly determined
by irradiation heating, not viscous energy dissipation. The disk mass therefore is roughly
M α−1. Moreover, for long wavelengths where much of the disk is
constant for constant
optically thin, the emergent fluxes also tend to scale in the same way. For the purposes of
fitting, we vary the parameter α as the dust opacities are varied, but note that varying α
is equivalent to varying the disk mass, because M is fixed. To first order, however, choices
with M α−1 ≈ constant are acceptable.
We use a dust mixture consisting of silicates, refractory organics, troilite and water ice,
following Pollack et al. (1994 = P94). The dust size distribution is taken to be n(a) ∼ a−p,
with p = 3.5, between given amin and amax. Optical properties for the compounds are taken
from Jager et al. (1994), P94, Begemann et al. (1994; also see Henning et al. 1999), and
Warren (1984). We consider the grains to be compact segregated spheres, and calculate the
opacity using a Mie scattering code (Wiscombe 1979). Sublimation temperatures for the
different grain types are taken from P94.
4. Disk model
The spectral energy distribution (SED) of TW Hya is shown in Figure 1. References
for the observational data are given in the figure caption. Note that our narrow-band L
observation overlaps with Sitko et al. (2000) fluxes.
In Figure 1, we compare the SED with the median SED for ∼ 1 Myr-old T Tauri stars
in the Taurus molecular cloud (Paper II), normalized to the TW Hya stellar photosphere
at H (1.6µm), thus compensating for the differing distances and stellar luminosities. Even
though the fluxes of TW Hya are relatively high compared with the median Taurus SED
at λ ∼> 20µm, there is a large flux deficit in TW Hya below 10 microns (Jayawardhana et
al. 1999); in particular, fluxes are essentially photospheric below ∼ 6µm (Sitko et al. 2000).
The flux deficit below 10 µm has lead to inferences of disk clearing inside a few AU from
the central star (Jayawardhana et al. 1999), but since the disk is still accreting mass onto
the star (Muzerolle et al. 2000), it has to extend all the way into the corotation radius at
– 5 –
least, so the inner disk radius has to be ∼< 0.03 AU (inferred from the 2 day photometric
period [Mekkaden 1998]).
We show below that models with a uniform dust well-mixed with the gas throughout
the disk extending all the way onto the inner radius cannot explain these features of the
SED. The observations are much better understood if the disk is divided into two regions:
the outer disk, which is more nearly comparable to the structure inferred for typical T Tauri
disks (Paper III), and the inner disk, which is much more optically thin than in typical T
Tauri disk models.
4.1. Outer disk
As discussed in Paper II, ISM dust mixtures (with small amax) cannot explain the
far-IR and mm-wave fluxes of T Tauri stars. The similarity between the median Taurus
SED and that of TW Hya at wavelengths λ ∼> 100 µm suggests disk models of the type
explored in Paper III, in which we allow for growth to large particles, can in principle
explain the observations.
Figure 2 shows results for flared, irradiated disk models calculated with the methods
of Paper III. The model SEDs shown are for amax = 1 mm (which fits the median SED
in Taurus; Paper III), 1 cm, and 10 cm, calculated with abundances in the dust mixture
usually assumed for protoplanetary disks (P94), which yield a dust-to-gas ratio of 0.013.
It can be seen that models where grains have grown to amax ∼ 1 − 10 cm provide a much
better fit to the long-wavelength SED than the amax = 1 mm model, requiring no or very
little additional emission from a wind or non-thermal sources. Models with amax ≪ 1 mm
fail to reproduce either the sub-mm and mm spectral slopes or the total flux levels for
reasonable disk masses.
Assuming an outer disk radius of ∼ 140 AU, comparable to the radius at which Krist
et al. (2000) find a rapid decline in disk density, the disk mass is ∼ 0.03, 0.06, and 0.11M⊙
for the amax = 1 mm, 1 cm, and 10 cm models, or α = 5 × 10−4, 3 × 10−4, and 1 × 10−3,
respectively. The amax = 1 cm model has a Toomre parameter Q ∼ 1 at its outer edge and
so is near the limit expected for gravitational stability. The high mass values are due to the
opacity at 7 mm, which decreases as a−1/2
max (for p = 3.5, see Paper III), so higher masses
are needed to account for the flux for larger grain mixtures. For example, for the amax =
10 cm mixture, the opacity at 7 mm is a factor of 3 lower than the frequently assumed law
κBS = 0.1(λ/250 µm)−1 (Beckwith & Sargent 1991) and the slope is slightly flatter, ∝ λ0.8
(cf. Paper III, Figure 2).
– 6 –
Our total disk mass estimates, using the dust-to-gas ratio of 0.013, are much higher
than the gas mass obtained by Kastner et al. (1997) from 12CO emission, 3.5 × 10−5M⊙.
This discrepancy may be attributed to a large degree of molecular depletion or to the
fact that optical depth effects may not have been properly included (Beckwith & Sargent
1993); it could also be due to molecules existing in the gas phase only in the hot upper
atmospheric layers of the disk where only a small amount of mass resides (Willacy &
Langer 2000). An alternative to high disk masses, not considered in this work, is to have
larger opacities at 7 mm than we obtain; porous aggregates, specially for amorphous carbon
particles, may result in such larger opacities (Stognienko, Henning & Ossenkopf 1995). Our
results for grain growth are not unique. However, within the assumptions we have made
concerning dust opacities, we cannot reproduce the mm-wave fluxes and spectral slope
without including dust particles much larger than those of a "standard ISM" mixture (see
Paper III). Beckwith & Sargent (1991) noted that optical depth effects could make the
mm-wave spectral slope flatter, and thus reduce or eliminate the need for grain growth.
However, we are unable to make TW Hya disk sufficiently optically thick over its large
radial extent.
4.2. Edge of the outer disk
The well-mixed grain-growth models of Figure 2, with self-consistently calculated
temperature structures, exhibit too little emission in the 20 − 60µm wavelength region in
comparison with observations. In principle, disk models with more small particles could
have higher fluxes. However, it is apparent that these models also predict far too much
flux at wavelengths ∼< 10µm, because they are too optically thick in the inner regions; this
suggest that some clearing has occurred in the inner disk. Moreover, it suggest that the
outer disk should have an optically-thick edge in which extra heating could be important.
In the case of irradiation of an optically thick disk with a smoothly-varying thickness,
the stellar flux captured by the disk (and thus the disk heating per unit area) depends
on the cosine of the angle between the direction of incidence of the stellar beam and the
normal to the disk surface µ0 (Kenyon & Hartmann 1987). In general, stellar radiation
penetrates the disk very obliquely, so µ0 is fairly small, ∼ 10−3 − 10−2 (cf. D'Alessio 1996).
However, if the disk had an inner edge, this portion of the disk would be illuminated by the
star more directly, increasing µ0 dramatically and thus increasing the amount of irradiation
heating. We propose that the outer regions of the TW Hya disk can be described as in §4.1,
but that it is truncated at a few AU by a steep, optically thick region, where most of the
mid-IR flux excess arises (Figure 3). Inside this region lies the inner optically thin disk,
– 7 –
which produces negligible continuum flux.
To calculate the emission of the disk edge, we have assumed that its irradiation surface,
that is the surface where most of the stellar energy is deposited, has a shape given by
zs = zo(Ro)exp[(R − Ro)/∆R], where zo is the height of irradiation surface of the outer disk
at radius Ro and ∆R is a characteristic width of the edge. The edge continuum emission
is produced in the photosphere, which for simplicity we take as the irradiation surface;
above this layer, there is a hotter optically thin region, the atmosphere, which we take as
isothermal (cf. Chiang & Goldreich 1997). We think that a more detailed treatment of the
structure of this region is not necessary given the exploratory nature of this study. The
temperature of the edge photosphere, assuming that half of the intercepted flux reaches the
photosphere, is given by
Tphot(R) ≈ T∗(cid:18)R∗
R (cid:19)1/2(cid:18)µ0
2 (cid:19)1/4
.
where µ0 = cos(θo) is obtained from θo = π/2 − tan−1(dzo/dR). The flux is evaluated as
Fν = Z Ro
Rth
Iν2πRdR,
(1)
(2)
where Rth is a the radius where the disk becomes optically thin, and Iν = Bν(Tphot).
The location of the outer disk edge and its width, characterized by Ro and ∼ ∆R, are
fairly well restricted by the mid-infrared excess, although the actual shape is not so well
constrained as long as most of it faces the star. In general we find ∆R ∼ 0.5 AU, because
the shape of the excess is fairly narrow and thus cannot be produced by a region with a
large range of temperatures. In addition, if Ro is much smaller than ∼4 AU, then the edge
is too hot and there is too much excess below 10 µm. Similarly, if Ro is much larger, there
is too little flux. With Ro ∼ 4 AU and ∆R ∼ 0.5 AU, µo varies from ∼ 0.2 at R = 3 AU
to ∼ 0.7 at R = 4 AU, and Tphot varies from ∼ 80 to 100 in this range. The resulting
continuum flux is shown in detail in the lower panel of Figure 4, while the upper panel
shows the fit to the SED of the composite disk model.
The temperature of the optically thin atmosphere of the edge is given by
W (R)κ∗
P T 4
∗ = κP (Tup)T 4
up
(3)
where W (R) is the geometrical dilution factor, W (R) = Ω∗/4π ≈ (R∗/2R)2, for R >> R∗,
and κ∗
P and κP are Planck mean opacities calculated at the stellar and local radiation fields,
respectively (cf. Paper I, Paper II, Paper II). The contribution to the flux from this region
is given by eq. (2), with Iν = Bν(Tup)τν and τν ≈ κντ∗/χ∗, where τ∗ and χ∗ are the optical
depth and extinction coefficient at the characteristic wavelength of the stellar radiation. We
– 8 –
follow Natta et al. (2000) taking τ∗ ≈ µ0, but in this work we include the effect of scattering
at the wavelengths at which the stellar radiation is absorbed. The neglect of scattering in
the calculation of χ∗ leads to artificially large values of τν and thus of Iν, and would be
only appropriate if scattering was completely forward. However, the asymmetry parameter
g = hcos Θi, with Θ the scattering angle, is in general < 1, and this approximation is not
valid. We have included the effect of scattering using χ∗ = κ∗ + (1 − g∗)σ∗, where κ∗, σ∗, are
the absorption and scattering coefficients and g∗ the asymmetry parameter for the assumed
dust mixture, all evaluated at the characteristic wavelength of the stellar radiation.
We assume that the atmosphere of the edge has small grains so it can produce emission
in the silicate feature. We considered different glassy and crystalline pyroxenes and olivines
(with optical properties from Laor & Draine 1993, Jager et al. 1994, and Dorschner et al.
1995). We find that glassy pyroxene Mg0.5F e0.43Ca0.03Al0.04SiO3 has one of the highest
ratios κν/χ∗ as required for producing a strong band (see also Natta et al. 2000). However,
even with the increasing heating at the disk edge, the conspicuous silicate emission feature
seen in the SED of TW Hya cannot be explained by emission from the atmospheric layers of
the edge. This is due in part to the inclusion of scattering of stellar light in the calculation
of χ∗, as discussed above, and in part to the adopted dust mixture; organics dominate the
opacity at the stellar wavelength (Paper III, Figure 2), yielding κ10µm/χ∗ ∼ 1. Note that
the atmosphere of the optically thick outer disk, which is irradiated less frontally by the
star and is thus cooler than the atmosphere of the edge, cannot produce significant silicate
emission either.
4.3.
Inner disk
Since the region inside the outer disk edge is not empty because material is still
accreting onto the star (Muzerolle et al. 2000), we have explored the possibility that a small
amount of particles coexists with the accreting gas of the inner disk, giving rise to silicate
feature emission, while on the other hand still resulting in negligible continuum flux.
We assume that the temperature of the dust in this region is given by equation (3). The
emergent intensity is calculated as Iν = Bνκντ10/κ(10µm), where τ10, the optical depth at
10µm, is a parameter. This region extends to ∼ 0.02 AU, the radius where grains sublimate.
We find that we can fit the silicate feature with τ10 ∼ 0.05, see Figure 4, lower panel; this
optical depth is small enough for the disk continuum emission in the near infrared to be
negligible. The best fit to the profile is achieved with glassy pyroxene, in agreement with
the fact that the profile is closer to that in young stars than in comets (Sitko et al. 2000).
The grain sizes are in the range amin ∼ 0.9µm and amax ∼ 2µm. Smaller grains have much
– 9 –
higher temperatures and produce too much emission at short wavelengths, resulting in a
narrower silicate profile than observed. Bigger grains produce too little emission.
From τ10 ∼ 0.05, we get a column density for the ∼ 1µm dust of Σd ∼ 4 × 10−3g cm−2,
which implies a mass inside 4 AU of ∼ 4 × 1025 g ∼ 0.5 lunar masses. A strong lower
limit to the mass of gas in the inner disk can be obtained from the observed rate of gas
accretion onto the star assuming that the material in the inner disk is in free-fall towards
the star, Mgas >> MR3/2/(2GM∗)1/2. Inside 4 AU, we obtain Mgas >> 6 × 10−10M⊙.
Another lower limit can be obtained from Σd, assuming a normal dust-to-gas ratio, yielding
Σg ∼ 0.4 g cm−2 and a mass Mgas > 2 × 10−6M⊙ ∼ 0.6 earth masses. If solid material is
hidden into large bodies, then the mass in gas could be much higher than this value. From
this limit, nonetheless, we can get an upper limit for the radial velocity of the gas at 1
AU, vR < M /2πRΣg ∼ 0.008 km s−1 << vK(1AU) ∼ 23 km s−1, so the gas probably drifts
inwards slowly, following nearly Keplerian orbits in the inner disk.
Weintraub, Kastner, and Bary (2000) have detected a flux of ∼ 1.0 × 10−15erg s−1 cm−2
in the 1-0 S(1) line of H2 towards TW Hya. Using Tin´e et al. (1997) models, we can estimate
a flux in this line at 55 pc, F (H2) ∼ 2 × 10−15(ǫ/10−22s−1)(MH2/10−8M⊙)erg s−1 cm−2,
where ǫ is the emissivity per H2 molecule. For a gas temperature of ∼ 1000K and a
ionization rate of 3 × 10−10s−1, estimated with typical parameters and an X-ray luminosity
LX ∼ 1030ergcm−3 (Kastner et al. 1997) using the Glassgold, Najita, & Igea (1997)
corrected expression, ǫ is 2.3 × 10−22erg s−1 for densities nH ≥ 107cm−3 (because the line
becomes thermalized; S. Lepp, personal communication). Assuming a disk height of 0.05
AU, we estimate nH >> 4 × 107cm−3 inside 4 AU, so from the mass limit estimated above,
we see that the observed H2 very likely arises in the inner disk. Firmer conclusions require
a determination of the gas temperature in the inner disk, which is left for future work.
4.4. Model tests: comparison with VLA 7 mm Data
In principle, observations of disk surface brightness distributions can be used to test
our physically self-consistent models based on SED fitting. The optical and near-infrared
scattered light distributions presented by Krist et al. (2000) and Trilling et al. (2001) do
not have sufficient resolution to probe our inner disk region; however, they do provide
constraints on our outer disk structure. We find that the scattered light fluxes predicted
by the amax = 1 cm model presented above matches the observations of Krist et al. (2000)
reasonably well, although Krist et al. find evidence for structure that cannot be reproduced
in detail with disk properties smoothly-varying with radius, as assumed here. This work
will be reported in a forthcoming paper (D'Alessio et al. 2001).
– 10 –
The high angular resolution data at 7 mm from the Very Large Array (Wilner et al.
2000) begin to approach the resolution needed to probe the inner disk, as well as constrain
the thermal dust emission of the outer disk instead of simply the scattering surface, and
thus provide an additional test of our model. Moreover, the imaging data provide an
independent test, since the model was constructed only to match the SED. The value of τ10
sets an upper limit to the opacity at λ > 10µm, so the inner disk should not contribute
significantly to the emission at 7 mm. We have calculated the intensity profile of the disk,
assuming different values for the brightness temperature Tb of the edge between 3 and 4
AU, since we do not know the properties of its interior. An upper limit for Tb is the surface
temperature, ∼ 100K, but it could be lower if it was optically thin at 7 mm, so we have
varied Tb between 60K and 100K.
Figure 5a shows the 7 mm images at two angular resolutions. The "low" resolution
image (∼ 0.′′6, about 35 AU) emphasizes extended low brightness emission, and the "high"
resolution (∼ 0.′′1, about 6 AU) shows the smallest size scales where emission remains
detectable with the available sensitivity. Figure 5b shows the result of imaging the model
brightness distribution at these two angular resolutions using the same visibility sampling as
the observations. For a more quantitative comparison, Figure 5c shows the residual images
obtained by subtracting the model visibilities from the 7 mm data, and then imaging the
difference with the standard algorithms.
The low resolution model image, with the centrally peaked dust emission (Tb ≥ 60K),
compares very favorably with the 7 mm images of Figure 5a; unfortunately, the higher
resolution is still not adequate to resolve directly the presence of the inner hole. But in
both cases, the residual images show only noise. The agreement between the model and
observations confirms the insightful parametric modeling of Wilner et al. (2000), who fitted
the observations with radial power laws for the temperature and surface density. This
was to be expected, since Wilner et al. adopted powers that were consistent with the
predictions of irradiated accretion disk models at large radii. However, since our models are
constructed from first principles, by solving the disk structure equations in the presence of
stellar irradiation subject to the constraint of steady accretion, we are able to predict and
confirm not only the radial dependences of the physical quantities (which are not exactly
power laws), but also their absolute values. The free parameters of our model are essentially
the disk mass (or equivalently, the parameter α, see §3) and the dust mixture. The models
were constructed primarily to fit the SED, but they can also fit the radial brightness
distribution, as the good agreement between the observations and the model indicate. On
the other hand, although there is no guarantee that the outer disk is completely steady, or
has a constant α, or ours is the correct prescription for the dust mixture, these assumptions
seem to provide a good description of the physical situation in the disk of TW Hya.
– 11 –
5. Discussion
The lack of infrared flux excess at wavelengths below 10 microns could be easily
understood if the inner disk was evacuated inside a few AU. However, the fact that material
is still being transfered onto the star, as evidenced by the broad emission line profiles and
the ultraviolet excess emission (Muzerolle et al. 2000), indicates that gas still exists in
the inner disk. However, this material must be optically thin to explain the flux deficit
in the near infrared. If the outer disk extended inwards to the magnetospheric radius,
then the surface density Σ at 2 AU would be ∼ 320g cm−2, implying a surface density
in solids of Σdust ∼ 3g cm−2 (for a dust to gas mass ratio of 0.01). To obtain an optical
depth at the near-infrared of < 0.05 with this amount of dust, the opacity should be of
order κdust < 0.05/Σdust ∼ 0.02cm2g−1. If we assume that the grains are large enough that
the opacity per gram of grain is ∼ Qextπa2/ (4/3)πρga3, with Qext the extinction cross
section and ρg the grain density, both or order 1, then κdust ∼ 1/a. Using this estimate, we
find that the original solids should have grown to sizes a > 50 cm to account for the low
optical depth in the near infrared. While this estimate is very rough, it suggests that very
significant grain growth has probably already taken place in the inner disk of TW Hya,
especially if the solid to gas ratio has increased significantly from solid matter left behind
as gas flows onto the star. This inside-out coagulation and settling is consistent with the
evolution of solid particles predicted in solar nebula theories. Dust grains are expected to
coagulate by sticking collisions at a rate ∝ 1/Ω2 ∝ P 2 and reach a maximum size at the
midplane in a few ×1000P , where Ω is the Keplerian angular velocity and P the orbital
period (Safronov 1972; Goldreich & Ward 1973; Weidenschilling 1997; Nakagawa 1981).
Since P ∼ 1.4 × 103(R/100AU)3/2 yrs, coagulation and settling are expected to occur first
in the inner disk.
In the outer disk, a dust mixture where grains have reached ∼ 1 cm and dust and gas
are well mixed can explain the observations, within the framework of our assumptions. This
interpretation is consistent with the Weidenschilling's (2000) argument that grain growth
to at least 1 cm is necessary before particles start settling to the midplane, because smaller
grains get stirred around by turbulence, and we suggest that turbulence plays an important
role in the evolution of dust in protoplanetary disks. Grains in the upper layers can absorb
stellar radiation and efficiently heat the disk and make it flare, resulting in the large far
infrared and millimeter fluxes observed. Nonetheless, grains in the outer disk have grown
significantly from ∼ 1 mm, the size that characterizes the median SED of Taurus (Paper
III). To get high mm fluxes for such large grains, the disk mass has to be high, ∼ 0.06M⊙,
since the opacity at 7 mm decreases as grains grow (cf. Paper III, §2). This high value for
the mass is surprising; it is similar or higher than values typically found for disks around 1
Myr old T Tauri stars (Beckwith et al. 1990), and comparable to that of objects emerging
– 12 –
from the embedded phase (D'Alessio, Calvet, & Hartmann 1997). On the other hand, as we
discussed in Paper III, disk masses may have been underestimated if the standard power
law opacity is applied indiscriminately to disks where significant grain growth may have
taken place.
It may be possible that the 3.6 cm data point contains a significant contribution from
hot (ionized) stellar plasma. This contribution is unlikely to come from a wind. If the
flux emitted by the wind is proportional to the mass loss rate, then scaling down from
M ∼ 10−6M⊙yr−1 to the
the wind/jet contribution in the case of HL Tau, which has
case of TW Hya, and assuming a ratio of mass loss rate to mass accretion rate of ∼ 0.1
(Calvet 1998), we find that this contribution at 3.6 cm should be ∼ 2 × 10−4 mJy, too
low to be important. A non-thermal contribution due to stellar surface activity is still
possible, given the high X-ray flux of TW Hya (Kastner et al. 1997). Second epoch
observations should resolve this question, because any non-thermal contribution should be
highly time-dependent. In any event, the high flux at 7 mm, at which a non-thermal plasma
contribution is much less likely, indicates that the grain mixture must contain grains that
have grown at least to ∼ 1 cm, otherwise the opacity would already turn to the optical
limit and drop, as is apparent in Figure 2 (cf. Paper III, Figure 2).
The peculiarities of the disk of TW Hya suggest an advanced state of evolution of
the dust. In addition, the presence of a steep transition between the inner and outer disk
suggest the action of an agent other than dust settling to the midplane, which in first
approximation would be expected to produce effects that vary monotonically with radius.
We speculate that we may be seeing the outer edge of a gap opened by the tidal action of a
growing protoplanet. Numerical models of gap formation in disks result in a sharp drop of
the surface density at the gap and enhancements near the edges, as disk material is pulled
away from the protoplanet and into the outer disk (Bryden et al. 2000; Nelson et al. 2000).
The edge of the gap would be facing the star, and if it was optically thick, would result
in excess emission similar to that modeled schematically by our truncation region. The
situation could be comparable to the one considered by Syer & Clarke (1995), who studied
the evolution of the disk spectrum under the presence of a planet formed in the disk, in
the case in which the mass of the protoplanet was larger than the mass of the disk at the
planet location. In this case, the disk cannot push the planet inwards at the local viscous
time-scale and material forms a reservoir upstream of the gap (Lin & Papaloizou 1986).
The surface density enhancement at the edge could make it optically thick, so the midplane
temperature would not increase as much as the surface temperature; the scale height, in
this case, would not increase significantly, justifying our assumption of similar heights for
the edge region and outer disk and the neglect of the effects of shadowing. The details of
the structure of this region, in any event, are of necessity very schematic. The extrapolated
– 13 –
mass of the disk inside ∼ 4 AU is ∼ 10−3M⊙ ∼ 1 Jupiter mass, so this mechanism would
work if a giant planet were forming in this region of the disk.
Finally, we address the question of what TW Hya tells us about protoplanetary disk
evolution. Although we do not have direct measurements of the gas mass of the outer
disk, it would be very surprising if the gas were strongly depleted there, as suggested by
Zuckerman et al. (1995). Both the SED and the scattered light emission as a function
of radius follow the predictions of models in which the dust is well-suspended in the gas,
producing finite scale heights of the dust disk. Based on this inference, we suggest that the
existence of the (fairly massive) TW Hya disk is evidence that neither UV radiation nor
stellar winds can efficiently remove gas at radii ∼> 5 AU over timescales of 10 Myr for stars
formed in dispersed, isolated environments.
The survival of disk material at 10 Myr also places constraints on viscous timescales.
c/ν, where Rc is a characteristic radius and ν the viscosity (cf. Pringle
The surface density of an accretion disk evolves with time, as viscous dissipation makes the
disk transfer mass onto the star and expand to conserve angular momentum in a viscous
time-scale given by R2
1981). Here we use the similarity solutions for disk evolution of Hartmann et al. (1998),
which assume temperature and surface density distributions comparable to our outer disk
models. The viscous time-scale in these solutions is ts ∼ R2
1/ν ∼ 8 × 104(10−2/α)(R1/10AU)
yrs, where R1 is the radius that contains ∼ 60 % of the initial mass (at t = 0), and the
viscosity ν is expressed in terms of the standard Shakura-Sunyaev (1973) α parameterization.
Adopting α ∼ 5 × 10−4 and R1 ∼ 100 AU, then ts ∼ 1.6 × 107 yrs. In the similarity
model, the mass of the disk falls in time as Md ∝ T −1/2, where T = 1 + t/ts. Thus,
if the total disk mass at 10 Myr is 0.06M⊙, then the initial disk mass must have been
0.08M⊙, comparable to that of objects still surrounded by infalling envelopes, like HL
Tau (D'Alessio, Calvet, & Hartmann 1997). The surface density drops below a Σ ∝ R−1
power-law by a factor of 1/e at a distance R1(T ) = R1T ∼ 160 AU, close to the ∼ 140 AU
transition in midplane density found in Krist et al. (2000). With these values of disk mass
and radius, the value of the Toomre parameter Q remains close to unity and so gravitational
instabilities (which could transfer angular momentum rapidly and thus invalidate the purely
viscous evolution model) need not arise.
The mass accretion rate for this constant-α model 6 at an age of 1 Myr is
M = Md(0)/(2tsT 3/2) ∼ 2 × 10−9M⊙yr−1, near the lowest values observed in accreting T
Tauri stars of this age (Gullbring et al. 1998; Hartmann et al. 1998). The mass accretion
rate at 10 Myr is then ∼ 1 × 10−9M⊙yr−1, which is slightly higher than estimated by
6Note that a factor (Md(0)/0.1M⊙) is missing in expression (39) of Hartmann et al. (1998).
– 14 –
Muzerolle et al. (2000). It would not be surprising if accretion is being halted in the inner
disk, and that inner and outer disk are becoming decoupled. If material is prevented from
reaching the inner disk, the material inside the gap would drain onto the central star on
a viscous time at ∼ 4AU, which with α ∼ 10−3, would be ∼ 3 × 105 yrs. Such a short
timescale suggests that the probability of observing a system in this stage is very low, but
still possible.
Instead of the α ∼ 10−3 adopted above, Hartmann et al. (1998) estimated a typical
value of α = 10−2 for T Tauri disks based on mm-wave disk mass estimates and assuming
that these disks expand to ∼ 100 AU in ∼ 1 Myr. If the TW Hya disk (the region containing
most of the mass) is not much larger than 200 AU in radius, then α cannot be this large. A
smaller viscosity could be consistent with the typical T Tauri data if disk masses are a few
times larger than typically estimated, as suggested by D'Alessio et al. (2000).
One remaining question is why TW Hya's disk evolution has been so much slower than
that of other members of the association with presumably similar ages. In part, this must
be due to the presence of binary (or multiple) stellar companions in several other systems
(e.g., HD 98800; Soderblom et al. 1998) which can disrupt disks. In addition, we conjecture
that TW Hya may simply have had a larger initial disk radius (equivalent to a large R1
in terms of the similarity solution above), and thus a lower average surface density, than
other systems. A larger radius would lead to slower viscous evolution for a given α. Also,
in contrast to timescales for dust coagulation and sedimentation, which depend mostly
on orbital period (see above), theoretical models of the runaway growth of giant planets
indicate that the timescale for such growth is a very sensitive function of surface density
(Pollack et at. 1996). The lower surface density of the disk around TW Hya would imply a
slower growth of giant planets. This would imply, in turn, that the final clearing of disks is
due to sweeping-up of material by large bodies.
6. Conclusions
The disk of TW Hya seems to be in advanced state of dust evolution. A planet or other
large perturbing body may have already formed, opening a gap with outer edge around ∼ 4
AU, and although there is still material in the inner disk, it is rapidly dissipating. In the
outer disk, grains may be reaching the size necessary to start settling towards the midplane.
This interpretation can consistently explain the large degree of activity of TW Hya, the
lack of disk emission at near-IR wavelengths and the large fluxes beyond ∼ 10 µm, as well
as the 7 mm images.
– 15 –
Evacuated regions in the disks are also known to result from the effects of companion
stars (GG Tau), and in some cases disk accretion can still occur, especially if the binary
has an eccentric orbit (DQ Tau). Indeed, HD98800, another member of the TW Hya
association, shows evidence for an inner disk hole; because this system is quadruple, it
is quite likely that the inner disk regions have been evacuated by the companion stars.
Although we cannot rule out such a possibility in TW Hya, the fact that accretion still
occurs onto the central star, which is not the case in HD 98800, suggests that the companion
body responsible for the gap is much less massive than a typical companion star, and thus
much less effective in clearing out the disk. Interferometric imaging should be attempted to
constrain the mass of any companion object, expected to have a separation of order 2-3 AU
(0.04 - 0.06 arcsec at 55 pc).
Acknowledgments. We gratefully acknowledge discussions with Edwin Bergin and
Stephen Lepp. This work has been supported by NASA Origins of Solar Systems grants
NAG5-9670 and NAG5-9475. P.D. has been supported by Conacyt grant J27748E and a
fellowship from DGAPA-UNAM, M´exico.
– 16 –
REFERENCES
Allard, F. & Hauschildt, P. H. 1995, ApJ, 445, 433
Begemann, B., Dorschner, J., Henning, T., Mutschke, H., & Thamm, E. 1994, ApJ, 423, 71
Beckwith, S.V.W., & Sargent, A.I. 1991 ApJ, 381, 250
Beckwith, S. V. W. & Sargent, A. I. 1993, ApJ, 402, 280
Beckwith, S.V.W., Sargent, A.I., Chini, R. S., & Guesten, R. 1990, AJ, 99, 1024 (BSCG)
Bryden, G., Chen, X., Lin, D. N. C., Nelson, R. P., & Papaloizou, J. C. B. 1999, ApJ, 514,
344
Bryden, G., R´ozyczka, M., Lin, D. N. C., & Bodenheimer, P. 2000, ApJ, 540, 1091
Calvet, N. 1998, Accretion Processes in Astrophysical Systems: Some Like it Hot!, 495
Chiang, E. I. & Goldreich, P. 1997, ApJ, 490, 368
D'Alessio, P. 1996, Ph.D. Thesis, Universidad Nacional Aut´onoma de M´exico, M´exic
D'Alessio, P., Calvet, N., & Hartmann, L. 1997, ApJ, 474, 397
D'Alessio, P., Cant´o, J., Calvet, N., & Lizano, S. 1998, ApJ, 500, 411 (Paper I)
D'Alessio, P., Calvet, N., Hartmann, L., Lizano, S. and Cant´o, J. 1999, ApJ, 527, 893
(Paper II)
D'Alessio, P., Calvet, N., & Hartmann, L. 2000, ApJ (in press) (Paper III)
de La Reza, R., Torres, C. A. O., Quast, G., Castilho, B. V., & Vieira, G. L. 1989, ApJ,
343, L61
Dorschner, J., Begemann, B., Henning, T., Jaeger, C., & Mutschke, H. 1995, A&A, 300, 503
Glassgold, A. E., Najita, J., & Igea, J. 1997, ApJ, 485, 920
Goldreich, P. & Ward, W. R. 1973, ApJ, 183, 1051
Gregorio-Hetem, J. Lepine, J. R. D., Quast, G. R., Torres, C. A. O., & de la Reza, R. 1992,
AJ, 103, 549
Gullbring, E., Hartmann, L., Briceno, C., & Calvet, N., 1998, ApJ, 492, 323
Hartigan, P., Hartmann, L., Kenyon, S. J., Strom, S. E., & Skrutskie, M. F. 1990, ApJ,
354, L25
Hartmann, L., Calvet, N., Gullbring, E., D'Alessio, P., 1998, ApJ, 495, 385
Henning, Th., Il'In, V. B., Krivova, N. A., Michel, B., & Voshchinnikov, N. V. 1999, A&AS,
136, 4
– 17 –
Jayawardhana, R., Hartmann, L., Fazio, G., Fisher, R. S., Telesco, C. M., & Pina, R. K.
1999, ApJ, 521, L129
Jager, C., Mutschke, H., Begemann, B., Dorschner, J., & Henning, T. 1994, A&A, 292, 641
Kastner, J. H., Zuckerman, B., Weintraub, D. A. & Forveille, T. 1997, Science, 277, 67
Kenyon, S.J., & Hartmann, L. 1987, ApJ, 323, 714
Kenyon, S. J. & Hartmann, L. 1995, ApJS, 101, 117
Krist, J. E., Stapelfeldt, K. R., M´enard, F. , Padgett, D. L., & Burrows, C. J. 2000, ApJ,
538, 793
Laor, A. & Draine, B. T. 1993, ApJ, 402, 441
Lin, D. N. C. & Papaloizou, J. 1986, ApJ, 309, 846
Lin, D. N. C. & Papaloizou, J. C. B. 1993, in Protostars and Planets III, eds. E.H. Levy &
J.L. Lunine (Tucson:Univ. Arizona Press), 749
Marcy, G. W. & Butler, R. P. 1998, ARA&A, 36, 57
McGregor P., Hart J., Downing M., Hoadley D., Bloxham G., 1994, Experimental
Astronomy, vol 3, p139
Meyer, M. R., Calvet, N., & Hillenbrand, L. A. 1997, AJ, 114, 288
Mekkaden, M. V. 1998, A&A, 340, 135
Muzerolle, J., Calvet, N., Briceno, C., Hartmann, L., & Hillenbrand, L. 2000, ApJ, 535, L47
Nakagawa, Y., Nakazawa, K., & Hayashi, C. 1981, Icarus, 45, 517
Natta, A., Meyer, M. R., & Beckwith, S. V. W. 2000, ApJ, 534, 838
Nelson, R. P., Papaloizou, J. C. B., Masset, F. ;. ;., & Kley, W. 2000, MNRAS, 318, 18
Pollack, J. B., Hollenbach, D., Beckwith, S., Simonelli, D. P., Roush, T. & Fong, W. 1994,
ApJ, 421, 615 (P94)
Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., & Greenzweig,
Y. 1996, Icarus, 124, 62
Pringle, J. E. 1981, ARA&A, 19, 137
Rucinski, S. M. & Krautter, J. 1983, A&A, 121, 217
Safronov, V.S. 1972, in Evolution of the Protoplanetary Cloud and Formation of the Earth
and Planets (Moscow: Nauke Press)
Shakura, N.I., & Sunyaev, R.A. 1973, A&A, 24, 337
Sitko, M. L., Lynch, D. K., & Russell, R. W. 2000, AJ, 120, 2609
– 18 –
Soderblom, D. R. et al. 1998, ApJ, 498, 385
Stognienko, R., Henning, T., & Ossenkopf, V. 1995, A&A, 296, 797
Syer, D. & Clarke, C. J. 1995, MNRAS, 277, 758
Tin´e, S., Lepp, S., Gredel, R., & Dalgarno, A. 1997, ApJ, 481, 282
Trilling, D. E., Koerner, D. W., Barnes, J. W., Ftaclas, C., & Brown, R. H. 2001, ApJ, 552,
L151
Warren, S.G., 1984, Applied Optics, vol. 23, p. 1206
Webb, R. A., Zuckerman, B., Platais, I., Patience, J., White, R. J., Schwartz, M. J. &
McCarthy, C. 1999, ApJ, 512, L63
Weidenschilling, S.J. 1997, Icarus, 127, 290
Weidenschilling, S.J. 2000, Space Science Reviews, 92, 295
Weintraub, D. A., Sandell, G., & Duncan, W. D. 1989, ApJ, 340, L69
Weintraub, D. A., Kastner, J. H., & Bary, J. S. 2000, ApJ, 541, 767
Wichmann, R., Bastian, U., Krautter, J., Jankovics, I. & Rucinski, S. M. 1998, MNRAS,
301, L39
Willacy, K. & Langer, W. D. 2000, ApJ, 544, 903
Wilner, D. J., Ho, P. T. P., Kastner, J. H., & Rodr´iguez, L. F. 2000, ApJ, 534, L101
Wilner, D. J. 2001, to appear in "Young Stars Near Earth: Progress and Prospects", eds.
R. Jayawardhana and T. Greene, ASP Conference Series.
Wiscombe, W.J., 1979, Mie scattering calculations: advances in technique and fast, vector-
speed computer codes, NCAR/TN-140 + STR, National Center for Atmospheric
Research, Boulder, Colorado
Wolk, S. J. & Walter, F. M. 1996, AJ, 111, 2066
Zuckerman, B., Forveille, T., & Kastner, J. H. 1995, Nature, 373, 494
This preprint was prepared with the AAS LATEX macros v4.0.
– 19 –
Fig. 1.- Spectral energy distribution of TW Hya. Observations are from Rucinski &
Krautter (1983; average of UBVRI measurements) (open circles), Webb et al.
(1999)
(triangles), Sitko et al. (2000) (small open circles), this paper (open square), Jayawardhana
et al. (1999) (squares), IRAS (de la Reza et al. 1989; Gregorio-Hatem et al. 1987; crosses),
Weintraub et al. (1989) (hexagons), Wilner (2001) (open circle), and Wilner et al. (2000)
(filled circles) The solid line is the median SED of classical T Tauri stars in Taurus (Paper
II). Note the excess flux at mid-infrared wavelengths and the flux deficit below 10 µm (see
text).
– 20 –
Fig. 2.- SED of TW Hya and disk models with gas and dust well mixed and extending
to the magnetospheric radius, for amax = 1 mm (solid line), 1 cm (dotted line), and 10 cm
(long dashed line). Larger maximum grain sizes produce a better match to the sub-mm and
mm-wave spectrum. The SED of the stellar photosphere is also shown for comparison (short
dashed line.)
– 21 –
Fig. 3.- Model disk adopted for TW Hya. The outer disk, where grains have grown to ∼
1 cm, has a edge at R ∼ 3 − 4 AU, and surrounds the inner optically thin disk, which has
a vertical optical depth at 10µm τ10 ∼ 0.05. Gas still exists in the inner disk accreting onto
the star through a magnetosphere. A minute amount of ∼ 1 µm dust permeates this gas.
The dotted line is the surface of the outer disk if it extended inward; the resulting SED for
this model is shown in Figure 4.
– 22 –
Fig. 4.- Above: Fit to the TW Hya SED with our composite disk model. Below: Detail
of the infrared region. The emission from each region is indicated: outer disk (dot-dash),
edge of outer disk (dashed), inner disk (dotted), star (light solid), total (heavy solid). The
stellar spectrum is taken from the Allard & Hauschildt (1995) M dwarf library, where a
model with appropriate effective temperature and gravity (log g =4) could be found. The
model has been scaled to the observations at K (2.2 µm). The excess near ∼ 4.5µm could
be CO fundamental emission; the amount of excess (if any) at λ < 5µm is very uncertain,
as it depends critically on the effective temperature (and model) for the stellar photosphere.
– 23 –
Fig. 5.- (a) VLA 7 mm images of TW Hya at two resolutions, from Wilner et al.
(2000). The synthesized beam sizes are 0.′′82 × 0.′′61 p.a. 20 (left) and 0.′′13 × 0.′′10, p.a. 39
(right). The contour levels are ±(2, 3, 4, 5, 6)× the rms noise levels of 0.5 and 0.4 mJy/beam.
Negative contours are dashed. (b) Simulated VLA 7 mm images of the disk model brightness
distribution described in the text. The contours and beam sizes are the same as in a. The
grey scale shows extended emission, most of which remains undetectable with the available
VLA sensitivity. (c) Difference images obtained by subtracting visibilities derived from the
model from the VLA 7 mm data. Again, the contours and beam sizes are the same as in a.
The residual images show no significant signal.
|
astro-ph/9710011 | 1 | 9710 | 1997-10-01T17:47:16 | Decoupled and inhomogeneous gas flows in S0 galaxies | [
"astro-ph"
] | A recent analysis of the "Einstein" sample of early-type galaxies has revealed that at any fixed optical luminosity Lb S0 galaxies have lower mean X-ray luminosity Lx per unit Lb than ellipticals. Following a previous analytical investigation of this problem (Ciotti & Pellegrini 1996), we have performed 2D numerical simulations of the gas flows inside S0 galaxies in order to ascertain the effectiveness of rotation and/or galaxy flattening in reducing the Lx/Lb ratio. The flow in models without SNIa heating is considerably ordered, and essentially all the gas lost by the stars is cooled and accumulated in the galaxy center. If rotation is present, the cold material settles in a disk on the galactic equatorial plane. Models with a time decreasing SNIa heating host gas flows that can be much more complex. After an initial wind phase, gas flows in energetically strongly bound galaxies tend to reverse to inflows. This occurs in the polar regions, while the disk is still in the outflow phase. In this phase of strong decoupling, cold filaments are created at the interface between inflowing and outflowing gas. Models with more realistic values of the dynamical quantities are preferentially found in the wind phase with respect to their spherical counterparts of equal Lb. The resulting Lx of this class of models is lower than in spherical models with the same Lb and SNIa heating. At variance with cooling flow models, rotation is shown to have only a marginal effect in this reduction, while the flattening is one of the driving parameters for such underluminosity, in accordance with the analytical investigation. | astro-ph | astro-ph | Decoupled and Inhomogeneous Gas Flows in S0 Galaxies
A. D'Ercole and L. Ciotti
Osservatorio Astronomico di Bologna, via Zamboni 33, 40126 Bologna, ITALY
Received
;
accepted
In press on ApJ (Main Journal)
7
9
9
1
t
c
O
1
1
v
1
1
0
0
1
7
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
– 2 –
ABSTRACT
A recent analysis of the Einstein sample of early-type galaxies has revealed
that at any fixed optical luminosity LB S0 galaxies have lower mean X-ray
luminosity LX per unit LB than ellipticals. Following a previous analytical
investigation of this problem (Ciotti & Pellegrini 1996), we have performed 2D
numerical simulations of the gas flows inside S0 galaxies in order to ascertain the
effectiveness of rotation and/or galaxy flattening in reducing the LX/LB ratio.
The flow in models without Supernova (SNIa) heating is considerably
ordered, and essentially all the gas lost by the stars is cooled and accumulated
in the galaxy center. If rotation is present, the cold material settles in a disk
on the galactic equatorial plane. Models with a time decreasing SNIa heating
host gas flows that can be much more complex. After an initial wind phase,
gas flows in energetically strongly bound galaxies tend to reverse to inflows.
This occurs in the polar regions, while the disk is still in the outflow phase.
In this phase of strong decoupling, cold filaments are created at the interface
between inflowing and outflowing gas. Models with more realistic values of the
dynamical quantities are preferentially found in the wind phase with respect
to their spherical counterparts of equal LB. The resulting LX of this class of
models is lower than in spherical models with the same LB and SNIa heating.
At variance with cooling flow models, rotation is shown to have only a marginal
effect in this reduction, while the flattening is one of the driving parameters for
such underluminosity, in accordance with the analytical investigation.
Subject headings: Galaxies: elliptical and lenticular, cD – Galaxies: ISM –
Galaxies: structure – X-rays: galaxies
– 3 –
1.
Introduction
It is well known that normal early-type galaxies are X-ray emitters, with 0.2–4 keV
luminosities ranging from ∼ 1040 to ∼ 1043 erg s−1. The X-ray luminosity LX is found to
correlate with the blue luminosity LB (LX ∝ L2.0±0.2
), but there is a large scatter of roughly
two orders of magnitude in LX at any fixed LB > 3 × 1010L⊙ (Fabbiano 1989; Fabbiano,
Kim, & Trinchieri 1992).
B
The trend and the scatter in the LX − LB diagram have been successfully reproduced
by the time-dependent, 1D hydrodynamical models of Ciotti et al. (1991, hereafter CDPR).
In these models the gas lost by the evolving stars is heated to X-ray temperatures by stellar
random motions and SNIa explosions. Given the assumed time dependence of both mass
loss rate (∝ t−1.36) and SNIa rate (∝ t−1.5), the CDPR models can evolve through up to
three different dynamical phases, i.e., from winds to subsonic outflows to inflows (in the
following we call this kind of evolution the WOI sequence). CDPR introduced the global
parameter χ in order to predict the dynamical status of the gas flows in non-rotating,
spherical galaxies. This parameter was defined as the ratio between the power required to
extract steadily the gas supplied by the mass loss of the evolving stellar population from
the galaxy potential well, Lg, and the power supplied to the gas by the thermalization of
the stellar random motions and by the supernova heating (Lσ and LSN, respectively). Due
to the temporal evolution of the various source terms, χ is an increasing function of time.
Numerical simulations showed the correspondence between χ < 1 and the presence of a
wind phase (i.e., of a low LX), and the onset of a high LX soon after χ becomes > 1.
A recent analysis of the complete Einstein sample of early-type galaxies showed that
on average S0 galaxies have lower LX and LX/LB than ellipticals (Es) (Eskridge, Fabbiano,
& Kim 1995a,b). A possible explanation of this fact could be that S0s and non-spherical
Es are less able to retain hot gaseous haloes than rounder systems of the same LB. This
– 4 –
could be due to a lower gravitational energy, because of their higher rotation rate which
decreases the effective potential, or their different mass distribution. A theoretical analysis
of this problem – based on the global energetics of the X-ray emitting gas – was carried out
by Ciotti & Pellegrini (1996, hereafter CP). They generalized the global parameter χ for
flows in flat and rotating two-component galaxy models. In particular, they added to χ the
contribution of the ordered rotation of the galaxy stellar component, Lrot, showing that, at
least from a global point of view, rotation alone cannot change an energetically bound gas
flow to an unbound one (see §2.3). On the contrary, using two-component Miyamoto-Nagai
models (1975, hereafter MN), and under the assumption of an identical rate of SNIa per
unit LB in S0s and Es, CP showed that flat galaxies can have a much lower χ than rounder
ones of the same luminosity, due essentially to their flattening, even in the presence of
massive spherical dark haloes. More precisely, they showed how a realistic flattening can
produce a decrease of Lg of ∼ 20%, a result which can be sufficient to change an inflow
into a wind. So, the effect of the flattening was suggested as a possible explanation for the
S0 underluminosity. Incidentally, note that the decrease in the gravitational energy that
is released by a stationary inflow at the bottom of the galactic potential well, associated
to the decrease of Lg, does not lower significantly the LX of pure cooling flow models (cfr.
Fig.1 in CDPR).
Despite the possible agreement between the observations and the results of CP, this
purely energetical approach can be seriously questioned considering the complexity of 2D
gas flows and the fact that χ is a global quantity. In fact decoupled flows – in which at the
same time different parts of the galaxy are in inflow and outflow – may in principle develop
in the CDPR scenario in such a way that LX remains high even if χ < 1; if this is the case,
χ < 1 cannot be used as a reliable indicator for a low X-ray luminosity, and the analysis of
CP is seriously questioned.
– 5 –
2D numerical simulations of gas flows in early-type galaxies were actually carried
out by Kley & Mathews (1995, hereafter KM) and Brighenti & Mathews (1996, hereafter
BM). These simulations, however, considered only pure cooling flows for which the flow
decoupling is not expected. KM considered rotating spherical King models plus a spherical
quasi-isothermal dark halo, without SNIa heating. Successively BM considered oblate
galaxies with different axial ratios, sustained by various amounts of ordered and disordered
kinetic energies. They assumed a SNIa rate ∝ t−1, with a present day value of one-tenth of
that proposed by Tammann (1982). Their models show that the gas flow and the associated
LX are only slightly dependent on the flattening of the galaxy (in agreement with the CP
analysis applied to the inflow case), and that LX is reduced by increasing the amount of
ordered rotation. These 2D hydrodynamical models do not solve problems such as the
disposal of excessive amounts of cold gas (a problem already encountered in 1D pure cooling
flow models) on an equatorial disk, which is actually unobserved, and the excessively flat
shape of the X-ray isophotes when compared with the observations.
Motivated by the given arguments, we adopt the same input physics used for the
hydrodynamical equations in CDPR, in order to explore whether the WOI sequence is still
achieved in flat galaxies, how such a sequence is influenced by galaxy flattening and rotation,
how important the decoupling in the flow is, which observational consequences are predicted
in this case, and finally whether the energetical explanation of CP is exhaustive. This is
accomplished with two different classes of models. In the first we compare the behavior of
the gas flow in spherical and flat galaxies that are energetically globally equivalent (both
with and without SNIa heating). In the second class of simulations we investigate the
gas flows in realistic galaxy models of S0s, also in order to explore the possibility of flow
decoupling in galaxies with χ < 1 and its effects on LX.
We thus study the evolution of gas flows in the galaxy models constructed and described
– 6 –
in detail by CP, i.e., axisymmetric stellar and dark mass distributions with the MN shape,
and in which the internal dynamics is determined by the solution of the associated Jeans
equations, allowing for variable amounts of ordered rotation and velocity dispersion.
In Section 2 we briefly describe the models used in the simulations. In Section 3
the results obtained for models with different amounts of ordered rotation, SN heating,
flattening, and characteristic scale-length are presented. A discussion, considering also some
recent observational results is given in Section 4, and in Section 5 the main conclusions are
summarized.
2. The models
2.1. The galaxy density distribution
The two-component mass models used in our simulations and their internal dynamics
are fully discussed in CP, which includes the explicit analytical formulae for the various
terms in the virial theorem and for χ are given. Here only the basic formulae are reported.
The stellar density distribution is given by
ρ∗ = M∗b2
4π ! aR2 + (a + 3ζ)(a + ζ)2
ζ 3[R2 + (a + ζ)2]5/2
,
(1)
(2)
and its potential is
Φ∗ = −
GM∗
,
qR2 + (a + ζ)2
where ζ = √z2 + b2, and (R, ϕ, z) are the cylindrical coordinates (MN)1. The dark matter
halo is described by another MN density distribution characterized by Mh, ah and bh The
1When a = 0 the density distribution (1) reduces to the Plummer (1911) model, and b
becomes its effective radius. So, b can be roughly considered a characteristic scale-length,
while the galaxy flattening is determined by the dimensionless ratio s = a/b.
– 7 –
associated Jeans equations are solved in the total potential Φtot = Φ∗ + Φh, and under
the assumption that the distribution function depends only on the specific energy and the
angular momentum component along the z-axis (see, e.g., Binney & Tremaine 1987, p.197).
With this assumption the radial and axial velocity dispersions are equal, i.e., σR = σz ≡ σ,
and the only non-zero streaming motion can be in the azimuthal direction (ϕ). The intrinsic
degeneracy in the azimuthal velocity component is removed adopting the Satoh (1980)
k-decomposition, i.e.,
and
vϕ
2 = k2(v2
ϕ − σ2),
σ2
ϕ ≡ v2
ϕ − vϕ
2 = k2σ2 + (1 − k2)v2
ϕ,
(3)
(4)
with 0 ≤ k ≤ 1, where vϕ is the modulus of the ordered rotational velocity of the galactic
stellar component. In the case k = 0 no net rotation is present and all the flattening is due
to σ2
ϕ; k = 1 corresponds to the isotropic rotator. The analytical solutions of the Jeans
equations can be obtained for bh = b. In simple case of ah = a the formulae for the velocity
components are expressed in terms of elementary functions and are given in CP together
with their edge-on projection.
2.2. The hydrodynamical equations
The equations describing the gas flows in our simulations are
∂ρ
∂t
+ ∇ · (ρv) = αρ∗,
∂m
∂t
+ ∇ · (m ⊗ v) = −(γ − 1)∇E + gρ + αρ∗v∗
∂E
∂t
+ ∇ · (Ev) = −(γ − 1) E∇ · v − L + αρ∗(ǫ◦ +
1
2kv − v∗k2),
(5)
(6)
(7)
where v is the fluid velocity, and ρ, m, and E are the mass, the momentum, and the
internal energy of the gas per unit volume, respectively. The source terms on the r.h.s. of
– 8 –
equations (5)-(7) describe the injection of mass, momentum and energy in the gas due to
the mass return and energy input from the stars. α(t) = α∗(t) + αSN(t) is the specific mass
return rate from stars and SNIa respectively, with α∗ ∝ t−1.36 and αSN ∝ t−1.5; ǫ◦ is the
injection energy per unit mass due to the stellar random motions and to SNIa explosions.
v∗ is the bulk velocity of the stars, whose modulus is given by equation (3), g = −∇Φtot
is the gravitational acceleration due to the total mass distribution, and γ = 5/3. Finally,
L = nenpΛ(T ) is the cooling rate per unit volume, where for the cooling law Λ(T ) we follow
the prescription of Mathews & Bregman (1978). For more details on the exact form of
α∗(t), αSN(t) and ǫ◦, see CDPR and KM.
To integrate the set of equations we used a second-order, upwind, 2D extension of the
numerical code described in CDPR, coupled with a staggered, spherical, Eulerian grid.
Given the symmetry of the problem, the grid covers polar angles 0◦ ≤ θ ≤ 90◦ with 40
equally-spaced angular zones, and reflecting boundary conditions are applied at θ = 0◦ and
θ = 90◦. The radial coordinate covers the range 0 ≤ r = √R2 + z2 ≤ 100 kpc with 80
logarithmically spaced zones, and a free outflow from the grid is allowed at the outer edge.
At the inner edge reflecting boundary conditions are set. We assume the model galaxies to
be initially devoid of gas due to the previous activity of the Type II SNe. A discussion on
the reliability and implications of this assumption is given in CDPR.
2.3. The χ parameter
The main ingredients of the global parameter χ are presented here. The SNIa heating
evolution is parameterized as LSN = 1051RSNLB erg s−1, with
RSN = 0.88h2 SNU ϑSN t
15Gyr!−1.5
,
(8)
– 9 –
where h = H◦/100 km s−1 Mpc−1, and 1 SNU = 1 SN per century per 1010L⊙ (cfr. CDPR).
When t = 15 Gyr, LB = 1010L⊙ and ϑSN = 1, the standard SNIa rate is recovered
(Tammann 1982)2.
The power Lg required to extract steadily from the galaxy the mass losses from the
aging stellar population is
The thermalization of the stellar random motions heats the gas at a power Lσ given by
Lg = α(t)Z ρ∗ΦtotdV.
(9)
Lσ =
α(t)
2 Z ρ∗(2σ2 + σ2
ϕ)dV,
while the power Lrot related to the ordered fraction of the kinetic energy is given by
Lrot =
α(t)
2 Z ρ∗vϕ
2dV.
(10)
(11)
Unfortunately this last term cannot be inserted in the χ definition in a simple way.
In fact, two opposite and extreme scenarios can be assumed: one in which also Lrot is
thermalized (contributing to the gas heating together with Lσ and LSN), and another in
which the only effect of Lrot is to reduce the depth of the effective galactic potential well.
As a consequence, the global parameter χ is defined as
χ =
Lg − γLrot
LSN + Lσ + (1 − γ)Lrot
,
(12)
with 0 ≤ γ ≤ 1. Note that if γ = 0, i.e., in the case of complete thermalization, the virial
theorem implies that Lσ + Lrot depends only on the galaxy structure and χ is independent
of the amount of ordered rotation (in our case of the value of the parameter k). When
instead γ = 1, i.e., in the case of cold rotation, the effect of rotation is maximum; CP
2Note that in our models the assumed LB is independent of h. As a consequence LSN
depends on h. In the paper we adopt h = 1/2
– 10 –
showed that also in this case, for realistic galaxy models the differences in the value of χ,
for a null and a maximum ordered velocity configuration, are small, usually less than 10%.
Moreover whenever γ > 0 the introduction of rotation decreases a χ < 1 and increases a
χ > 1. In conclusion, from the CP analysis it results that rotation alone cannot invert the
global energy balance of a gas flow, but can help the occurrence of the wind when χ < 1.
Due to the small dependence of χ on γ, through the paper we assume γ = 0.
3. The results
3.1. Model A1: ϑSN = 0.3 and k = 1
We describe here in detail the main characteristics of the flow hosted by a galaxy model
in which ordered rotation (k = 1) and SNIa heating in accordance with the latest estimates
of Cappellaro et al. (1997, ϑSN = 0.3) are present. We assume M∗ = 2.75 × 1011 M⊙,
M∗/LB = 5.5, so that LB = 5 × 1010 L⊙, and a = b = 3 kpc. The dark halo has the same
profile of the stellar distribution, i.e., ah = a and bh = b, and Mh/M∗ = 2 (see Table 1).
The three-dimensional central velocity dispersion is σc = 330 km s−1, and the maximum of
the ordered rotational velocity is vrot = 360 km s−1. This set of parameters is such that
χ = 1 at the present epoch, in order to compare as much as possible the present results
with those obtained for the so-called reference model discussed in CDPR (hereafter KRM,
χKRM = 1.1). LB is also the same of the KRM, so that the mass return rate of the two
models is exactly the same. Note how it turns out that, in order to keep χ = 1, model
A1 must have very high stellar σc and vrot. This is not a surprise:
in fact, as already
pointed out by CP, in a realistic flat model with a non-negligible ϑSN the present-epoch χ
is substantially lower than unity. The same comment applies to models A1-A4.
– 11 –
3.1.1. Hydrodynamics
As a consequence of the early high SNIa energy injection, due to the assumed ϑSN
value and αSN ∝ t−1.5, at the very beginning the flow is in the wind supersonic phase, i.e.,
the radial velocity ur(R, z) > vsound(R, z) throughout the galaxy (Fig.1, top left panel).
As time increases and the energetic input decreases according to equation (8), the
sonic surface moves outward, with ur decreasing earlier in the central regions of the galaxy,
and the gas density increases. Finally, due to the rapidly increasing radiative losses (the
so called cooling–catastrophe), at tcc ≃ 2 Gyr the fluid starts to revert its motion in a
conical region containing the polar axis. We call this phenomenon flow inversion, and the
subsequent flow decoupled inflow. The most striking feature of this evolutionary phase is
the formation of cold, dense filaments at the boundary where two streams of inflowing and
outflowing gas converge, as can be seen in Fig.1 (top right panel). This filamentary phase,
which in model A1 suddenly develops at tfil ≃ tcc, is rather short (<∼2 Gyr). Later on, no
more filaments are generated, and the previous ones are accreted at the galactic center.
However, regions of gas colder than X-ray temperature and with a low density contrast are
present for a much longer time. During this evolutionary phase two qualitatively different
regimes are present: an inner zone (r<∼15 kpc) in which the fluid is rather turbulent, and an
outer zone where the velocity field is smooth (Fig.1, bottom left panel). The inflow region
becomes progressively larger and larger. At t = 15 Gyr the gas is still outflowing in the
outskirts of the disk. The flow in the rest of the galaxy is in a turbulent state, characterized
by very subsonic velocities (v < 5 106 cm s−1, Fig.1, bottom right panel).
In order to understand better the structure of the cold filaments, we performed a
high-resolution simulation (with 300 radial meshes and 140 angular meshes covering the
same physical space as the normal resolution simulations). Because of the severe limitation
due to the azimuthal time-step, we could not extend the simulation over 15 Gyr. We thus
– 12 –
remapped onto the finer grid the hydrodynamical variables computed on the coarse grid at
a time immediately before the filament formation, and we carried on the simulation for the
entire lifetime of the filament. Fig.2 shows the temperature and velocity field at t = 2 Gyr
Density and temperature vary along the filament. For r<∼20 kpc, ρfil ∼ 6.3 10−27 g cm−3,
and the density contrast between the cold filaments and the surrounding density is of the
order of 30. The mean temperature is Tfil ≃ 104K (the lowest allowed temperature in our
simulations). At large r the density decreases and the temperature increases, but never
reaches X-ray temperatures. We did not made convergence tests, so it is plausible that the
filaments are denser and thinner. We stress that in our 2D simulations these filaments are
in fact surfaces. This fact should be kept in mind in all the following discussions3.
The cold gas accreted during the entire lifetime of the galaxy forms a small disk,
whose radius oscillates around a mean value of 3 kpc. The total mass of the cold disk is
1.5 1010 M⊙, while the hot X-ray gas is 6.3 108 M⊙ (Fig.3b, dotted line 1), much less than
the hot gas in the equivalent KRM in CDPR (∼ 5.4 109 M⊙).
3.1.2. Energetics
The comparison with the KRM is instructive. Also for this spherically symmetric
model the present day χ ≃ 1, but tcc ≃ 10 Gyr, much later than in model A1. From Fig.3d
we note that the time evolution of LX (dotted line 1) is qualitatively similar to that of
spherical models. After an initial decrease during the wind phase, LX increases up to a
maximum value at t = 2.5 Gyr. Later on, LX decreases again, in pace with the decreasing
3The formation of such filaments seems to be a rather general feature in decoupled
flows such as this. Melia, Zylstra, & Fryxell (1991), for instance, find astonishingly similar
structures in their studies on accretion disk coronae around black holes.
– 13 –
mass return from the stars. When LX reaches its maximum, a non-negligible fraction of gas
is still outflowing from the galaxy, as in the 1D models. Note that at the flow inversion the
energy budget (χ < 1) indicates that the SNIa energy injection would be able to sustain a
global outflow. This means that, because of the strong decoupling of the flow, in this model
the global energetic argument (i.e., the χ value), is not a useful indicator of the dynamical
state of the gas.
A second important difference between spherical models and model A1 must be
stressed: in the latter, LX is a factor of four fainter than LSN (Fig.3d, solid line), at variance
with 1D models in which LX may be even greater than LSN in the inflow phase. The
fainter LX of model A1 is due to its lower content of hot gas with respect to the KRM (see
§3.1.1) . This lower fraction of hot gas is due to the higher efficiency of the cooling near
the equatorial plane of model A1, a natural consequence of a higher gas density in this
region, which in turn is due to the higher local stellar density (the effect of rotation will be
discussed in §3.1.4)
We note however that, although the hot gas mass in model A1 is lower by about a
factor ten with respect to the KRM, LX does not diminish by the factor one hundred that
one could expect as a rule of thumb (the cooling per unit volume being ∝ ρ2): the much
lower than expected reduction of LX is thus due to different hot gas distributions between
the two models.
We finally point out that the noisy temporal behavior of LX is a consequence of the
density and temperature fluctuations associated to large eddies in the flow.
– 14 –
3.1.3. Surface Brightness
The X-ray surface brightness (ΣX) evolution of model A1 is shown in Fig.4 at the
same epochs of the shots in Fig.1. The optical isophotes (Σ∗) are also superimposed for
comparison, and the galaxy is seen perfectly edge-on.
During the wind phase, the X-ray isophotes are peanut-shaped, with regular lobes
aligned with the galactic polar direction (Fig.4a). The elongation of the isophotes decreases
steadily with time, they become rounder and rounder, and, in the inner regions, eventually
flatter than the optical isophotes (Fig.4b,c). The isophotes of ΣX at the present epoch are
rather flat in the central regions, and become progressively circular moving outward. This
is due to the well known fact that the emitting gas accommodates in a potential well that
is more spherical at large r. The irregularities in the isophotal shapes are clear evidence of
the inner turbulence (Fig.4d).
Note that once the cold filaments appear (Fig.4b,c), ΣX becomes very disturbed: we
stress however that such an effect is magnified in our simulations by the imposed geometry,
in which the filaments are in fact surfaces. We also computed ΣX under different viewing
angles. It turns out that at early epochs the X-ray isophotes appear still elongated along
the optical minor axis, even for values of the viewing angle as low as 45◦ (90◦=edge-on).
Instead, at late epochs, Σ∗ and ΣX become very similar as soon as the inclination angle is
slightly lower than 90◦.
3.1.4. Model A2: ϑSN = 0.3 and k = 0
We describe here a model very similar to A1, but with k = 0, all the galaxy flattening
being due to σ2
ϕ (see Table 1). For this model the value of χ is independent of any
assumption on the thermalization of ordered motions [see eq. (12) with γ = 0], and is equal
– 15 –
to unity.
From Fig.3b it is apparent how this non-rotating model has a higher total gas mass
(dashed line 2) than the rotating model A1 (dashed line 1). This is because rotation - if
not entirelly thermalized - decreases a χ < 1 (see §2.3), and in fact the largest difference
between dashed line 1 and dashed line 2 takes place during the initial wind phase, when
χ is substantially lower than unity. Also the hot gas mass of model A2 (dotted line 2) is
higher than that of model A1 (dotted line 1): this is due to the combined effect of rotation,
already discussed, and of the presence of a slightly more massive cold disk in model A1, a
consequence of angular momentum conservation. This effect is quite small in the models
heated by SNIa, but much more important in the cooling-flow models (a more detailed
discussion is postponed to §§3.2.2-3.2.3).
The main features of the temporal evolution of LX are similar for models A1 and A2
(Fig.3d dotted line 2). We note however that the LX of model A2 is slightly higher than
that of model A1, in accordance with its higher content of hot gas. The flow dynamical
evolution of model A2 is also quite similar to that of model A1. The transient phase
of filamentary instabilities – although to a minor extent – is still present, and develops
approximately at the same time as in model A1. The absence of rotation does not affect the
presence of oscillations in the temporal evolution of LX, thus indicating that the turbulence
is mainly due to a decoupled inversion of the gas flow, rather than to rotation. Finally, ΣX
evolves mainly as in model A1, although the X-ray isophotes appear more circular in this
non-rotating case.
– 16 –
3.1.5. Models A3 and A4: ϑSN = 0.3 and k = 1, but different a/b's
We focus here on the relationship between the galaxy's flattening and the development
of the cold instabilities. To investigate this point we run two models similar to model A1,
except for the ratio a/b (see Table 1). A quantitative measure of the flattening of a MN
model in its central regions is obtained with a series expansion of eq. (1). Retaining the
quadratic terms the resulting isodensity surfaces are similar ellipsoids whose ellipticity is
Eρ = 1 −s
s + 5
(s + 1)(s2 + 4s + 5)
,
(13)
and s = a/b. Thus s = 1 adopted in models A1 and A2 corresponds to Eρ = 0.45. As
a consequence, with s = 0.7 (and Eρ = 0.36) model A3 is slightly more spherical than
model A1, and its χ = 1.2 is higher, as expected. The result is that the flow inversion
occurs earlier, and the filamentary phase is already completely developed at tfil = 1.4 Gyr.
On the contrary model A4, with s = 1.3 and and Eρ = 0.52 is flatter. Its χ = 0.94, and
the filamentary phase starts now at tfil = 3 Gyr. For a short time this model shows the
contemporary presence of two cold structures similar to that in Fig. 2. These two models
show how small changes in the galaxy flattening can change the flow evolution in the CDPR
scenario.
3.2. Model B1: ϑSN = 0 and k = 1
A very different flow evolution is obtained by suppressing the SNIa heating in a model
equal to A1: now χ = 4 since the beginning, and the behaviour is that of a pure cooling
flow. The velocity field is very smooth, thus showing that SNIa's play a fundamental role
in maintaining a boiling hot gaseous halo. In this case we do not obtain the transient
filamentary structure, and LX decreases smoothly. The present day LX is more than one
order of magnitude lower than in the previous SNIa heated cases (Fig.3c, dashed line),
– 17 –
because there is considerably less hot gas in the galaxy (Fig.3a, dashed line). The cooling
gas accumulates now on an extended cold disk of ≃ 1011 M⊙, approximately ten times more
massive than that of model A1. Such a disk, present also in the simulations of BM and KM,
is not actually observed in real galaxies. The X-ray isophotes of model B1 closely follow the
optical ones.
3.2.1. Model B2: ϑSN = 0 and k = 0
Model B2 is analogous to model B1, but no ordered rotation is applied. In this case LX
is higher than in model B1 at every time (Fig.3c, dotted line), just due to the larger amount
of hot gas (Fig.3a, dotted line). At first sight it could seem contradictory that the model
with the lower LX (B1) has cooled a larger amount of gas, i.e., one could expect a higher
LX, at least in the past. However, although more mass cools per unit time in the rotating
model, the lack of thermalization of the gravitational energy of the gas which stops on the
rotating cold disk instead of falling into the galaxy center produces a net quenching in the
global LX (as pointed out by BM). The ΣX isophotes in the non-rotating case are more
concentrated towards the center and rounder than in the rotating case, as found by BM.
It is of interest to compare our cooling flow models with those of BM. These authors
find a dramatic reduction of LX when introducing rotation, but at variance with us, the hot
gas mass in their rotating models decreases only slightly compared to the non-rotating ones.
Thus, the lower luminosity in their rotating models is due essentially to a redistribution
of the hot gas mass, rather than to its scarcity as in ours. The origin of this different
behaviour is in the higher flattening of the models used here (i.e., in the presence of the disk
in the MN density distribution). In fact, in our models a larger fraction of gas is produced
at low-z over the galactic equatorial plane. When the model is rotating this gas falls on the
disk approximately at the same distance from the center where it was formed, bacause of
– 18 –
angular momentum conservation. So, on one hand it releases a much lower gravitational
energy, and, on the other, the hot gas distribution is changed only slightly with respect to
the non-rotating case. The accumulation of the gas on the disk also significantly increases
its cooling, and the final result is a reduction of the hot gas mass. Note that also in the BM
models the amount of cold gas increases with an increasing ellipticity of the parent galaxy.
3.3. Models C1 and D1: ϑSN = 0.3 and k = 1
As mentioneded in the Introduction, we also ran models which are similar in many
aspects to A1, but with larger b's, in order to have realistic values for the velocity dispersion
(see Table 1). Since ρ∗(0) = M∗(3 + s)/4πb3(1 + s)3, increasing b at fixed s produces a
reduction in the gas density and thus in the cooling, because of the lower stellar density.
At the same time, because Φ∗(0) = GM∗/b(1 + s), Lg is reduced too, favoring the escape of
the gas.
Model C1 (b = 4 kpc, σc = 290 km/s, vrot = 309 km/s, χ = 0.85) develops a decoupled
inflow at 5.5 Gyr, as can be seen from the evolution of LX (Fig.3d, heavy dotted line 3).
The most striking feature is again the formation of cold filaments (Fig.5). Surprisingly
enough, these filaments do not form at the moment of flow inversion (tcc ≃ 5.5 Gyr), but
quite later (tfil ≃ 10.5 Gyr). At variance with model A1, they form close to the galactic
equatorial plane (cfr. Fig.2). Once on the plane, they slowly drift towards the center and
accrete on a small cold disk (r ∼ 4 kpc, Mdisk ≃ 5 109 M⊙). This dynamical phase is
characterized by a conspicuous and steady decrease of LX, a consequence of the decrease
of the hot gas mass (Fig.3b, heavy dotted line 3). While this effect is certainly real, it is
magnified in our simulations for two reasons. First, as already pointed out, the filaments
are in fact cylindrical surfaces; second, as usual in 2D simulations, reflecting boundary
conditions on the disk introduce non-physical effects favoring cold material accumulation.
– 19 –
The ΣX evolution is similar to that of model A1, but postponed in time.
For model D1 (b = 5 kpc, σc = 258 km/s, vrot = 277 km/s, χ = 0.7), the resulting flow
is a pure wind all over the galaxy, lasting for all the Hubble time without any decoupling.
All the gas is hot, and its mass – as is common in the wind phase – decreases steadily and
is very low (Fig.3b, heavy line 4). LX is also very low and decreases proportionally to the
gas mass (Fig.3d, heavy dotted line 4). ΣX is similar to that shown in Fig. 4a. This run
clearly shows how in a realistic flat model, even with a low ϑSN, the flow can be in the wind
phase during its entire evolution.
3.3.1. Models C2 and D2: ϑSN = 0.3 and k = 1, but different a/b's
We now briefly describe the behaviour of two models similar to C1 and D1, but with
different flattening (see Table 1). The first (C2, s = 1.3, Eρ = 0.52, χ = 0.75) is flatter than
model C1, and consequently has a lower χ. The resulting flow is a wind during all the time.
This model (with b = 4 kpc) is energetically very similar to D1 (s = 1 and b = 5 kpc), as
testified by their χ values, and its flow evolution is also similar.
Model D2 (b = 5 kpc, s = 0.6, Eρ = 0.33, χ = 0.86) is obtained from D1 by decreasing
its flattening. As a consequence the gas is more bound as testified by a χ similar to that
of model C1 (s = 1 and b = 4 kpc). The resulting evolution is similar to that of model
C1, with the flow inversion and the following development of a decoupled inflow around
tcc ≃ 5.5 Gyr. However, in this case no cold filaments are developed. In conclusion, a very
small change in the flattening can compensate for a variation of the characteristic scale
length, showing again the strong sensitivity of models to variations of galactic structural
parameters in the CDPR scenario.
– 20 –
4. Discussion
Although the number of models we ran is not sufficient for an accurate statistical and
quantitative analysis of the gas flow properties in flat and rotating galaxies, a qualitative
comparison with the spherical models discussed in CDPR and with cooling flow models
is possible; it is also interesting to compare our results with recent observational findings
(Pellegrini, Held & Ciotti 1997).
In order to compare the behavior of S0 models with the equienergetical spherical
models discussed in CDPR, we performed calculations for model galaxies with χ = 1 at the
present epoch (models A1-A2). We stress again that while χ = 1 is a quite natural product
of realistic spherical models, it turns out to be rather extreme for flat models. In any case,
in flat models a clear X-ray underluminosity with respect to their spherical counterpart
is found, mainly due to a drastic reduction of the hot gas mass. The amount of galactic
rotation contributes only slightly to such a reduction.
We also ran a few models with ϑSN = 0 (models B1-B2). These have a remarkably low
LX, in contrast with spherical cooling flows which in general show excessively high X-ray
luminosities. Although the model without rotation displays a present day LX similar to
that of the rotating model, it was much more luminous in the past. As a consequence, with
different choices of the structural and dynamical parameters, a large difference in the LX
can be obtained. Such a possibility is supported also by the models of BM, where different
degrees of rotation lead to a spread in LX.
When the galaxy structure and dynamics are more similar to those of observed galaxies,
and SNIa's are present, χ is low, and the galaxy is permanently in a global wind phase or
loses a large amount of gas during a Hubble time; the flow decoupling takes place at late
times, even for ϑSN as low as 0.3. The qualitative predictions made in CP are thus correct:
it is much easier to extract gas from a flat, realistic galaxy model, than from a rounder
– 21 –
system of the same LB. LX may span all the values in the range between that of a pure
wind and approximately LSN by, for instance, varying b or the galaxy flattening, similarly to
what happens for spherical models in the CDPR scenario for small variations of the model
parameters.
Moreover, the X-ray underluminosity of flat galaxies can naturally be accounted for
because their LX never reaches LSN, even in the inflow phase. A possible explanation of
the observations could be that at low LB, where the CDPR scenario predicts a significant
number of objects to have an energy budget sufficient to unbind the gas, flat galaxies are
preferentially found in the wind phase with respect to spherical galaxies of the same optical
luminosity, while at high LB, where a significant number of galaxies contain inflows, the
underluminosity of flat galaxies could be due to the underluminosity of their inflow phase,
as found (albeit for different reasons) in our models and by BM.
We computed also a few models (not shown here) with a different flattening of the dark
matter distribution with respect to the stellar component. The results are qualitatively
similar to those discussed above. As a general rule, for a given stellar distribution, a rounder
halo increases χ and reduces tfil. For example, a model similar to A1 but with a spherical
halo (a/b)h = 0, has χ = 1.14 and tfil ≃ tcc = 1 Gyr; a model similar to C2 but with a
rounder halo (a/b)h = 0.6, has χ = 0.8 and tfil ≃ 6.2 Gyr. As expected, as the flattening of
the dark halo increases, χ decreases and tfil increases: for example, a model similar to D2
but with (a)
¯h
of χ = 0.79.
= 1.3 always remains in the wind phase, in accordance with the lower value
We conclude that, on the basis of LX alone, it is not possible to decide whether the
CDPR scenario or the cooling flow scenario is more suitable to describe the flows in S0s
and Es: both can reproduce the X-ray underluminosity of flat galaxies. The scatter in
the LX − LB diagram (at a fixed optical luminosity) is mainly produced by the galactic
– 22 –
flattening (and rotation to a minor extent) in the CDPR scenario, and by the amount of
rotation in the cooling flow models.
In any case, the current explanations of the observational results are unsatisfactory. In
fact, Pellegrini et al. (1996) pose unsolved problems, especially concerning the explanation
of the scatter in the LX − LB diagram, as a function of the flattening and rotation. It is
found that LX/LB does not correlate with rotation and/or flattening. On the contrary, a
very strong segregation effect is found: only galaxies with small flattening and low vrot/σc
show a scatter in LX/LB, while flat objects and strongly rotating ones are systematically
found only at low LX/LB. The main factor in determining the LX/LB segregation (and
scatter, as a consequence) could be related to the slope of the galaxy surface brightness
profile in the very center, as measured by HST observations (Pellegrini 1997, and references
therein). This would tell us that a nuclear origin for the X-ray behavior of early-type
galaxies should be taken into proper account (Ciotti & Ostriker 1997). It is remarkable
how the independently chosen separation of early-type galaxies in cuspy and core galaxies,
a separation operated using only photometric criteria, reproduces exactly the segregation
in the LX − LB plane: while core galaxies show the entire scatter in LX at fixed LB, cuspy
galaxies are invariably found only at low LX.
Other problems affect both the CDPR models and the cooling flow models. For
example, the absence of massive cold disks in Es seems to support the CDPR scenario,
while the claimed very low metallicity in the hot gas is against it, because a low metallicity
contrasts with the assumption of a non negligible value of ϑSN. In fact, in principle,
constraints on the SNIa rate can be given by estimates of the iron abundance in galactic
flows (see, e.g., Renzini et al. 1993, and references therein). Under the assumption of solar
abundance ratios, the analysis of the available data suggests a very low iron abundance,
consistent with no SNIa's enrichment and even lower than that of the stellar component
– 23 –
(Ohashi et al. 1990; Awaki et al. 1991; Ikebe et al. 1992; Serlemitsos et al. 1993;
Loewenstein et al. 1994; Awaki et al. 1994; Arimoto et al. 1997; Matsumoto et al. 1997).
This is rather puzzling, in that a value as high as ϑSN = 0.3 agrees with the current
optical estimates of the SNIa rates. However, some authors have found that more complex
multi-temperature models with higher abundance give a better fit to the data (Kim et
al. 1996; Buote & Fabian 1997). So, there are arguments on both sides of the abundance
question, which is far from being a closed issue. In any case it is clear that the CDPR
scenario is ruled out in case a null SNIa rate is clearly determined.
An interesting feature of our models, that could be used to discriminate between the
CDPR and the cooling flow scenario, is the spontaneous occurrence of the flow decoupling.
This is at the origin of the striking feature of the transient cold filaments, which – if caught
during their formation – should be observed first as soft X-ray emitting structures, and
subsequently as Hα filaments. In model A1 this takes place at tfil ≃ 2 Gyr, while in model
C1 (the model with b = 4 kpc but similar to A1 in all other input parameters) it occurs
at tfil ≃ 10 Gyr. Due to this sensibility to the galaxy structure, it is plausible that for a
few galaxies the decoupling of the flow takes place at the present epoch, and so these Hα
features could be observed4. Indeed, it is tempting to identify the transient cold filaments
with the soft X-ray clumps (Kim & Fabbiano 1995) or Hα filaments (Trinchieri & di Serego
Alighieri 1991, Trinchieri, Noris, & di Serego Alighieri 1997) recently observed. We note
how the size of the numerical and observed structures (of the order of ∼ 10 kpc) are
surprisingly similar. This could be an alternative explanation to the galactic drips proposed
by Mathews (1997).
4Note that a flow decoupling – due to different physical reasons – is shown also in spherical
power-law galaxies with low dark matter content and low SNIa heating (Pellegrini & Ciotti
1997).
– 24 –
A final comment is on the X-ray surface brightness distributions of the computed
models. As already found by BM for cooling flow models, rotation is the key factor in
determining whether ΣX is more or less round than Σ∗, even in the presence of SNIa
heating: rotation produces a flatter ΣX. A favourable feature of the CDPR scenario is
that outflowing galaxies have rounder X-ray isophotes, which alleviates the problem of
excessively flat ΣX profiles, affecting our cooling flow models as well as the models of KM
and BM. A detailed observational analysis of the relationship between optical and X-ray
surface brightness for the E4 galaxy NGC 720 and S0 galaxy NGC 1332 has been carried out
by Buote & Canizares (1997, and references therein). While data for NGC 1332 are more
uncertain, an analysis of the X-ray isophotes of NGC 720 shows that they are misaligned
with and rounder than the optical ones. The authors conclude that the shape of ΣX is due
to the flattening of the dark halo, which is higher than that of the stellar component. In
fact, our simulations shows that, if rotation is the origin of flattening of the X-ray isophotes,
then a decreasing flattening of ΣX with galactocentric radius is expected, at variance with
the observations. On the other hand, in our simulations flatter dark halos produce (during
the inflow phase) flatter X-ray isophotes, as expected.
5. Conclusions
We have performed 2D numerical simulations of gas flows in S0 galaxy models with
different flattening, following the same recipe for the input physics adopted in CDPR, and
exploring the effects of different amounts of ordered rotation and SNIa energy injection
rates, and structural parameters. As a general rule, the WOI sequence which occurs in the
1D models is replaced by a wind phase followed by a decoupled inflow. The main results
can be summarized as follows:
1. As expected models without SNIa heating (χ = 4) host a cooling flow from the
– 25 –
beginning. The flow is considerably ordered, and essentially all the gas lost by the stars
is cooled and accumulated in the galactic center. If rotation is present, the cold material
settles in a disk on the galactic equatorial plane.
2. Models with SNIa heating and χ ≃ 1 at the present epoch, invert their flow much
earlier than the spherical ones with the same energy budget. In particular, as the SNIa
rate decreases, the flow tends to revert its motion starting from the galactic polar regions.
In this phase of strong decoupling, cold filaments are created at the interface between
inflowing and outflowing gas. This strongly unstable phase lasts ∼ 2 Gyr, after which the
filaments are definitively accreted on the galactic center. Later on, the turbulences are
unable to create regions of very cool, dense gas. The temperature may however fluctuate
and gas cooler than X-ray temperatures is present. In the rotating model, the presence of
SNIa greatly reduces the formation of the cold disk. In general, less cold matter is present
in these models (≃ 1010 M⊙) than in those with χ = 4 (≃ 1011 M⊙).
3. Models with realistic values of the dynamical quantities are characterized by low
χ values. While for model D1 (b = 5 kpc, χ = 0.7) the flow is always escaping without
displaying any interesting feature, in model C1 (b = 4 kpc, χ = 0.85), the flow inversion
occurs at tcc = 5.5 Gyr, and filaments form at tfil ≃ 10.5 Gyr. Of course, no cold mass is
present in the first model; in the second, 5 109 M⊙ of cold gas form a small disk.
4. For any fixed value of ϑSN, rotating models are X-ray fainter than the non-rotating
ones, for the presence of the cold disk as well as for a different distribution of the hot gas.
When compared to 1D models, 2D models are much fainter in the X-rays. The former
flows, in fact, have LX similar to or even higher than LSN after the flow inversion; the latter,
instead, always have LX < LSN. This is due to several combined factors: the loss of a SNIa
energy fraction which is emitted by the gas below the X-ray temperature in high density
regions (which are much larger than in spherical models); the loss of a SNIa energy fraction
– 26 –
used by the gas still outflowing from the external regions of the disk; the corresponding
gravitational energy not released by this escaping material, and, for rotating models, the
loss of thermalization of the gravitational energy of the cold disk. Another main difference
of the temporal evolution of LX is its noisy behaviour in 2D models. This is due to the
presence of turbulence and to the formation of cold filaments.
5. We have shown that the flattening of the galaxy in the CDPR scenario allows for
a decoupling of the flow, in which different flow regimes are present in different part of the
galaxy at the same time. We have found also that the flow decoupling is a general feature,
i.e., all our models, except those without SNIa heating and the model with χ substantially
lower than unity, show a flow decoupling. We note however that the LX associated to a
decoupled flow when χ < 1, is not high:
it is true, on the contrary, that the flows are
decoupled even when χ > 1, as the galaxy loses gas from the disk even with a low SNIa rate.
So, the energetical analysis carried out by CP is shown to be qualitatively verified: objects
with a low χ value are found preferentially at low LX, and at variance with the cooling
flow models the effect of the flattening is shown to be important in determining the flow
evolution.
6. For non-rotating, inflow models, the edge-on ΣX is rounder than the corresponding
Σ∗ of the parent galaxy; when rotation is present, ΣX becomes flatter, and very similar to
Σ∗. For models with SNIa heating, the ΣX morphology evolves with time, from a smooth,
peanut-shaped geometry to a generally rounder shape than the stellar profile, through
disturbed phases during the presence of cold filaments. When rotation is present, in the
galagtic center the ΣX isophotes are somewhat flatter than Σ∗, at variance with non-rotating
models in which ΣX remains more spherical up to the center. As a general rule, even for
low line-of-sight inclination angles over the galactic equatorial plane, the isophotes of ΣX
and Σ∗ tend to become similar.
– 27 –
All these new features considerably enrich the old CDPR scenario, and represent a step
further in the direction of a more realistic simulation of a situation that is proving to be
very complex. In future, we plan to perform numerical simulations of gas flows in models
tailored on observed X-ray emitting S0 galaxies, in order to improve our understanding of
the much richer phenomenology of aspherical gas flows.
We would like to thank James Binney, Fabrizio Brighenti, Silvia Pellegrini, and Ginevra
Trinchieri for useful discussions, and the referee for comments that improved the paper. We
are indebted to Giovanna Stirpe who carefully read the manuscript. We would also like to
thank the CINECA Computing Center for having kindly provided the CRAY C90 for our
numerical computations. This work was partially supported by the Italian MURST and the
Italian Space Agency (ASI) trough grant ASI-95-RS-152.
– 28 –
REFERENCES
Arimoto, N., Matsushita, K., Ishimaru, Y., Ohashi, T., Renzini, A., 1997, ApJ, 477, 128
Awaki, H., Koyama, K., Kunieda, H., Takano, S., Tawara, Y., & Ohashi, T., 1991, ApJ,
366, 88
Awaki, H., et al., 1994, PASJ, 46, L65
Binney, J., Tremaine, S., 1987, "Galactic Dynamics", Princeton Univ. Press: Princeton
Brighenti, F., Mathews, W.G., 1996, ApJ, 470, 747 (BM)
Buote, D.A., & Canizares, C.R., 1997, ApJ, 474, 650
Buote, D.A., & Fabian, A.C., 1997, submitted to MNRAS, (astro-ph/9707117)
Cappellaro, E., Turatto, M., Tsvetkov, D.Y., Bartunov, O.S., Pollas, C., Evans, R., Hamuy,
M., 1997, A&A, 322, 431
Ciotti, L., D'Ercole, A., Pellegrini, S., Renzini, A., 1991, ApJ, 376, 380 (CDPR)
Ciotti, L., Ostriker, J.P., 1997, ApJ, 487, L105
Ciotti, L., Pellegrini S., 1996, MNRAS, 279, 240 (CP)
Eskridge, P., Fabbiano, G., Kim, D.W., 1995a, ApJS, 97, 141
Eskridge, P., Fabbiano, G., Kim, D.W., 1995b, ApJ, 442, 523
Fabbiano G., 1989, ARA&A, 27, 87
Fabbiano, G., Kim D.W., Trinchieri G., 1992, ApJS, 80, 531
Ikebe, Y., et al., 1992, ApJ, 384, L5
– 29 –
Kim, D.W., Fabbiano, G., 1995, ApJ, 441, 182
Kim, D. W., Fabbiano, G., Matsumoto, H., Koyama, K., & Trinchieri, G., 1996, ApJ, 468,
175
Kley, W., Mathews, W.G., 1995, ApJ, 438, 100 (KM)
Loewenstein, M., Mushotzky, R., Tamura, T., Ikebe, Y., Makishima, K., Matsushita, K.,
Awaki, H., & Serlemitsos, P.J., 1994, ApJ, 436, L75
Mathews, W.G., 1997, AJ, 113, 755
Mathews, W.G., Bregman, J.N., 1978, ApJ, 224, 308
Matsumoto, H., Koyama, K., Awaki, H., Tsuru, T., Loewenstein, M., Matsushita, K., 1997,
ApJ, 482, 133
Melia, F., Zylstra, G.J., Fryxell, B., 1991, ApJ, 377, L101
Miyamoto, M., Nagai, R., 1975, PASJ, 27, 533 (MN)
Ohashi, T., et al., 1990, in Windows on Galaxies, ed. G. Fabbiano, J.A. Gallagher, & A.
Renzini (Dordrecht: Kluwer), 243
Pellegrini, S., Held, E.V., Ciotti, L., 1997, MNRAS, 288, 1
Pellegrini, S., Ciotti, L., 1997, A&A, submitted
Pellegrini, S., 1997, Third ESO Whorkshop "Galaxy scaling relations: origins, evolution
and applications", Garching bei Muenchen, 1996, in press
Plummer, H.C., 1911, MNRAS, 71, 460
Renzini, A., Ciotti, L., D'Ercole, A., Pellegrini, S., 1993, ApJ, 419, 52
– 30 –
Satoh, C., 1980, PASJ, 32, 41
Serlemitsos, P. J., Loewenstein, M., Mushotzky, R., Marshall, F., & Petre, R., 1993, ApJ,
413, 518
Tammann, G., 1982, in "Supernovae: A Survey of Current Research", ed. M. Rees & R.
Stoneham (Dordrecht: Reidel), p.371
Trinchieri, G., di Serego Alighieri, S., 1991, AJ, 101, 1647
Trinchieri, G., Noris, L., di Serego Alighieri, S., 1997, A&A, in press
van den Bergh, S., McClure, R.D., 1994, ApJ, 425, 205
This manuscript was prepared with the AAS LATEX macros v4.0.
– 31 –
Fig. 1.- The gas density and velocity fields of model A1 at four different times (0.46, 2.16,
3.37, 15) Gyr (top left, top right, bottom left, bottom right). In the top left panel the dashed
lines represent the stellar density distribution. In the second panel the large high density
transient structure is apparent. Note how the velocity field in the second and third panels
is dominated by large vortices, while in the fourth panel the velocity is very low, and nearly
random.
Fig. 2.- The gas temperature and velocity for model A1 (high resolution simulation) at
t ≃ 2 Gyr, when the cold filament is completely developed. Note how the disk is still strongly
degassing.
Fig. 3.- The time evolution of the mass and luminosity of gas flows for the models discussed
in the text. Panel (a) illustrates the behaviour of the hot gas of model B1 (ϑSN = 0, k = 1;
dashed line), and of model B2 (ϑSN = 0, k = 0; dotted line). The solid line represents the
total mass lost by the stars. Panel (c) represents the LX evolution for the same models.
Panel (b) shows the mass budget for the models with ϑSN = 0.3. Dotted lines represent the
evolution of the hot gas mass for model A1 (k = 1, line 1); A2 (k = 0, line 2); C1 (k = 1,
line 3); D1 (k = 1, line 4). Dashed lines show the total gas mass inside the galaxy. The solid
line again represents the total mass lost by the stars. Panel (d) shows the evolution of LX
for these models (dotted lines), and the solid line represents LSN.
Fig. 4.- The optical (solid lines) and X-ray (dotted lines) surface brightness distributions
for model A1, at the same representative times as in Fig. 1 (time increases from top to
bottom). The optical isophotes are separated by one magnitude. The X-ray isophotes are
logarithmically spaced: the lowest value (-8.51 in cgs units) is the same in all panels, the
maximum is -3.34 (a), -2.97 (b), -3.37 (c), -4.47 (d).
Fig. 5.- The gas temperature and velocity for model C1 at t ≃ 10 Gyr, when the cold
– 32 –
filament is completely developed.
– 33 –
Table 1. Models parameters.
Modela
bb
a/b
c
σc
vrot
d ϑSN
k
χe
f
tcc
f
tfil
A1
A2
A3
A4
B1
B2
C1
C2
D1
D2
3
3
3
3
3
3
4
4
5
5
1
1
0.7
1.3
330
360
330
0
376
365
300
347
1
1
330
360
330
0
1
290
309
1.3
260
300
1
258
277
0.6
305
283
0.3
0.3
0.3
0.3
0.0
0.0
0.3
0.3
0.3
0.3
1
0
1
1
1
0
1
1
1
1
1
1
1.20
0.94
4
4
2
1.7
1.4
3
0
0
2
1.7
1.4
3
...
...
0.85
5.5
10.5
0.75
0.70
...
...
0.86
5.5
...
...
...
aAll models have LB = 1010L⊙, M∗ = 1011 M⊙, and Mh/M∗ = 2
bin kpc
cthree-dimensional central velocity dispersion in km s−1
dmaximum rotational velocity in km s−1
efor the rotating models computed under the assumption of
complete thermalization of ordered motions
f in Gyr
60
40
20
0
40
20
0
0
20
40
0
20
40
60
R
R
1
0
-1
-2
3
2
1
0
a
-1
c
b
d
0
5
10
0
0
5
5
10
10
15
15
t (Gyr)
t (Gyr)
t (Gyr)
30
20
10
30
20
10
30
20
10
30
20
10
a
b
c
d
20
40
R
60
|
astro-ph/9901160 | 1 | 9901 | 1999-01-13T10:12:00 | The Cosmic Ray Proton Spectrum determined with the Imaging Atmospheric Cherenkov-Technique | [
"astro-ph"
] | The HEGRA system of 4 Imaging Atmospheric Cherenkov Telescopes (IACTs) has been used to determine the flux and the spectrum of cosmic ray protons over a limited energy range around 1.5 TeV. Although the IACT system is designed for the detection of gamma-rays with energies above 500 GeV, it has also a large detection area of $ \simeq 10^6 m^2 \cdot 3 msr$ for primary protons of energies above 1 TeV and the capability to reconstruct the primary proton energy with a reasonable accuracy $\Delta$E/E of 50% near this threshold. Furthermore, the principle of stereoscopic detection of air showers permits the effective suppression of air showers induced by heavier primaries already on the trigger level, and in addition on the software level by analysis of the stereoscopic images. The combination of both capabilities permits a determination of the proton spectrum almost independently of the cosmic ray chemical composition. The accuracy of our estimate of the spectral index at 1.5 TeV is limited by systematic uncertainties and is comparable to the accuracy achieved with recent balloon and space borne experiments. In this paper we describe in detail the analysis tools, namely the detailed Monte Carlo simulation, the analysis procedure and the results. We determine the local (i.e. in the range of 1.5 to 3 TeV) differential spectral index to be $\gamma_p = 2.72\pm 0.02_{stat.}{\pm 0.15}_{syst.}$ and obtain an integral flux above 1.5 TeV of $F(> 1.5 TeV) =3.1\pm 0.6_{stat.}\pm 1.2_{syst.} 10^{-2}/s sr m^2$. | astro-ph | astro-ph |
The Cosmic Ray Proton Spectrum
determined with the
Imaging Atmospheric Cherenkov-Technique
F. Aharonian1, A.G. Akhperjanian7,1, J.A. Barrio3,2, A.S. Belgarian7, K. Bernlohr1,9,
J.J.G. Beteta3, H. Bojahr6, S. Bradbury2,8, I. Calle3, J.L. Contreras3, J. Cortina3, A.
Daum1,11, T. Deckers5, S. Denninghoff2, V. Fonseca3, J.C. Gonzalez3, G. Heinzelmann4, M.
Hemberger1,11, G. Hermann1,13, M. Hess1,11, A. Heusler1, W. Hofmann1, H. Hohl6, I. Holl2,
D. Horns4, A. Ibarra3, R. Kankanyan1,7, M. Kestel2, O. Kirstein5, C. Kohler1, A.
1Max-Planck-Institut fur Kernphysik, Saupfercheckweg 1, D-69117 Heidelberg, Germany
2Max-Planck-Institut fur Physik, Fohringer Ring 6, D-80805 Munchen, Germany
3Universidad Complutense, Facultad de Ciencias F´ısicas, Ciudad Universitaria, E-28040 Madrid,
Spain
4Universitat Hamburg, II. Institut fur Experimentalphysik, Luruper Chaussee 149, D-22761 Ham-
burg, Germany
5Universitat Kiel, Institut fur Physik, Leibnitzstr. 15, D-24118 Kiel, Germany
6Universitat Wuppertal, Fachbereich Physik, Gaussstr. 20, D-42097 Wuppertal, Germany
7Yerevan Physics Institute, Yerevan, Armenia
8Now at Department of Physics University of Leeds, Leeds LJ2 9JT, UK
9Now at Forschungszentrum Karlsruhe, P.O. Box 3640, 76021 Karlsruhe, Germany
10On leave from Altai State University, Barnaul, Russia
11Now at SAP AG, Neurottstr. 16, D-69190 Walldorf, Germany
12Now at Universidad Aut´onoma de Barcelona, Institut de F´ısica d'Altes Energies, E-08193 Bel-
laterra, Spain
13Now at Enrico Fermi Institute, The University of Chicago, 933 East 56th Street, Chicago, IL,
60637, USA
1
Konopelko1,10, H. Kornmeyer2, D. Kranich2, H. Krawczynski1,4, H. Lampeitl1, A. Lindner4,
E. Lorenz2, N. Magnussen6, H. Meyer6, R. Mirzoyan2, A. Moralejo3, L. Padilla3, M.
Panter1, D. Petry2,12, R. Plaga2, A. Plyasheshnikov1,10, J. Prahl4, C. Prosch2, G.
Puhlhofer1, G. Rauterberg5, C. Renault1, W. Rhode6, A. Rohring4, V. Sahakian7, M.
Samorski5, D. Schmele4, F. Schroder6, W. Stamm5, H.J. Volk1, B. Wiebel-Sooth6, C.A.
Wiedner1, M. Willmer5, H. Wirth1
(April 22, 2018)
The HEGRA system of 4 Imaging Atmospheric Cherenkov Telescopes
(IACTs) has been used to determine the flux and the spectrum of cosmic ray
protons over a limited energy range around 1.5 TeV. Although the IACT sys-
tem is designed for the detection of γ-rays with energies above 500 GeV, it has
also a large detection area of ≃ 106 m2 · 3 msr for primary protons of energies
above 1 TeV and the capability to reconstruct the primary proton energy with
a reasonable accuracy ∆E/E of 50% near this threshold. Furthermore, the
principle of stereoscopic detection of air showers permits the effective suppres-
sion of air showers induced by heavier primaries already on the trigger level,
and in addition on the software level by analysis of the stereoscopic images.
The combination of both capabilities permits a determination of the proton
spectrum almost independently of the cosmic ray chemical composition. The
accuracy of our estimate of the spectral index at 1.5 TeV is limited by sys-
tematic uncertainties and is comparable to the accuracy achieved with recent
balloon and space borne experiments. In this paper we describe in detail the
analysis tools, namely the detailed Monte Carlo simulation, the analysis pro-
cedure and the results. We determine the local (i.e. in the range of 1.5 to 3
TeV) differential spectral index to be γp = 2.72±0.02stat.±0.15syst. and obtain
an integral flux above 1.5 TeV of F ( > 1.5 TeV) = 3.1±0.6stat. ±1.2syst. ·10−2/s
sr m2.
96.40, 95.85.S, 98.70.S
2
I. INTRODUCTION
The stereoscopic system of Imaging Atmospheric Cherenkov telescopes (IACT-system)
of the HEGRA collaboration [1] is a powerful tool for detecting TeV γ-ray sources and for
performing detailed spectroscopic studies in the energy range from 500 GeV to ∼ 50 TeV,
where the latter limit is determined by event statistics alone. With the nearly background-
free detection of γ-rays from the Crab Nebula [1], an energy flux sensitivity ν Fν of ≃ 10−11
ergs/cm2 s at 1 TeV for one hour of observation time has been estimated. The high signal to
noise ratio together with the energy resolution of better than 20% for primary photons makes
it possible to study the spectra of strong sources on time scales of one hour, as demonstrated
by the observation of the BL Lac object Mkn 501 during its 1997 state of high and variable
emission [2].
The IACT system can not only be used for γ-ray astronomy. It can also contribute to
the study of charged cosmic rays (CR) for energies between a few TeV and possibly ∼ 100
TeV, a key energy region for the understanding of the sources of CRs and their propagation
through our galaxy (see e. g. [3,4] and references therein for reviews).
The measurement with the IACT system described in this paper has systematic uncer-
tainties comparable to recent measurements of satellite and balloon borne experiments (see
e. g. [5] for a recent compilation). A clear advantage of the IACT technique is the large
effective area of ≃ 3·103 m2 sr for TeV cosmic rays combined with a field of view of ≃ 3 msr,
corresponding to a detection rate of around 12 Hz for > 1 TeV cosmic rays.
In an earlier paper [6] we explored the possibilities to use the IACT technique to mea-
sure the energy spectra and mass composition of CRs and especially CR protons. The
stereoscopic observation of air showers with at least two IACTs suppresses heavier primaries
already on the trigger level. This is because the energy threshold Ethr, defined as the energy
where the differential detection rate peaks, increases substantially with the nucleon num-
ber A, approximately as Ethr ∝ A0.5. The stereoscopic detection of the air shower under
different viewing angles with high resolution imaging cameras permits us to unambiguously
3
reconstruct the air shower axis in three dimensions. Knowing the location of the shower
core with a precision of 30 m, the energy of a primary proton can be determined with an
accuracy ∆E/E of 50% and the different projections of the longitudinal and lateral shower
development can be used to obtain an event sample enriched with particles of a certain pri-
mary species. The net effect of the trigger scheme and of the software cuts is a suppression
of heavier nuclei by a factor larger than 10 at TeV energies. This makes the extraction of
an almost pure proton data sample possible and permits the determination of its energy
spectrum, at least in a narrow range around 1.5 TeV. Even a rather limited knowledge of
the CR chemical composition significantly extends this dynamical range.
In this paper we give a detailed description of the principles underlying a proton measure-
ment. Then we apply the method to data from the HEGRA experiment, that automatically
accumulates CR air shower data in the form of background events during γ-ray observa-
tions. The HEGRA experiment is introduced in Section II, the analysis tools are described
in Section III and the results and the systematic errors are presented in Section IV. Section
V discusses the results. This paper is based on the results of the PhD thesis [7].
II. THE HEGRA IACT-SYSTEM
The HEGRA experiment, located on the Canary Island La Palma at the Observatorio
del Roque de los Muchachos (2200 m a.s.l., 28.75◦ N, 17.89◦ W), is a large detector complex
dedicated to the study of cosmic rays and γ-ray astronomy [8]. In particular, the HEGRA
collaboration operates two air shower arrays on a surface of ≃ 4 · 104 m2. The first one is
an array of 243 scintillation detectors of one square meter area each [9] which samples the
particle cascade reaching the observation level. The other one is the AIROBICC array of
97 wide angle Cherenkov counters [10] which samples the atmospheric Cherenkov photons
emitted by the particle cascade. Apart from γ-ray astronomy in the energy range above
15 TeV, the arrays are used to measure the all particle spectrum and the chemical compo-
sition in the energy range above ≃ 200 TeV [11,12]. The third element in operation is the
4
stereoscopic IACT system together with two IACTs observing in single telescope mode. One
of these telescopes has very recently been incorporated into the stereoscopic system. Here
we concentrate on results obtained by the IACT system.
At the time when the data used in this analysis were taken, the stereoscopic IACT-
System consisted of 4 telescopes with 8.5 m2 mirror area each. Each telescope is equipped
with a 271 pixel camera, covering a field of view of 4.3◦. The pixel size is 0.25◦. The cameras
are readout by an 8 bit 120 MHz Flash-ADC system.
The telescope system uses a multi level trigger scheme [13]. A coincidence of two neigh-
boring pixels above a given threshold triggers an individual telescope. This trigger condition
is called 2NN/271 > q0 hereafter, where NN denotes the next-neighbour condition and q0 is
the threshold in units of registered photoelectrons. A coincidence of at least two telescopes
(named hereafter 2/M-telescope multiplicity, with M=4) triggers the telescope system and
results in the readout of the buffered FADC information of all telescopes.
An absolute calibration of the system has been performed with a laser measurement
and a calibrated low-power photon detector [14]. This measurement has determined the
conversion factor from photons to FADC counts with an accuracy of 12%. The error on the
energy scale is estimated to be 15% which derives from the uncertainty in the conversion
factor from Cherenkov photon counts to FADC counts, and from the uncertainty in the
atmospheric absorption.
To obtain the data used in the following analysis, the photomultipliers are operated in a
regime where saturation effects due to space charges are smaller than 10%, for less than 400
photoelectrons per pixel. A total amplitude of the image, the Size, of 400 photoelectrons
represents an energy of protons of around 15 TeV.
III. ANALYSIS TOOLS
5
A. Monte Carlo Simulations
The CR-induced extensive air showers have been simulated with the ALTAI code [15–17].
The simulation of the electromagnetic shower development models the elementary processes
of bremsstrahlung, ionization losses and Coulomb scattering of charged particles as well as
pair production and Compton scattering of photons. The effect of multiple scattering of
the charged particles is simulated with a fast semi-analytical algorithm which computes the
probability distributions of the lateral and angular distributions of charged particles in a
given volume in space. The simulation of the hadron component is based on accelerator
data of pp- and np-interactions using, where necessary, extrapolations of the cross sections
to TeV energies. The code uses a modified version of the radial-scaling model [18]. Taking
into account the probability coefficients for the different fragmentation channels, the model
of independent nucleon interactions was used to describe the fragmentation of the nuclear
projectile.
In order to study the model dependence of the observable parameters, a second air
shower library was generated, using the CORSIKA code (Version 4.50) [19,20] to simulate the
hadronic interactions of the air shower cascade. CORSIKA offers several interaction models.
High energy interactions (ECM > 80 GeV) were simulated with the HDPM ('hadronic
interactions inspired by the Dual Parton Model') [21]. Low energy interactions (ECM < 80
GeV) were modeled with the GHEISHA code ('Gamma Hadron Electron Interaction Shower
Algorithm'). HDPM is known to describe, also for heavier primary particles, reasonably well
the available accelerator data in the energy region relevant here (1011 < ELab < 1014 eV).
Instead of EGS our variant of CORSIKA uses the ALTAI code to model the electromagnetic
shower development.
A first comparison of the essential characteristics was performed using 5 · 105 showers of
vertical incidence, simulated both with the ALTAI and the CORSIKA hadronic interaction
models in an energy range of 0.3 to 50 TeV and a distance scale of 250 m to the central
telescope of the system. The construction of an energy spectrum relies on the determination
6
of the effective areas. In Figure 1 the effective areas for proton- and helium-induced showers,
as computed with the two interaction models, are compared with each other. The difference
between the two models is smaller than 10% over the full energy range. Although completely
different interaction models have been used, the agreement is excellent. Predicted HEGRA
detection rates have been computed, weighting the individual showers according to the
chemical composition of the nuclei as known from the literature ( [22], see Table I). The
predictions of both models, summarized in Table II, are in very good agreement.
To characterize the telescope images, a 2nd-moment analysis is used to derive the stan-
dard Hillas parameters [23], i.e. the Width-parameter which reflects the lateral development
of the air shower, and the Length-parameter which is related to the longitudinal shower
development. Figure 2 compares the Width and Length parameter distribution as derived
with the two interaction models for proton- and helium-induced air showers. The agreement
is good.
In the following a set of ≈ 106 CORSIKA generated showers in the energy and distance
range given above is used to analyze the data, comprising simulations for air showers induced
by proton and helium as well as by light and medium nuclei (with mass numbers 6 to 19, in
the following abbreviated with LM), and finally by heavy and very heavy nuclei (with mass
numbers 20 to 56, abbreviated as HVH). For the LM- and HVH-groups the atomic numbers
of the primary nuclei were randomly distributed inside the group.
In addition to vertical proton-showers, proton showers incident under z = 20◦ zenith
angles were simulated in order to interpolate the effective area for z ∈ [0◦, 20◦]. The effective
area varies in this range only weakly according to the expected cos z-dependence.
After the air shower simulation, the showers are processed with a new detector simulation
of the HEGRA-System of IACTs. This improved detector simulation includes a full detector
simulation, taking into account Cherenkov photon losses due to atmospheric absorption and
scattering and due to the telescope mirror, the mirror geometry and the arrival times of
the Cherenkov photons, the photomultiplier (PM) response and the characteristics of the
electronic chain to derive the trigger decision and the digitized signal. The new simulations
7
permit an identical treatment of Monte Carlo simulated showers and real data. A detailed
description can be found in [7].
B. Proton Enrichment of the Data Sample
The proton component can effectively be separated from heavier cosmic rays over the
energy range from 1 to more than 10 TeV [6]. The suppression of heavier CRs is based
on the following air shower characteristics: At a given energy, showers induced by heavier
nuclei develop at substantially greater heights in the atmosphere since the cross section σA
for an inelastic hadronic interaction of a primary of nucleon number A with the air nuclei
increases wit A: to first order approximation, σA is given by σg = σ0 · Aα, with σg being the
geometric cross section, σ0 ≈ 30 − 50 mb, and α ≈ 2/3. In addition, the ratio of transverse
momentum to total momentum in the first interaction increases with increasing nucleon
number. Also, the momentum of the primary is for heavy primaries shared among several
nucleons and the typical transverse momentum generated in interactions is fixed, leading to
a larger lateral extension of showers induced by heavy particles. Furthermore, the fraction of
energy channeled into electromagnetic subshowers, responsible for the emission of Cherenkov
photons, decreases with increasing nucleon number [12]. The combination of all these effects
results in a larger but less intensive Cherenkov light pool, increasing the threshold energy
of heavier particles.
A first suppression of the heavier nuclei occurs on the trigger level. Figure 3 shows the
detection rates for different particles, assuming for all nuclei a differential spectrum dF/dE
= 0.25 · E −2.7 s−1 sr−1 m−2 TeV−1. As can be seen, at 1.5 TeV, the energy threshold for
protons, heavier nuclei are suppressed by more than one order of magnitude.
Note that apparently the suppression of heavier nuclei is best at the trigger threshold.
Remarkably a similar suppression occurs de facto also at higher energies by sorting the events
into bins according to their reconstructed energies. Since at a given energy heavier particles
produce a smaller Cherenkov light density, their energy is estimated (see next section) to
8
be smaller by a factor η with η ≈ 3, 5, 6 for Helium, Oxygen and Iron induced air showers,
respectively. In an energy bin centered at the reconstructed energy E, protons of the mean
true energy E are contained, but also heavier particles with the mean true energy η · E.
Since the flux of all primary particles rapidly decreases with increasing energy, to first order
approximation according to dF/dE ∝ E−2.7, heavier particles are suppressed by a factor
η−2.7. This effect is slightly counterbalanced by a relatively larger effective area for heavier
particles at higher energies (> 10 TeV), due to the larger (although less intensive) light pool.
Detailed studies of the separation capabilities at higher energies (> 10 TeV) are still under
way.
A further important suppression of heavier particles is achieved by an analysis of the
stereoscopic IACT images which mirror the longitudinal and lateral shower development,
described by the Hillas-parameters [23]. Pixels with a small S/N-ratio are excluded from
the analysis, by computing the image parameters only from the so called "picture" and
"boundary" pixels [24]. Picture-pixels are all pixels with an amplitude above the "high
tailcut" (here 6 photoelectrons). Boundary pixels are all pixels with an amplitude below the
"high tailcut" but above the "low tailcut" (here 3 photoelectrons) which are neighbours of
a picture-pixel.
Figure 4 shows the distribution of the most important Hillas parameters for data and
for Monte Carlo generated events. Both in the data and the Monte Carlo events, a software
threshold has been applied, requiring two or more telescopes with at least two pixels above
10 photoelectrons, and a sum "Size" of at least 40 photoelectrons recorded in the picture
and boundary pixels. The Monte Carlo events have been weighted according to the chemical
composition from the literature (Table I). The Conc-parameter, measuring the concentration
of the amplitude in the image, is defined as the amplitude in the two most prominent pixels
divided by the total amplitude in the image. Proton images are more concentrated than
images of heavier nuclei. The Distance represents the position of the image centroid in
the camera. Since hadronic showers fall in isotropic, the Distance distribution should rise
linearly until it is cut by the edge of the camera. The agreement between Monte Carlo and
9
data image parameter distributions in Figure 4 is very good.
Note that, since heavier particles are suppressed already on the trigger level, the distri-
butions in Figure 4 depend only slightly on the assumed chemical composition. In future
work we will try to use these small differences to determine the CR chemical composition.
As outlined already in [6] and as seen in Figure 2, the Width parameter, reflecting the
lateral extent of the air shower, is sensitive to the relatively larger transverse momentum
in showers induced by heavier particles and can therefore be used to extract a data sample
enriched with primaries of a certain species. Figure 5 shows the distributions of the Width-
parameter for the different particle groups. The heavier the primary particle, the larger is
the Width parameter.
In the following the parameter mean scaled width, introduced in [26] and first applied
successfully to γ-ray data in [1], is used. For each telescope i the Width-value is normalized
to the value expected for a proton shower h W (Sizei, ri) iMC,p given the sum of photoelectrons
of the image, Sizei, and the distance ri of the shower core from the telescope. The values
obtained from the ntel triggered telescopes are combined to the quantity
Wscal = 1/ntel
ntel
X
i
Wi(Sizei, ri)/h W (Sizei, ri) ip
MC.
(1)
Figure 6 shows the distribution of the Wscal-parameter for the different groups of pri-
maries (assuming a chemical composition as given in Table I). The Wscal-parameter, taking
into account the distance and amplitude dependence of the image width, allows one to en-
hance proton induced showers among showers induced by all particles. The acceptances
of the different cuts are shown in Table III. A cut in Widthscal < 0.85, for example, ac-
cepts ∼ 48% of the primary protons, but only 20% of the primary helium, and ≃ 10% of
the heavier nuclei. The main advantage of scaling the Width-parameter consists in energy
independent cut efficiencies for proton-induced air showers and almost energy independent
cut efficiencies for the heavier primaries, as shown in Figure 7. Since the image widths of
proton-induced showers and showers induced by heavier particles become more similar at
higher energies, the acceptance of heavier nuclei increases slightly with their energy.
10
To summarize, at energies between 1 and 10 TeV the combined effect of suppression of
heavier nuclei by the detection principle and by the image analysis enriches the data sample
with proton-induced showers by a large factor up to one order of magnitude, depending on
the used image shape cuts. In future we shall investigate if additional image parameters can
be used to obtain a similar effective suppression also at energies above 10 TeV.
C. Energy Determination
For each triggered telescope, an energy estimate Ei of the primary particle is computed
under the hypothesis of the primary particle being a proton, from the image Size, Sizei, mea-
sured in the ith telescope at the distance ri of the telescope from the shower core. Averaging
over all triggered telescopes gives a common energy estimate. The energy estimate Ei is de-
termined by inversion of the relation Sizei = h Size(E, r) iMC,p between primary energy E,
impact distance ri and expected image Size Sizei, as computed from the Monte Carlo simula-
tions for proton induced showers. For illustration purposes, the function h Size(E, r) iMC,p is
shown in Figure 8 for 4 broad energy bins. The expected number of photoelectrons decreases
with increasing distance from the shower core. The higher the proton's primary energy is,
the more pronounced is the light concentration near the shower axis. More energetic show-
ers penetrate more deeply into the atmosphere. The tails of these showers give rise to the
increased light intensity near the shower axis in contrast to the flat light pool of primary
photons [27].
This method leads to an energy resolution ∆E/E of ≈ 50% for primary protons, as
shown in Figure 9 for proton induced showers. The energy resolution is determined by the
accuracy of the shower core reconstruction of σri = 30 m and by the variations of the image
size (which is a function of ri and E). Cores which are reconstructed too far away from the
telescopes are partly responsible for the long tail towards large values of ∆E/E. A second
cause are the fluctuations of the image size due to fluctuations in the shower development.
As shown in Table IV the energy resolution ∆E/E as a function of primary proton energy
11
is rather constant, an important requirement for a robust and reliable deconvolution of the
spectrum.
The filled points in Figure 9 show the distribution of ∆E/E for the helium-induced air
showers. As mentioned above, showers induced by heavier nuclei produce, in comparison to
proton-induced air showers of the same energy, a lower Cherenkov light density at observation
level. This effect effectively suppresses, due to the steeply falling spectra of CR primaries,
the contamination of certain energy bins with heavier nuclei. In Figure 10 the differential
detection rates after the cut Widthscal < 0.85 are shown as a function of the reconstructed
energy, assuming the chemical composition from the literature (Table I). As can be seen,
even if the helium to proton ratio would exceed the value given in the literature by a factor
of 2, the contamination of the data sample by heavier particles is small, i.e. < 20%, taking
into account also the cut efficiencies as given in Figure 7.
D. Method to determine the proton spectrum
The proton spectrum is determined using the standard method of forward folding.
The Monte Carlo events of the particle group i are weighted to correspond to a power
law for the flux dF/dE = α · ni E−γi where the ni and the γi (except the γi of the proton
component) reflect the chemical composition taken from the literature ( [22], see Table
I). The fitted parameters are the common scaling parameter α and the spectral index of
the proton component γp. These two parameters are varied until the χ2-difference of the
observed histogram of reconstructed energies and the corresponding histogram predicted
with the weighted Monte Carlo events is minimized. The fit is performed in the range from
1.5 to 3 TeV of the reconstructed energy.
As we have shown in the previous sections, the contamination of the data sample with
heavier particles is small, especially in the energy range from 1 to 3 TeV, and therefore
the result depends only slightly on their assumed abundances and spectral index. This
dependence has been studied in detail and will be discussed in detail below.
12
Flux estimates at given energies are derived as follows: Knowing the best fit value of the
spectral index γp of the proton component, a correction function U(E) is computed from
Monte Carlo simulations so that the differential flux of protons at the reconstructed energy
E can be computed from the number ni of observed events in the ith energy bin by
dF/dE(Ei) = U(Ei) ·
ni
∆t · ∆Ei · κp(Ei) · Aeff(Ei)
,
(2)
where ∆t is the observation time, ∆Ei is the width of the ith energy bin, κp(Ei) is the
acceptance for protons of the Widthscal-cut, and Aeff(Ei) is the effective area for proton
registration. In this ansatz, the effect of the energy resolution and the sample contamination
by heavier particles is accounted for by the function U(E) which depends, for the reasons
mentioned above, only slightly on the assumed chemical composition and on the Widthscal-
cut in the energy range of 1.5 to 3 TeV. Eq. 2 can strictly only be used if the proton
spectrum indeed follows the power law determined in the forward folding fit. However, since
the correction function U(E) depends only weakly on the spectral index, the method gives
reasonable results, also for spectra which deviate from the power law shape.
IV. RESULTS
A. Data Set
For the following analysis, the data primarily taken for the observation of Mkn 501 during
1997 have been used. Only runs taken under excellent weather and hardware conditions were
accepted. Table V gives a summary of the data set.
The Mkn 501 data set was used because of its large fraction of small zenith angle data.
Furthermore, the solid angle region around Mkn 501 does not contain very bright stars
which cause excessive additional noise. As a matter of fact, the strong γ-ray beam from
Mkn 501 in 1997 did not only supply informations of astrophysical interest, but made it
in addition possible to test the simulation of electromagnetic showers and the simulation
of the detector response to these showers with unprecedented statistics (in 1997, 38,000
13
photons were recorded). The strong γ-ray beam could easily be excluded from the analysis
by rejecting all showers reconstructed within 0.3◦ from the source direction.
Identical cuts were applied to the measured data and the Monte Carlo data. In addition
to the cuts already mentioned above, a cut on the distance r of the shower axis from the
central telescope of r < 175 m was applied. Only telescopes with a distance ri smaller
than 200 m from the shower axis entered the analysis, suppressing by these means images
close to the edge of the camera. We apply a mean scaled width cut of Widthscal < 0.85,
which minimizes, to our present understanding, the systematic uncertainties caused by the
contamination of the data sample by heavier particles and by the limited accuracy of the
Monte Carlo simulations.
B. The proton spectrum
The forward folding method described above gives a best power law fit to the data in
the energy range from 1.5 to 3 TeV for:
dF
dE
= (0.11 ± 0.02stat ± 0.05sys) · E −(2.72±0.02stat ±0.15sys) / s sr m2 TeV.
(3)
In Figure 11 the differential energy spectrum is shown assuming the chemical composition
from Table I. This assumption allows to extend the energy range of our measurement to
energies above > 10 TeV, as will be explained later. As can be seen, a single power fits
the data very well. The systematic error on the spectral index is dominated by the Monte
Carlo dependence of the results and by the contamination of the data sample by heavier
particles. The systematic error of the absolute flux is affected in addition by an uncertainty
in the energy scale of 15%. We obtain an integral flux above 1.5 TeV of F (> 1.5 TeV) =
3.1 ± 0.6stat. ± 1.2syst. · 10−2/ s sr m2.
A rough estimate of the systematic errors can be derived by varying the Widthscal-cut.
The different cuts lead to a varying percentage of heavier nuclei in the remaining data
sample. Table VI summarizes the results for cut-values between 1.15 and 0.65. The derived
14
spectral index varies between 2.68 and 2.76 and the flux amplitude (differential flux at 1
TeV) varies between 0.08 and 0.13 /s sr m2 TeV.
We have performed the following studies to estimate the systematic error on the spectral
index.
The dependence of the results on the assumed spectrum for helium, LM- and HVH-
particles was determined in a Monte Carlo study by varying the assumed abundance of the
heavier particles over a wide margin. Setting the assumed flux of one of the groups to zero
or increasing it by a factor of 2, yields the proton spectral indices given in Table VII. Since
Helium has the lowest HEGRA energy threshold of the heavier elements, the spectral index
is most sensitive to the abundance of the Helium component. Setting the assumed Helium
flux to zero results in a proton spectral index of γp = 2.79, and doubling it decreases the
index to 2.73 from an assumed spectral index of γp = 2.75. (see also Figure 12).
Table VIII compares the measured CR detection rates with the rates predicted by weight-
ing the simulated events with the CR spectra of Table I. The rates are in very good agree-
ment. A much higher relative abundance of the heavier particles than assumed in Table I is
therefore not probable. Furthermore, if the Helium abundance would be much larger than
assumed here (more than two times the assumed abundance), the image parameter distribu-
tions found in the data (see Figure 2 and compare with Figure 4) would not fit anymore the
Monte Carlo predictions. Consequently, to our present understanding, the systematic error
in the spectral index due to the uncertainty in the chemical composition is already estimated
conservatively by varying the relative abundances of the heavier elements by factors between
zero and two, and is in the order of 0.05.
Table IX shows the results obtained from experimental data under the extreme hypothesis
of a pure CR proton flux as function of the cut in Widthscal. Comparison with Table VI
shows, that after applying tight cuts (Widthscal < 1.0 or tighter) the results agree nicely
and consequently depend only weakly on the assumed chemical composition.
The dependence of the spectral index on the detector performance and on the atmo-
spheric conditions has been derived as follows. First, the data was divided into 4 parts of
15
equal event statistics, and the analysis was performed for each of the 4 subsamples. Second,
the data was divided into 4 seasonal parts and for each group the spectrum was determined.
The derived spectral indices are given in Table X. They are constant within ∼ 0.05.
The dependence of the spectral index on the details of the Monte Carlo simulation
(mainly threshold effects), has been examined in the framework of determining the system-
atic error on γ-ray spectra measured with the HEGRA IACT system. The studies will be
published elsewhere. The uncertainties on the spectral index are currently estimated to be in
the order of 0.1. The quadratic sum of these systematic errors (dependence on assumed CR
chemical composition 0.05, changing atmospheric and detector conditions 0.05, threshold
effects 0.1) gives a total systematic error on the spectral index of 0.15.
Two effects dominate the systematic error on the flux amplitude.
The uncertainty in the energy scale of 15% [2] translates into an uncertainty of 30% in the
differential flux at a given energy. The uncertainty of the differential flux in the energy range
from 1.5 to 3 TeV from threshold effects is estimated to be 10%, because with increasing
energy the slope of the effective area changes only slowly (compare with Figure 1). Note
that since the energy threshold for heavier particles (Helium to Iron) is much higher than
for protons, the reconstructed proton flux between 1.5 and 3 TeV is essentially independent
of the assumed contamination of the data sample by heavier particles. Figure 12 shows
the reconstructed proton spectrum varying the Helium flux from zero to 3 times the value
from the literature. As can be seen, from 1.5 to 3 TeV, the reconstructed flux is to a good
approximation independent of the assumed Helium flux.
The quadratic sum of the systematic errors (energy scale 30%, threshold effects and cut
efficiencies 15%) gives a total systematic of 35%.
We also investigated wether broken power law models fit our data in the energy range
from 1 to 10 TeV better than single power law spectra. Due to the limited energy resolution
of ∆E/E = 50% for proton induced showers we would be able to detect a break in the 1 to
10 TeV spectrum only for changes in the differential index that are larger than ∼ 1. The
data do not indicate such a break.
16
V. DISCUSSION
In this paper we used a new method to determine the Cosmic Ray proton spectrum in
the energy range from 1.5 to 3 TeV with the stereoscopic IACT system of HEGRA.
As shown in Figure 13, the results are in very good agreement with recent results of
satellite and balloon-borne experiments (see the Figure for references). We have shown that
the new technique yields a similar accuracy as achieved with the present day satellite and
ballon-borne experiments, i.e. an error on the absolute flux of ∼50% and an error on the
spectral index of 0.15.
Earlier claims about a possible cutoff in the proton spectrum at energies below 10 TeV
are clearly not confirmed (e.g. [28,29] and references therein).
Our measurement of the proton spectrum is based on the large effective area of the
atmospheric Cherenkov Technique of ≃ 3·103 m2 sr for a field of view of ≃ 3 msr, and the
stereoscopic imaging technique which permits to reconstruct the protons' primary energy
with the reasonable accuracy of ∆E/E of 50%. The extraction of an almost pure proton
data sample is possible due to a suppression of the number of heavier primaries by more than
one order of magnitude using the multi-telescope trigger and the stereoscopic image analysis.
The accuracy of the measurements is limited by an uncertainty in the energy scale of 15%,
by uncertainties of the detector acceptance, and by a residuum of heavier particles which
could contaminate the data sample, if the relative abundance of heavier particles is much
higher than presently believed. In future work we shall attempt to extend the measurement
of the proton spectrum to higher energies. This might be possible by increasing the software
threshold despite decreasing statistics. Improved cuts should also yield information about
the spectrum of heavier nuclei.
17
ACKNOWLEDGMENTS
The support of the German Ministry for Research and technology BMBF and of the
Spanish Research Council CYCIT is gratefully acknowledged. We thank the Instituto de
Astrofisica de Canarias (IAC) for supplying excellent working conditions at La Palma. We
gratefully acknowledge the work of the technical support staff of Heidelberg, Kiel, Munich
and Yerevan. We thank H. Rebel and D. Muller for fruitful discussions.
18
]
r
s
2
m
[
a
e
r
A
e
v
i
t
c
e
f
f
E
1e+05
1e+04
1e+03
1e+02
Helium, CORSIKA
Proton, CORSIKA
Proton, ALTAI
Helium, ALTAI
1
Energy [TeV]
10
FIG. 1. Comparison of effective areas for the System of HEGRA-Cherenkov-telescopes for pro-
ton- and helium-induced showers, simulated with the ALTAI hadronic interaction model and the
CORSIKA-HDPM code. The differences are smaller than 10%. The trigger required 2NN/271 > 10
ph.e. and a 2/4 telescope coincidence.
19
s
t
i
n
u
0.07
0.06
.
b
r
a
0.05
Proton
CORSIKA
ALTAI
s
t
i
n
u
.
b
r
a
0.06
0.05
0.04
Helium
CORSIKA
ALTAI
0.04
0.03
0.02
0.01
0.03
0.02
0.01
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.9
1
0.8
Width [deg]
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.9
1
0.8
Width [deg]
s
t
i
n
u
.
0.3
b
r
a
0.25
0.2
0.15
0.1
0.05
Proton
CORSIKA
ALTAI
s
t
i
n
u
.
0.3
0.25
b
r
a
0.2
0.15
0.1
0.05
Helium
CORSIKA
ALTAI
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Length [deg]
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.9
0.8
1
Length [deg]
FIG. 2. Comparison of the Width and Length parameter distribution for the proton- and he-
lium-induced showers. The two different interaction models show nearly identical distributions of
these image parameters, reflecting the lateral and longitudinal shower development respectively.
20
]
V
e
T
/
z
H
[
E
d
/
R
d
1
-1
10
-2
10
-3
10
Proton
Helium
LM-particles
HVH-particles
1
10
Energy [TeV]
FIG. 3. Differential detection rates for different nuclei according to individual spectra following
an identical power law. For a single telescope trigger a 2NN/271 > 10 ph.e. condition was applied.
For the System trigger a 2/4 coincidence was required. Already on the trigger level, a clear
suppression of heavier nuclei against protons can be seen. At the energy threshold for protons, this
suppression amounts to at least a factor of 10.
21
s
t
i
n
u
.
b
r
a
0.07
0.06
0.05
0.04
0.03
0.02
0.01
Monte Carlo
Data
s
t
i
n
u
0.3
0.25
.
b
r
a
0.2
0.15
0.1
0.05
Monte Carlo
Data
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.9
1
0.8
Width[deg]
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.9
0.8
1
Length[deg]
0.4
s
t
i
n
u
0.35
.
b
r
a
0.3
0.25
0.2
0.15
0.1
0.05
Monte Carlo
Data
Monte Carlo
Data
0.25
s
t
i
n
u
0.2
.
b
r
a
0.15
0.1
0.05
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Conc
0
0.25
0.5
0.75
1
1.25
1.5
2
2.5
1.75
Distance[deg]
2.25
FIG. 4. Comparison of Monte Carlo and measured image parameters for cosmic rays for an
assumed chemical composition according to the compilation of [22]. A very good agreement between
simulated and measured image parameter distributions can be seen. Remind that identical cuts
where applied for the parameter calculations.
22
Proton
Helium
LM-particle
HVH-particle
s
t
i
n
u
.
b
r
a
0.07
0.06
0.05
0.04
0.03
0.02
0.01
0
0
0.1
0.2
0.3
0.4
0.5
0.6
0.8
0.7
Width [deg]
FIG. 5. The Width-distribution for the particle groups from Table I after the trigger condition
2NN/271 > 10 photoelectrons in each telescope, requiring at least two triggered telescopes in each
event. The distributions are normalized to equal area.
23
Proton
Alpha
LM-particle
HVH-particle
s
t
i
n
u
.
b
r
a
0.05
0.04
0.03
0.02
0.01
0
0
0.5
1
1.5
2
2.5
3
Scaled Width [deg]
FIG. 6. Scaled Width parameter for the different groups of nuclei (assuming a chemical compo-
sition from Table I) as derived from the simulations (normalized to equal area). The same trigger
conditions as in Figure 5 was applied.
24
Proton
Alpha
LM-particles
HVH-particles
y
t
i
l
i
b
a
b
o
r
p
e
c
n
a
t
p
e
c
c
A
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
5
10
15
20
25
35
30
Energy [TeV]
FIG. 7. Acceptance probability as a function of the energy for a cut on the scaled Width
(W < 0.85) for different nuclei. The lines are drawn to guide the eye.
25
]
.
e
.
h
p
[
1000
e
z
i
S
e
g
a
m
I
n
a
e
M
800
600
400
200
0
0
20
40
60
80
100
120
E > 15 TeV
10 < E < 15 TeV
5 < E < 10 TeV
1 < E < 5 TeV
160
200
140
Impact Distance [m]
180
FIG. 8. Dependence of the image amplitude S of the impact distance r for different primary
energies E for proton showers.
26
s
t
n
e
v
e
f
o
r
e
b
m
u
n
e
v
i
t
a
l
e
R
0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
0
-0.5
0
Helium
Helium, 3-5 TeV
Proton, 3-5 TeV
0.5
2
Relative error of energy measurement d E
1
1.5
FIG. 9. Energy resolution of proton induced air showers with an initial energy between 3 and
5 TeV. The distribution is highly asymmetric. For an explanation see text. The given values of
Mean and rms error relate to primary protons.
27
]
V
e
T
/
z
H
[
e
t
a
R
.
f
f
i
0.25
0.2
D
0.15
0.1
0.05
0
1.5
2.5
4.0
6.3
39.8
10.0
Energy (reconstructed) [TeV]
25.1
15.8
FIG. 10. Differential detection rate of proton and the group of helium, LM- and HVH-particles
as function of the reconstructed energy, for a cut on the scaled Width < 0.85, assuming the chemical
composition given in Table I.
28
P1
P2
9.944 / 11
0.1149
-0.2589E-01
0.8474E-02
0.3349E-01
0.3
1
-
)
5
7
1
-
.
V
e
T
2
m
r
s
s
(
.
5
7
2
E
•
E
d
/
F
d
0.2
0.1
0.09
0.08
0.07
0.06
1.5
2.5
4.0
6.3
10.0
15.8
Energy [TeV]
FIG. 11. Differential energy spectrum of protons, obtained using Eq. 2 and assuming the
chemical composition from Table I, multiplied by E2.75. The cut in the scaled Width was
Widthscal. < 0.85. Error bars are statistical only.
29
]
1
-
)
V
e
T
2
m
r
s
s
(
[
E
d
/
F
d
10
10
10
9.944 / 11
0.1149
2.724
Ap
p
0.7574E-02
0.2993E-01
-2
-3
-4
(a
/p)Standard = 0.61
-5
10
-6
10
1.5
/p)Standard
/p < 2(a
0 < a
/p = 2.5(a
/p = 3(a
/p)Standard
/p)Standard
2.5
4.0
6.3
10.0
15.8
Energy [TeV]
FIG. 12. From experimental data reconstructed energy spectrum of protons for a cut on the
scaled Width < 0.85 and an assumed chemical composition according to [22] (black dots). The
hatched area represents the systematics connected with an over-estimation (no helium) and un-
der-estimation (doubled helium content) of the relative proton content. Additional ratios are also
given by the lines.
30
g
a
a
10 4
10 3
10 2
1
-
)
s
u
e
l
c
u
n
/
V
e
T
r
s
s
m
(
x
u
l
2
F
l
a
i
t
n
e
r
e
f
f
i
D
10
1
-1
10
-2
10
-3
10
-4
10
-5
10
-6
10
-7
10
-8
10
-9
10
10
-10
-2
10
-1
10
1
10
31
2
10
10
Energy [ TeV/nucleus ]
3
FIG. 13. Comparison of our proton spectrum measurement with other experiments. The black
points are our measurements around the threshold region of the HEGRA-CT-System. For compar-
ison also indicated are the results of previous satellite and balloon-borne instruments. The shaded
area represents the systematic error of our measurement caused by a variation of the assumed
α/p-ratio over the range 0 < α/p < 3(α/p)Standard relative to (α/p)Standard = 0.61. The shaded
area can be compared to the extreme assumptions of Figure 12.
32
TABLE I. Parameters for the differential energy spectra of different nuclei, taken from [22],
using dF
dE = φ0· E−γ /s sr m2 TeV.
Nucleus
Atomic Number A
Mean Atomic Number hAi
p
1
1
He
4
4
LM
6-19
14
HVH
20-56
40
φ0
γ
0.109±0.32
0.066±0.15
0.028±0.06
0.050±0.19
2.75±0.02
2.62±0.02
2.67±0.02
2.61±0.03
Proportion [at 1 TeV]
0.43
0.26
0.11
0.20
TABLE II. Comparison of the integral rates for ALTAI (RSM) and CORSIKA (HDPM) for the
trigger 2NN/271 > q0 ph.e. and a 2/4 telescope coincidence.
CR primary
p, CORSIKA
p, ALTAI
He, CORSIKA
He, ALTAI
q0, ph.e.
7
10
12
15
20
30
R [Hz]
11.96
6.63
4.78
3.62
2.23
1.22
R [Hz]
R [Hz]
3.65
2.14
1.57
1.21
0.79
0.46
3.74
2.19
1.63
1.28
0.83
0.47
R [Hz]
11.96
6.61
4.68
3.56
2.23
1.22
33
TABLE III. Acceptance probabilities for protons after different scaled Width cuts and the pro-
portion for different nuclei in the residual rate.
Cut, Widthscal [deg]
1.15
1.0
0.85
0.75
0.65
0.55
Acceptance prob. for p
0.854
0.706
0.477
0.293
0.143
0.049
Proton proportion
Helium proportion
LM-proportion
HVH-proportion
0.723
0.215
0.032
0.031
0.762
0.193
0.024
0.021
0.815
0.158
0.016
0.011
0.858
0.124
0.012
0.007
0.889
0.100
0.007
0.004
0.918
0.074
0.005
0.003
TABLE IV. Energy resolution for proton induced air shower.
Energy [TeV]
1.75
2.5
4.0
6.0
8.5
12.5
17.5
25.0
40.0
δE
0.113
0.026
-0.072
-0.118
-0.109
-0.199
-0.295
-0.403
-0.648
Resolution
0.565
0.575
0.535
0.531
0.542
0.516
0.461
0.396
0.251
TABLE V. The data set.
Runs
Period
max. z [deg]
¯z [deg]
t, s
t, h
Events
Events (e.g. Widthscal < 0.85)
79
March-August 1997
20
14.0
191630
53.2
∼ 2 · 106
∼ 6 · 105
34
TABLE VI. Summary of proton spectrum for different scaled Width cuts, according to
dF
dE = Ap · E−γp /s sr m2 TeV.
Widthscal [deg] 1.15
1.0
0.85
0.75
0.65
Ap
γp
0.0829±0.0040 0.0975±0.0052 0.1149±0.0076 0.1206±0.0101 0.1274±0.0149
2.675±0.022
2.709±0.024
2.726±0.030
2.726±0.038
2.758±0.053
TABLE VII. Reconstructed spectral indices of the proton component with no or with doubled
content of heavier particles after according to the standard composition [22] calculated corrections.
The assumed proton spectral index was 2.75 (W < 0.85). This leads to a systematic error of ∼
0.04 due to an incorrectly assumed chemical composition.
Content
Double
No
Helium
2.733
2.793
LM-particles
HVH-particles
2.755
2.762
2.755
2.762
35
TABLE VIII. Comparison of detection rates (given in [Hz]) of the telescope system derived from
Monte Carlo (with an assumed chemical composition after [22]), measurements and data runs. The
trigger condition was always 2 pixel above a threshold q0. NN signifies the next neighbour condition,
MJ the majority decision, which requires only two pixel not necessarily neighboured for the trigger.
The measured values come from [13]. The data values were derived directly from Mkn501 data
runs.
System
Trigger
q0, ph.e. →
2/4
Measurement
Monte Carlo
NN
3/4
Measurement
Monte Carlo
4/4
Measurement
Monte Carlo
2/4
Measurement
Monte Carlo
Data Runs
MJ
3/4
Measurement
Monte Carlo
4/4
Measurement
Monte Carlo
7
16.2
18.1
8.5
9.0
3.8
3.6
18.8
20.8
8.8
10.4
3.9
4.2
10
9.6
10.1
4.7
5.1
1.8
1.9
11.1
11.1
9.1
5.9
5.5
2.5
2.2
12
7.3
7.3
3.7
3.6
1.6
1.3
8.3
7.8
7.7
4.5
3.9
1.9
1.4
15
5.5
5.6
3.0
2.7
1.3
1.0
5.5
5.9
5.9
2.9
2.9
1.1
1.0
20
4.0
3.5
2.1
1.7
0.8
0.6
3.9
3.6
3.9
2.1
1.7
0.8
0.6
30
2.4
2.0
1.2
0.9
0.5
0.3
2.5
2.0
2.2
1.4
0.9
0.6
0.3
TABLE IX. Summary of reconstructed indices from experimental data for different scaled Width
cuts, assuming a pure proton sample in the simulations.
Widthscal [deg] 1.15
1.0
0.85
0.75
0.65
Ap
γp
0.1216±0.0074 0.1322±0.0085 0.1397±0.0107 0.1538±0.0145 0.1552±0.0197
2.647±0.028
2.676±0.030
2.690±0.035
2.727±0.044
2.756±0.059
36
TABLE X. Reconstructed spectral index for different data samples for a scaled Width cut of
∆γp,stat.
± 0.03
± 0.04
0.85.
Sample
Random
γp
Periods
1
13.6 h
2.73
March-May
2
3
12.9 h
13.6 h
2.72
4
13.2 h
2.71
2.73
May
May-July
July-August
Observation time
14.0 h
13.8 h
12.5 h
γp
2.71
2.70
2.72
11.8 h
2.77
37
[1] A. Daum et al., Astroparticle Physics 8, 1 (1997).
[2] F. Aharonian et al., submitted to Astronomy and Astrophysics, 1998.
[3] S. Swordy, Proc. of the 23rd ICRC, Calgary, Invited, Rapporteur and Highlight Papers (1993),
p. 243.
[4] T. Shibata, Proc. of the 24th ICRC, Rome, Invited, Rapporteur and Highlight Papers (1995),
published in Il Nuovo Cimento, Vol. 19 C, N. 5, p. 713, 1996.
[5] B. Wiebel-Sooth, P. L. Biermann, H. Meyer, Astronomy and Astrophysics 330, p. 389-398
(1998).
[6] A. Plyasheshnikov et al., J. Phys. G: Nucl. Part. Phys. 24, 653 (1998).
[7] M. Hemberger, Ph.D. thesis, 1998, Universitat Heidelberg, available as preprint MPI H – V
19 – 1998, MPI fur Kernphysik, Heidelberg.
[8] A. Lindner (HEGRA-Collaboration), in Proc. 25th ICRC, Durban (1997), Vol. 5, p. 113.
[9] H. Krawczynski et al. 1996, NIM A 383, 431.
[10] Karle, A., et al. 1995, Astroparticle Physics 3, 21.
[11] Cortina, J., et al. 1997, in Proc. 25th ICRC, Durban, South Africa, Vol. 4, p. 69
[12] Lindner, A. 1998, Astroparticle Physics 8, 235.
[13] N. Bulian et al., Astroparticle Physics 8, pp. 223–233 (1998).
[14] A. Frass et al., Astroparticle Physics 8, pp. 91–99 (1997).
[15] A. Plyashesnikov, A. Konopelko, preprint Moscow Lebedev Phys. Inst. (1988).
[16] A. Konopelko, A. Plyasheshnikov, A. Schmidt, preprint Moscow Lebedev Phys. Inst. (1992)
38
(in russian).
[17] A. Konopelko, A. Plyasheshnikov, in preparation.
[18] A. M. Hillas, in Proc. 16th ICRC, Kyoto (1979), Vol. 6, p. 13.
[19] J. N. Capdevielle et al., Technical Report No. KfK 4998, Kernforschungszentrum Karlsruhe,
Institut fur Kernphysik (unpublished).
[20] D. Heck et al., CORSIKA: A Monte Carlo Code to Simulate Extensive Air Showers,
Forschungszentrum Karlsruhe, wissenschaftliche Berichte, FZKA 6019.
[21] J. N. Capdevielle et. al., in Proc. VIIth ISVHECRI, Ann Arbor (1992).
[22] B. Wiebel, Wuppertal preprint WUB 94-08 (1994).
[23] A. M. Hillas, in Proc. 19th ICRC, La Jolla (1985), Vol. 3, p. 445.
[24] M. Punch et al., Nature 358, 477 (1992).
[25] P. T. Reynolds et al., Astrophys. J. 404, 206 (1993).
[26] A. Konopelko, in Proc. Towards a Major Atmos. Cherenkov Detector IV, Padova, edited by
M. Cresti (1995), p. 373.
[27] F. Aharonian et al., submitted to Astroparticle Physics, 1998.
[28] V. I. Zatsepin et al., in Proc. 23rd ICRC, Calgary (1993), Vol. 2.
[29] I. P. Ivanenko et al, in Proc. 23rd ICRC, Calgary (1993), Vol. 2.
39
|
0707.3413 | 2 | 0707 | 2007-10-19T13:44:20 | The Sixth Data Release of the Sloan Digital Sky Survey | [
"astro-ph"
] | With the Sixth Data Release of the Sloan Digital Sky Survey, the imaging of the Northern Galactic Cap is now complete. The survey contains images and parameters of roughly 287 million objects over 9583 deg^2, and 1.27 million spectra of stars, galaxies, quasars and blank sky (for sky subtraction) selected over 7425 deg^2. This release includes much more extensive stellar spectroscopy than previously, and also includes detailed estimates of stellar temperatures, gravities, and metallicities. The results of improved photometric calibration are now available, with uncertainties of roughly 1% in g, r, i, and z, and 2% in u, substantially better than the uncertainties in previous data releases. The spectra in this data release have improved wavelength and flux calibration, especially in the extreme blue and extreme red, leading to the qualitatively better determination of stellar types and radial velocities. The spectrophotometric fluxes are now tied to point spread function magnitudes of stars rather than fiber magnitudes, giving a 0.35 mag change in the spectrophotometric flux scale. Systematic errors in the velocity dispersions of galaxies have been fixed, and the results of two independent codes for determining spectral classifications and redshifts are made available. (Abridged) | astro-ph | astro-ph |
Draft version August 14, 2019
Preprint typeset using LATEX style emulateapj v. 08/22/09
THE SIXTH DATA RELEASE OF THE SLOAN DIGITAL SKY SURVEY
Jennifer K. Adelman-McCarthy, Marcel A. Agueros,, Sahar S. Allam,, Carlos Allende Prieto, Kurt S. J.
Anderson,, Scott F. Anderson, James Annis, Neta A. Bahcall, C.A.L. Bailer-Jones, Ivan K. Baldry,, J. C.
Barentine, Bruce A. Bassett,, Andrew C. Becker, Timothy C. Beers, Eric F. Bell, Andreas A. Berlind,
Mariangela Bernardi, Michael R. Blanton, John J. Bochanski, William N. Boroski, Jarle Brinchmann, J.
Brinkmann, Robert J. Brunner, Tam´as Budav´ari, Samuel Carliles, Michael A. Carr, Francisco J. Castander,
David Cinabro, R. J. Cool, Kevin R. Covey, Istv´an Csabai,, Carlos E. Cunha,, James R. A. Davenport, Ben
Dilday,, Mamoru Doi, Daniel J. Eisenstein, Michael L. Evans, Xiaohui Fan, Douglas P. Finkbeiner, Scott D.
Friedman, Joshua A. Frieman,,, Masataka Fukugita, Boris T. Gansicke, Evalyn Gates, Bruce Gillespie, Karl
Glazebrook, Jim Gray, Eva K. Grebel,, James E. Gunn, Vijay K. Gurbani,, Patrick B. Hall, Paul Harding,
Michael Harvanek, Suzanne L. Hawley, Jeffrey Hayes, Timothy M. Heckman, John S. Hendry, Robert B.
Hindsley, Christopher M. Hirata, Craig J. Hogan, David W. Hogg, Joseph B. Hyde, Shin-ichi Ichikawa, Zeljko
Ivezi´c, Sebastian Jester, Jennifer A. Johnson, Anders M. Jorgensen, Mario Juri´c, Stephen M. Kent, R.
Kessler, S. J. Kleinman, G. R. Knapp, Richard G. Kron,, Jurek Krzesinski,, Nikolay Kuropatkin, Donald Q.
Lamb,, Hubert Lampeitl, Svetlana Lebedeva, Young Sun Lee, R. French Leger, S´ebastien L´epine, Marcos
Lima,, Huan Lin, Daniel C. Long, Craig P. Loomis, Jon Loveday, Robert H. Lupton, Olena Malanushenko,
Viktor Malanushenko, Rachel Mandelbaum,, Bruce Margon, John P. Marriner, David Mart´ınez-Delgado,
Takahiko Matsubara, Peregrine M. McGehee, Timothy A. McKay, Avery Meiksin, Heather L. Morrison,
Jeffrey A. Munn, Reiko Nakajima, Eric H. Neilsen, Jr., Heidi Jo Newberg, Robert C. Nichol, Tom Nicinski,,
Maria Nieto-Santisteban, Atsuko Nitta, Sadanori Okamura, Russell Owen, Hiroaki Oyaizu,, Nikhil
Padmanabhan,, Kaike Pan, Changbom Park, John Peoples Jr., Jeffrey R. Pier, Adrian C. Pope, Norbert
Purger, M. Jordan Raddick, Paola Re Fiorentin, Gordon T. Richards, Michael W. Richmond, Adam G. Riess,
Hans-Walter Rix, Constance M. Rockosi, Masao Sako,, David J. Schlegel, Donald P. Schneider, Matthias R.
Schreiber, Axel D. Schwope, Uros Seljak,, Branimir Sesar, Erin Sheldon,, Kazu Shimasaku, Thirupathi
Sivarani, J. Allyn Smith, Stephanie A. Snedden, Matthias Steinmetz, Michael A. Strauss, Mark SubbaRao,,
Yasushi Suto, Alexander S. Szalay, Istv´an Szapudi, Paula Szkody, Max Tegmark, Aniruddha R. Thakar,
Christy A. Tremonti, Douglas L. Tucker, Alan Uomoto, Daniel E. Vanden Berk, Jan Vandenberg, S. Vidrih,
Michael S. Vogeley, Wolfgang Voges, Nicole P. Vogt, Yogesh Wadadekar, David H. Weinberg, Andrew A.
West, Simon D.M. White, Brian C. Wilhite,, Brian Yanny, D. R. Yocum, Donald G. York,, Idit Zehavi, Daniel
B. Zucker
Draft version August 14, 2019
ABSTRACT
This paper describes the Sixth Data Release of the Sloan Digital Sky Survey. With this data
release, the imaging of the Northern Galactic Cap is now complete. The survey contains images
and parameters of roughly 287 million objects over 9583 deg2, including scans over a large range of
Galactic latitudes and longitudes. The survey also includes 1.27 million spectra of stars, galaxies,
quasars and blank sky (for sky subtraction) selected over 7425 deg2. This release includes much
more extensive stellar spectroscopy than previously, and also includes detailed estimates of stellar
temperatures, gravities, and metallicities. The results of improved photometric calibration are now
available, with uncertainties of roughly 1% in g, r, i, and z, and 2% in u, substantially better than the
uncertainties in previous data releases. The spectra in this data release have improved wavelength
and flux calibration, especially in the extreme blue and extreme red, leading to the qualitatively
better determination of stellar types and radial velocities. The spectrophotometric fluxes are now tied
to point spread function magnitudes of stars rather than fiber magnitudes. This gives more robust
results in the presence of seeing variations, but also implies a change in the spectrophotometric scale,
which is now brighter by roughly 0.35 mags. Systematic errors in the velocity dispersions of galaxies
have been fixed, and the results of two independent codes for determining spectral classifications and
redshifts are made available. Additional spectral outputs are made available, including calibrated
spectra from individual 15-minute exposures and the sky spectrum subtracted from each exposure.
We also quantify a recently recognized under-estimation of the brightnesses of galaxies of large angular
extent due to poor sky subtraction; the bias can exceed 0.2 mag for galaxies brighter than r = 14
mag.
Subject headings: Atlases -- Catalogs -- Surveys
1 Fermi National Accelerator Laboratory, P.O. Box 500, Batavia, IL 60510.
2 Columbia Astrophysics Laboratory, 550 West 120th Street, New York, NY 10027.
3 NSF Astronomy and Astrophysics Postdoctoral Fellow.
4 Department of Physics and Astronomy, University of Wyoming, Laramie, WY 82071.
5 McDonald Observatory and Department of Astronomy, The University of Texas, 1 University Station, C1400, Austin, TX 78712-0259.
6 Apache Point Observatory, P.O. Box 59, Sunspot, NM 88349.
2
Adelman-McCarthy et al.
1.
INTRODUCTION
The Sloan Digital Sky Survey (SDSS; York et al. 2000) is a comprehensive imaging and spectroscopic survey of the
optical sky using a dedicated 2.5-meter telescope (Gunn et al. 2006) at Apache Point Observatory in southern New
Mexico. The telescope has a 3◦ diameter field of view, and the imaging uses a drift-scanning camera (Gunn et al.
1998) with 30 2048 × 2048 CCDs at the focal plane which image the sky in five broad filters covering the range from
3000A to 10,000A (Fukugita et al. 1996; Stoughton et al. 2002). The imaging is carried out on moonless and cloudless
nights of good seeing (Hogg et al. 2001), and the resulting images are calibrated photometrically (Tucker et al. 2006;
Padmanabhan et al. 2007) to a series of photometric standards around the sky (Smith et al. 2002). After astrometric
7 Department of Astronomy, MSC 4500, New Mexico State University, P.O. Box 30001, Las Cruces, NM 88003.
8 Department of Astronomy, University of Washington, Box 351580, Seattle, WA 98195.
9 Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544.
10 Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany.
11 Astrophysics Research Institute, Liverpool John Moores University, Twelve Quays House, Egerton Wharf, Birkenhead CH41 1LD,
UK.
12 Center for Astrophysical Sciences, Department of Physics and Astronomy, Johns Hopkins University, 3400 North Charles Street,
Baltimore, MD 21218.
13 South African Astronomical Observatory, Observatory, Cape Town, South Africa.
14 University of Cape Town, Rondebosch, Cape Town, South Africa.
15 Dept. of Physics & Astrophysics, CSCE: Center for the Study of Cosmic Evolution, and JINA: Joint Institute for Nuclear Astrophysics,
Michigan State University, E. Lansing, MI 48824, USA.
16 Center for Cosmology and Particle Physics, Department of Physics, New York University, 4 Washington Place, New York, NY 10003.
17 Department of Physics and Astronomy, University of Pennsylvania, 209 South 33rd Street, Philadelphia, PA 19104.
18 Centro de Astrof´ısica da Universidade do Porto, Rua das Estrelas - 4150-762 Porto, Portugal.
19 Department of Astronomy, University of Illinois, 1002 West Green Street, Urbana, IL 61801.
20 Institut de Ci`encies de l'Espai (IEEC/CSIC), Campus UAB, E-08193 Bellaterra, Barcelona, Spain.
21 Department of Physics and Astronomy, Wayne State University, Detroit, MI 48202.
22 Steward Observatory, 933 North Cherry Avenue, Tucson, AZ 85721.
23 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge MA 02138.
24 Department of Physics of Complex Systems, Eotvos Lor´and University, Pf. 32, H-1518 Budapest, Hungary.
25 Department of Astronomy and Astrophysics, University of Chicago, 5640 South Ellis Avenue, Chicago, IL 60637.
26 Kavli Institute for Cosmological Physics, The University of Chicago, 5640 South Ellis Avenue, Chicago, IL 60637.
27 Department of Physics, University of Chicago, 5640 South Ellis Avenue, Chicago, IL 60637.
28 Institute of Astronomy, Graduate School of Science, The University of Tokyo, 2-21-1 Osawa, Mitaka, 181-0015, Japan.
29 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218.
30 Institute for Cosmic Ray Research, The University of Tokyo, 5-1-5 Kashiwa, Kashiwa City, Chiba 277-8582, Japan.
31 Department of Physics, University of Warwick, Coventry CV4 7AL, United Kingdom.
32 Centre for Astrophysics & Supercomputing, Swinburne University of Technology, P.O. Box 218, Hawthorn, VIC 3122, Australia.
33 Microsoft Research, 455 Market Street, Suite 1690, San Francisco, CA 94105.
34 Astronomical Institute, Department of Physics and Astronomy, University of Basel, Venusstrasse 7, CH-4102 Binningen, Switzerland.
35 Astronomisches Rechen-Institut, Zentrum fur Astronomie, University of Heidelberg, Monchhofstrasse 12-14, D-69120 Heidelberg,
Germany.
36 Bell Laboratories, Alcatel-Lucent, 2701 Lucent Lane, Rm. 9F-546, Lisle, Illinois 60532.
37 Dept. of Physics & Astronomy, York University, 4700 Keele St., Toronto, ON, M3J 1P3, Canada
38 Department of Astronomy, Case Western Reserve University, Cleveland, OH 44106.
39 Lowell Observatory, 1400 W Mars Hill Rd, Flagstaff AZ 86001.
40 Institute for Astronomy and Computational Sciences, Physics Department, Catholic University of America, Washington DC 20064
41 Code 7215, Remote Sensing Division, Naval Research Laboratory, 4555 Overlook Avenue SW, Washington, DC 20392.
42 Institute for Advanced Study, Einstein Drive, Princeton, NJ 08540.
43 National Astronomical Observatory, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan.
44 Department of Astronomy, Ohio State University, 140 West 18th Avenue, Columbus, OH 43210.
45 Electrical Engineering Department, New Mexico Institute of Mining and Technology, 801 Leroy Place, Socorro, NM 87801.
46 Enrico Fermi Institute, University of Chicago, 5640 South Ellis Avenue, Chicago, IL 60637.
47 Gemini Observatory, 670 N. A'ohoku Place, Hilo, HI 96720, USA
48 Obserwatorium Astronomiczne na Suhorze, Akademia Pedogogiczna w Krakowie, ulica Podchor¸azych 2, PL-30-084 Krac´ow, Poland.
49 Department of Astrophysics, American Museum of Natural History, Central Park West at 79th Street, New York, NY 10024
50 Astronomy Centre, University of Sussex, Falmer, Brighton BN1 9QH, UK.
51 Hubble Fellow.
52 Department of Astronomy & Astrophysics, University of California, Santa Cruz, CA 95064.
53 Instituto de Astrofisica de Canarias, La Laguna, Spain.
54 Department of Physics and Astrophysics, Nagoya University, Chikusa, Nagoya 464-8602, Japan.
55 IPAC, MS 220-6, California Institute of Technology, Pasadena, CA 91125.
56 Department of Physics, University of Michigan, 500 East University Avenue, Ann Arbor, MI 48109.
57 SUPA, Institute for Astronomy, Royal Observatory, University of Edinburgh, Blackford Hill, Edinburgh EH9 3HJ, UK.
58 US Naval Observatory, Flagstaff Station, 10391 W. Naval Observatory Road, Flagstaff, AZ 86001-8521.
59 Department of Physics, Applied Physics, and Astronomy, Rensselaer Polytechnic Institute, 110 Eighth Street, Troy, NY 12180.
60 Institute of Cosmology and Gravitation (ICG), Mercantile House, Hampshire Terrace, Univ. of Portsmouth, Portsmouth, PO1 2EG,
UK.
61 CMC Electronics Aurora, 84 N. Dugan Rd. Sugar Grove, IL 60554.
62 Department of Astronomy and Research Center for the Early Universe, Graduate School of Science, The University of Tokyo, 7-3-1
Hongo, Bunkyo, Tokyo 113-0033, Japan.
63 Lawrence Berkeley National Laboratory, One Cyclotron Road, Berkeley, CA 94720.
64 Korea Institute for Advanced Study, 207-43 Cheong-Nyang-Ni, 2 dong, Seoul 130-722, Korea
65 Institute for Astronomy, 2680 Woodlawn Road, Honolulu, HI 96822.
66 Department of Physics, Drexel University, 3141 Chestnut Street, Philadelphia, PA 19104.
SDSS DR6
3
calibration (Pier et al. 2003) the properties of detected objects in the five filters are measured in detail (Lupton et
al. 2001; Stoughton et al. 2002). Subsets of these objects are selected for spectroscopy, including galaxies (Strauss et
al. 2002; Eisenstein et al. 2001), quasars (Richards et al. 2002), and stars. The spectroscopic targets are assigned to
a series of plates containing 640 objects each (Blanton et al. 2003), and spectra are measured using a pair of double
spectrographs, each covering the wavelength range 3800 -- 9200A with a resolution λ/∆λ which varies from 1850 to
2200. These spectra are wavelength- and flux-calibrated, and classifications and redshifts, as well as spectral types for
stars, are determined by a series of software pipelines (Subbarao et al. 2002). The data are then made available both
through an object-oriented database (the Catalog Archive Server, hereafter "CAS"), and as flat data files (the Data
Archive Server, hereafter "DAS").
The SDSS telescope saw first light in May 1998, and entered routine operations in April 2000. We have issued
a series of yearly public data releases, which have been described in accompanying papers (Stoughton et al. 2000,
hereafter the Early Data Release, or EDR paper; Abazajian et al. 2003, 2004, 2005; hereafter the DR1, DR2, and DR3
papers respectively, and Adelman-McCarthy et al. 2006, 2007; hereafter the DR4 and DR5 papers, respectively). The
current paper describes the Sixth Data Release (DR6), which includes data taken through June 2006. Access to the
data themselves may be found on the DR6 website85. This website includes links to both the CAS and DAS websites,
which contain extensive documentation on how to access the data.
When the SDSS started routine operations, the budget funded operations for five years, i.e., through summer
2005. Additional funding from the National Science Foundation, the Alfred P. Sloan Foundation, and the member
institutions secured another three years of operations, and the present data release includes data from the first year
of this extended period, termed SDSS-II. SDSS-II has three components: Legacy, which aims to complete the imaging
and spectroscopy of a contiguous ∼ 7700 deg2 region in the Northern Galactic Cap, SEGUE (Sloan Extension for
Galactic Understanding and Exploration), which is carrying out an additional 3500 deg2 of imaging and spectroscopy
of 240,000 stars to study the structure of our Milky Way, and Supernovae (Frieman et al. 2007), which repeatedly
images a ∼ 300 deg2 equatorial stripe in the Southern Galactic Cap to search for supernovae in the redshift range
0.05 < z < 0.35 for measurement of the redshift-distance relation.
DR6 is cumulative, in the sense that it includes all data that were included in previous data releases. However, as we
describe in detail in this paper, we have incorporated into this data release a number of improvements and additions
to the software. These include:
• Improved photometric calibration, using overlaps between the imaging scans;
• Improved wavelength and flux calibration of the spectra;
• Improved velocity dispersion measurements for galaxies;
• Results of an independent determination of galaxy and quasar redshifts and stellar radial velocities;
• Effective temperatures, surface gravities and metallicities for many stars with spectra.
All DR6 data, including those included in previous releases, have been reprocessed with the new software.
In § 2, the sky coverage of the data included in DR6 is presented. Section 3 describes new features of the imaging data,
including extensive low-latitude imaging, target selection of the SEGUE plates, improved photometric calibration, and
a recently recognized systematic error in sky subtraction which affects the photometry of bright galaxies. Section 4
describes the extensive reprocessing we have done of our spectra, including improved flux and wavelength calibration,
the determination of surface temperatures, metallicities and gravities of stars with spectra, the availability of two
independent determinations of object redshifts, and improved velocity dispersions of galaxies. We summarize DR6 in
§ 5.
67 Department of Physics, Rochester Institute of Technology, 84 Lomb Memorial Drive, Rochester, NY 14623-5603.
68 UCO/Lick Observatory, University of California, Santa Cruz, CA 95064.
69 Kavli Institute for Particle Astrophysics & Cosmology, Stanford University, P.O. Box 20450, MS29, Stanford, CA 94309.
70 Department of Astronomy and Astrophysics, 525 Davey Laboratory, Pennsylvania State University, University Park, PA 16802.
71 Universidad de Valparaiso, Departamento de Fisica y Astronomia, Valparaiso, Chile.
72 Astrophysical Institute Potsdam, An der Sternwarte 16, 14482 Potsdam, Germany.
73 Joseph Henry Laboratories, Princeton University, Princeton, NJ 08544.
74 Institute for Theoretical Physics, University of Zurich, Zurich 8057 Switzerland.
75 Department of Physics and Astronomy, Austin Peay State University, P.O. Box 4608, Clarksville, TN 37040.
76 Adler Planetarium and Astronomy Museum, 1300 Lake Shore Drive, Chicago, IL 60605.
77 Department of Physics and Research Center for the Early Universe, Graduate School of Science, The University of Tokyo, 7-3-1
Hongo, Bunkyo, Tokyo 113-0033, Japan.
78 Dept. of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139.
79 Observatories of the Carnegie Institution of Washington, 813 Santa Barbara Street, Pasadena, CA 91101.
80 Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK.
81 Max-Planck-Institut fur extraterrestrische Physik, Giessenbachstrasse 1, D-85741 Garching, Germany.
82 Astronomy Department, 601 Campbell Hall, University of California, Berkeley, CA 94720-3411.
83 Max Planck Institut fur Astrophysik, Postfach 1, D-85748 Garching, Germany.
84 National Center for Supercomputing Applications, 1205 West Clark Street, Urbana, IL 61801.
85 http://www.sdss.org/dr6
4
Adelman-McCarthy et al.
TABLE 1
Coverage and Contents of DR6
Imaging
Imaging area in CAS
Imaging catalog in CAS
Legacy footprint area
Legacy imaging catalog
SEGUE footprint area, available in DASa
SEGUE footprint area, available in CAS
SEGUE imaging catalog
M31, Perseus scan area
Southern Equatorial Stripe with > 50 repeat scans
Commissioning ("Orion") data
9583 deg2
287 million unique objects
8417 deg2 (5% increment over DR5)
230 million unique objects
1592 deg2
1166 deg2
57 million unique objects
∼ 26 deg2
∼ 300 deg2
832 deg2
Spectroscopy
Spectroscopic footprint area
Legacy
SEGUE
Total number of plate observations (640 fibers each)
Legacy survey plates
SEGUE plates
Special program plates
Repeat observations of plates
Total number of spectra
Galaxiesb
Quasars
Stars
Sky
Unclassifiable
Spectra after removing skies and duplicates
7425 deg2 (20% increment over DR5)
6860 deg2
565 deg2
1987
1520
162
226
79
1,271,680
790,860
103,647
287,071
68,770
21,332
1,115,971
a Includes regions of high stellar density, where the photometry is likely to be poor. See text
for details.
b Spectral classifications from the spectro1d code; numbers include duplicates. The complete
MAIN sample (Strauss et al. 2002) includes 585,719 galaxies after duplicates are removed,
while the luminous red galaxy sample (Eisenstein et al. 2001) contains 79,891 galaxies.
2. THE SKY COVERAGE OF THE SDSS DR6
In the Spring of 2006, the imaging for the SDSS Legacy survey was essentially completed. The Northern Galactic
Cap in DR6 is now contiguous, with the exception of 10 deg2 spread among several holes in the survey; these have
since been imaged, and will be included in the Seventh Data Release. The Northern Galactic Cap imaging survey
covers 7668 deg2 in DR6; the additional Legacy scans in the Southern Galactic Cap bring the total to 8417 deg2. The
sky coverage of the imaging data is shown in Figure 1, and is tabulated in Table 1. The images, spectra, and resulting
catalogs are all available from the DAS; with a few exceptions noted below, all the catalogs are available from the CAS
as well.
The imaging data are the union of three data sets:
• Legacy data, which includes the large contiguous region in the Northern Galactic Cap, as well as three 2.5◦ wide
stripes in the Southern Galactic Cap. These are shown in gray. The lighter gray indicates those regions new to
DR6, containing 417 deg2; the entire Legacy area available in DR6 is 8417 deg2.
• Imaging stripes (also 2.5◦ wide) as part of the SEGUE survey. These do not aim to cover a contiguous area,
but are separated by roughly 20◦ and are designed to sparsely sample the large-scale distribution of stars in
the Galactic halo. These cover just under 1600 deg2, and are all available in the DAS. Notice that many of
these stripes go to quite low Galactic latitude, and some cross the Galactic Plane. As we describe in § 3.1, the
SDSS photometric pipeline is not optimized for crowded fields, and thus the photometry of objects at the lowest
Galactic latitudes is not reliable. Of these data, 1166 deg2 are available in the CAS in a separate database from
the Legacy imaging; these are the regions in which the outputs of the photometric pipeline are most reliable, and
which have been used for spectroscopic targeting (§ 3.2). The SEGUE imaging that is available in both CAS
and DAS is indicated in red; purple indicates the area only available in the DAS.
• Additional imaging taken as part of various auxiliary programs as part of the SDSS, including scans of the region
around M31 and Perseus (see the description in the DR5 paper), and adding up to roughly 26 deg2. These scans
are indicated in blue. These data are not included in the CAS, but are available in the DAS.
In addition, the 2.5◦ wide Equatorial Stripe ("Stripe 82") in the Southern Galactic Cap has been imaged multiple
times through the course of the SDSS, and again as part of the Supernova component of SDSS-II (Frieman et al. 2007).
Sixty-five scans of Stripe 82 observed through Fall 2004 are of survey quality, i.e., they were taken under moonless and
SDSS DR6
5
Fig. 1. -- The distribution on the sky of the data included in DR6 (upper panel:
imaging; lower panel: spectra), shown in an Aitoff
equal-area projection in J2000 Equatorial Coordinates. The Galactic Plane is the sinuous line that goes through each panel. The center
of each panel is at α = 120◦ ≡ 8h, and that the plots cut off at δ = −20◦. The Legacy imaging survey covers the contiguous area of the
Northern Galactic Cap (centered roughly at α = 200◦, δ = 30◦), as well as three stripes (each of width 2.5◦) in the Southern Galactic Cap.
The regions new to DR6 are shown in lighter shading than the rest in both panels. In addition, several stripes (indicated in blue in the
imaging data) are auxiliary imaging data in the vicinity of M31 and the Perseus Cluster, while the SEGUE imaging scans are available
in the DAS and CAS (red) and DAS only (purple). The green scans are additional runs as described in Finkbeiner et al. (2004). In the
spectroscopy panel, special plates (in the sense of the DR4 paper) are indicated in blue, while SEGUE plates are in red. Note that many
plates overlap; for example, there are SEGUE plates in the contiguous area of the Northern Galactic Cap, and the Equatorial Stripe in the
Southern Galactic Cap, which appears solid blue, is also completely covered by the Legacy survey.
cloudless skies in good seeing. As in DR5, we make the calibrated object catalogs and the images corrected for bias,
flatfield, and image defects available through the DAS. There were an additional 171 supernova runs taken in the Fall
seasons of 2005 and 2006. Much of these data were taken under non-photometric conditions, poor seeing, or during
bright moon, and thus the photometry is not reliable at face value (although Ivezi´c et al. 2007 have demonstrated that
it can be calibrated quite well after the fact). The images and the uncalibrated object catalogs for these runs are made
available through the DAS as well. Stripe 82 is composed of two overlapping strips (York et al. 2000), and Figure 2
shows the number of times each right ascension of the two strips is covered in the data through 2004 and as part of
the Supernova survey.
Finkbeiner et al. (2004) made available 470 deg2 of imaging on the Southern Equatorial Stripe taken early in the
survey but not included in either the DAS or the CAS. With DR6, we release an additional 362 deg2 of imaging data;
these runs are indicated in green in Figure 1.
The DR6 spectroscopy contains 1,271,680 spectra over 1987 plate observations. Of these, 1520 plates are from the
main Legacy survey, and there are 64 repeat observations ("extra plates") of 55 distinct Legacy plates. In addition,
6
Adelman-McCarthy et al.
Fig. 2. -- Stripe 82, the Equatorial stripe in the South Galactic Cap, has been imaged multiple times. The lower pair of curves in black
show the number of scans covering a given right ascension in the North and South strip through Fall 2004 (these data were also included in
DR5); these data are available through the DAS. Since that time, Stripe 82 has been covered many more times as part of a comprehensive
survey for 0.05 < z < 0.35 supernovae, although often in conditions of poor seeing, bright moon, and/or clouds; the number of additional
scans at each right ascension in the North and South strip is indicated in red. These latter data have not been flux-calibrated.
there are 234 observations of 226 distinct "special" plates of the various programs described in the DR5 paper86,
indicated in blue in Figure 1, and 169 observations of 162 distinct special plates taken as part of SEGUE (see §3.2)
(indicated in red). In total, these plates cover 7425 deg2 (not including overlaps). Thirty-two fibers (64 fibers for
the SEGUE plates) are dedicated to background sky subtraction on each plate, about 0.7% of spectra are repeat
observations on overlapping plates for quality assurance (and science; see e.g., Wilhite et al. 2005) and roughly 1%
of spectra are too low in signal-to-noise ratio (S/N) for unambiguous classification, so there are roughly 1.1 million
distinct objects with useful spectra in the DR6. This represents a roughly 20% increase over DR5. The areas of sky
new to DR6 are represented in lighter gray in Figure 1. We plan to complete the spectroscopy of the contiguous area
of the Northern Galactic Cap in the Spring of 2008.
The average seeing (see Figure 4 of the DR1 paper) and limiting magnitude of the imaging data, as well as the
typical S/N of the main survey spectra, are essentially unchanged from previous data releases; see the summary of
survey characteristics in Table 1 of the DR5 paper.
The SDSS photometric processing pipeline has been stable since DR2, and thus the quantities measured for all
objects included in DR5 have been copied wholesale into DR6. This version of the pipeline has been used for the small
3. CHARACTERIZATION AND USE OF THE IMAGING DATA
86 An updated special-plate list is at http://www.sdss.org/dr6/products/spectra/special.html .
SDSS DR6
7
amount of Northern Galactic Cap data new to DR6, as well as the SEGUE imaging scans shown in Figure 1. The
magnitudes quoted in the SDSS archives are asinh magnitudes (Lupton et al. 1999).
3.1. SEGUE data at low Galactic latitudes
The SEGUE imaging survey is designed to explore the structure of the Milky Way at both high and low Galactic
latitudes, and thus extends to lower latitudes than did the Legacy survey. This extension gives us better leverage on
the spatial distribution of stars in the disk components of the Milky Way, and on the three-dimensional shape of the
stellar halo. Eighty-six of the 162 SEGUE plates were targeted off SEGUE imaging, while the remainder were targeted
off Legacy imaging. The SEGUE imaging scans are made available in a separate database, termed "SEGUEDR6",
within the CAS.
The SEGUE imaging data close to the Galactic plane have regions of higher dust extinction and object density
than does the high-latitude SDSS. The SDSS imaging reduction pipelines used to reduce the data for DR6 were not
designed for optimal performance in crowded fields, and are known to fail for some of these data. In particular:
• When the images are sufficiently crowded, the code has trouble finding suitable isolated stars from which to mea-
sure the point spread function (PSF). Without a suitable determination of the PSF, the brightness measurements
by the pipeline (Stoughton et al. 2002) are inaccurate.
• The pipeline attempts to deblend objects with overlapping images, but the deblend algorithm fails when the
number of overlapping objects gets too large, such as happens in sufficiently crowded fields. In such fields, the
number of detected objects reported by the pipeline can be a dramatic underestimate.
• At low latitudes, the dust causing Galactic extinction (as measured by Schlegel, Finkbeiner & Davis 1998,
hereafter SFD) cannot be assumed to lie completely in front of the stars in the sample. This has an effect on
the interpretation of quality assurance tools based on the position of the stellar locus, as we describe below.
Therefore, it is necessary to check that the quality of the reductions in any area of the sky of interest is adequate to
address a particular science application of the data.
As Ivezi´c et al. (2004) and the DR3 paper explain, we use a series of automated quality checks on the imaging data
to determine whether the data meet our science requirements; the results of these tests are made available in the CAS.
These checks are available for the SEGUE imaging as well. The best indicator of bad PSF photometry is the difference
between PSF and large aperture magnitudes for stars brighter than 19th magnitude. If the median difference between
the two is greater than 0.03 mag, the PSF photometry will not make the survey requirement of 2% calibration error
in g, r, or i. About 2.3% of the fields of the SEGUE imaging data loaded into the CAS in DR6 fail this criterion87.
For comparison, about 1.6% of all fields in the SDSS Legacy footprint in DR6 fail this criterion.
The automated overall measurement of the quality in a given field also takes into account the location of the stellar
locus in the ugr and gri color-color diagrams, and how it differs in each field from the average value over the entire
survey (see the discussion in Ivezi´c et al. 2004). These color-color diagrams are made with SFD extinction-corrected
magnitudes, so even for very good photometry they may vary from the survey average if that extinction correction
is not valid for any reason. The user should apply appropriate caution in interpretation of the stellar locus location
diagnostics in the quality assurance for these data.
Finally, the photometric pipeline performs poorly for a stellar density greater than ∼ 5000 objects brighter than the
detection limit per 10′ × 13′ field, or about 140, 000 objects deg−2, a density roughly ten times the density at high
latitudes. The outputs of the photometric pipeline are quite incomplete (and indeed, confusingly, can fall well below
5000 objects per field) and can be unreliable for more crowded fields. Almost all the SEGUE data affected by this
problem are in the DAS only; the SEGUE imaging in the CAS (which is the subset used for SEGUE target selection;
see below) largely avoids these crowded areas of the sky.
3.2. SEGUE target selection
SEGUE has as one of its goals a kinematic and stellar population study of the high-latitude thick disk and halo of
the Milky Way. The halo is sampled sparsely with a series of tiles each of seven deg2 in both the SEGUE imaging
stripes and the main Legacy survey area, with centers separated by roughly 10 deg. Each such tile is sampled with
two pointings, one plate for stars brighter than r = 17.8 (approximately the median target magnitude), and one
plate, which typically gets double the standard exposure time, for fainter stars. The target selection categories and
criteria are summarized in Table 2 (listed roughly in order from bluest to reddest targets); see the DR4 paper for a
description of an earlier version of SEGUE targeting. Most of the target selection categories are sparsely sampled,
with a sampling rate that depends on magnitude; see the on-line documentation for more details. The target selection
bits in the PrimTarget flag are indicated in the table. Spectra with target selection bits set by the SEGUE target
selection algorithm have PrimTarget bit 0x80000000 and SecTarget bit 0x40000000 set.
Half of the science targets on each line of sight are selected using color-color and color-magnitude cuts designed to
sample at varying densities across the main sequence from g − r = 0.75 (K dwarfs at Teff < 5000K). To this sample we
add metal-poor main sequence turnoff stars selected by their blue ugr colors, essentially an ultraviolet excess cut that
87 Of course, a much larger fraction of the additional SEGUE data available in the DAS also fail this criterion.
8
Adelman-McCarthy et al.
TABLE 2
SEGUE targeting algorithms
Category
Bit (Hex)
Color cuts
White dwarf
A, BHB stars
Metal-poor MS turnoff
F/G stars
G stars
Cool white dwarf
0x80080000
0x80020000
0x80100000
0x80000200
0x80040000
0x80020000
g < 20.3, −1 < g − r < −0.2, −1 < u− g < 0.7, u− g + 2(g − r) < −0.1
g < 20.5, 0.8 < u − g < 1.5, −0.5 < g − r < 0.2
g < 20.3, −0.7 < P1 < −0.25, 0.4 < u − g < 1.4, 0.2 < g − r < 3.0
14.0 < g < 20.2, 0.2 < g − r < 0.48
14.0 < r < 20.2, 0.48 < g − r < 0.55
14.5 < r < 20.5, −2 < g − i, Hg > max[17.5, 16.05 + 2.9(g − i)],
Low metallicity
K giant
K dwarf
MS/WD pairs
0x80010000
0x80040000
0x80008000
0x80001000
1.7
otherwise
g − i < 0.12 neighbor with g < 22 within 7′′
r < 19.5, −0.5 < g − r < 0.75, 0.6 < u − g < 3.0, l > 0.135
r < 20.2, 0.7 < u − g < 4.0, 0.5 < g − r < 0.9, 0.15 < r − i < 0.6,
l > 0.07, µ < 0.011′′/yr
14.5 < r < 19.0, 0.55 < g − r < 0.75
15 < g < 20, u − g < 2.25, −0.2 < g − r < 1.2, 0.5 < r − i < 2.0,
−19.78(r − i) + 11.13 < g − r < 0.95(r − i) + 0.5,
i − z >
0.68(r − i) − 0.18 otherwise
0.5 if r − i > 1.0
M subdwarf
High µ M subdwarf
Brown dwarf
AGB
0x80400000
0x80400000
0x80200000
0x80800000
14.5 < r < 19.0, g − r > 1.6, 0.95 < r − i < 1.3
µ > 0.04′′/yr, r − z > 1.0, 15 + 3.5(g − i) > Hr > 12 + 3.5(r − z)
z < 19.5, u > 21, g > 22, r > 21, i − z > 1.7
14.0 < r < 19.0, 2.5 < u − g < 3.5, 0.9 < g − r < 1.3, s < −0.06
#/tile
25
≤155
200
50
375
10
150
95
95
5-10
5
60
<5
10
Note. -- This table describes Version 4 2 of the SEGUE target selection algorithm. The hex bit in the second column
is set in the PrimTarget flag. All magnitudes above are PSF magnitudes which have been corrected for Galactic extinction
following SFD. The one exception is the MS/WD pair algorithm, which uses PSF magnitudes without extinction correction.
The quantity l ≡ −0.436u + 1.129g − 0.119r − 0.574i + 0.1984 is a metallicity indicator following Lenz et al. (1998). The
quantities s ≡ −0.249u + 0.794g − 0.555r + 0.234 and P1 ≡ 0.91(u − g) + 0.415(g − r) − 1.280 are defined by Helmi et al.
(2003). The proper motion µ is in units of arcsec/yr, and the reduced proper motion is defined as Hg ≡ g + 5 log µ + 5 and
similarly for Hr. The fourth column lists the typical number of targets selected in each category per spectroscopic tile.
is highly efficient at separating the halo from the thick disk near the turnoff. At the faint end, r = 19.5, the average
star that makes this selection is at a heliocentric distance of 10 kpc for [Fe/H] = −1.54. To reach to greater distances,
we use the strength of the Balmer jump to select field blue horizontal branch (BHB) stars in the ugr color-color
diagram (Lenz et al. 1998, Sirko et al. 2004; Clewley et al. 2004). The halo BHB sample extends to distances of 40
kpc at g = 19 (corresponding to the S/N limit we use for detailed spectroscopic classification; see § 4.3). We select all
available BHB candidates in our high-latitude fields, and all candidates with g − r < 0 irrespective of latitude.
We select distant halo red giant candidates by the photometric offset in the ugr color-color diagram, as quantified by
the l color (Lenz et al. 1998; see the notes to Table 2). This offset is caused by their ultraviolet excess and weak Mg Ib
and MgH at at 5175A relative to foreground disk dwarfs (Morrison et al. 2001, Helmi et al. 2003). This is augmented
by a 3σ proper motion cut using a recalibrated version of the USNO-B catalog (Munn et al. 2004). Spectroscopic
identification of true giants using the methodology in Morrison et al. (2003) has shown that the giant selection is
roughly 50% efficient at g < 17, the current limit to which we can reliably distinguish giants from dwarfs in the
spectra. The halo giant sample identified in this way reaches distances of 40 kpc from the Sun. We select candidate
low-metallicity stars using a more extreme l-color cut, and without any proper motion cut.
The spectroscopic selection also includes smaller categories of rare but interesting objects. These include cool white
dwarfs selected with the recalibrated USNO-B reduced proper motion diagram, which can be used to date the age of
the Galactic disk (Gates et al. 2004; Harris et al. 2006), high proper motion targets from the SUPERBLINK catalog
(L´epine & Shara 2005), which have uncovered some of the most extreme M subdwarfs known (L´epine et al. 2007)
and have aided in the calibration of their metallicity scale using common proper motion pairs, and white dwarf/main
sequence binaries containing cool white dwarfs, which are predicted to be the dominant population among this type of
binaries (Schreiber & Gansicke 2003). These rare object categories also include color-only selections for cool subdwarfs,
brown dwarfs (using cuts similar to those employed by Chiu et al. 2006), and the SEGUE "AGB" category that selects
metal-rich, cool giants that separate readily from the ugr stellar locus.
Table 2 describes Version 4 2 of the SEGUE target selection algorithms. The algorithms have evolved throughout
the survey, and users wishing to understand the detailed selection associated with each target category should examine
the SEGUE documentation off the DR6 survey page. The user should also know that SEGUE target selection has
been run only on those chunks used to design SEGUE plates, and has not yet been run on the bulk of the Legacy
survey imaging.
As described in the DR2 paper, the proper motions of stars in the SDSS are taken from the measurements of the
USNO-B1.0 (Monet et al. 2003), based primarily on the POSS-I and POSS-II. However, this catalog is incomplete
3.3. A caveat on high proper motion stars
SDSS DR6
9
at the highest proper motions, greater than 100 milli-arcsec per year. Confusingly, objects with no proper motion
measurement in the USNO-B1.0 catalog have their proper motion listed as 0.0 in the CAS ProperMotions table,
meaning that a query for low proper motion stars will be contaminated by a small number of the highest proper
motion stars. The best available catalog of high proper motion stars can be found in the SUPERBLINK catalog of
L´epine & Shara (2005) and references therein; we hope to incorporate this catalog into the proper motion data in the
SDSS in future data releases.
3.4. Low Galactic latitude SDSS commissioning data
During commissioning and subsequent tests of the SDSS observing system, additional data were obtained outside of
the nominal survey region. These data consist of 28 runs (see Finkbeiner et al. 2004, Table 1) at low Galactic latitude,
mostly in the star-forming regions of Orion, Cygnus, and Taurus. There are 832 deg2 of data, 470 deg2 of which have
been previously released88 as flat files. There are three types of files: calibrated images (one calibImage per field),
calibrated object files (one calibObj per field), and condensed "sweep" files (one star or galaxy file per run/camcol).
The remaining 362 deg2 are hereby released in the same format, but they are not available in the DAS or CAS.
These data have been photometrically calibrated using the ubercalibration algorithm (§ 3.5)89. Ubercalibration takes
advantage of the Apache Wheel calibration scans (not shown in Figure 1) to tie the photometry of disjoint regions
of the sky together; nevertheless, because the overlap with other runs is less than in the main survey area, their
calibration may not be quite as good.
3.5. Improved photometric calibration
Photometric calibration in SDSS has been carried out in two parallel approaches. The first uses an auxiliary 20′′
photometric telescope (PT) at the site, which continuously surveys a series of US Naval Observatory standard stars
which are used to define the SDSS u′g ′r′i′z ′ photometric system (Smith et al. 2002). Transformations between the
u′g ′r′i′z ′ and native SDSS 2.5-meter ugriz photometric systems and zeropoints for stars in patches surveyed by the
2.5-meter telescope are determined with these data (Tucker et al. 2006, Davenport et al. 2007). These secondary
patches are spaced roughly every 15◦ along the imaging stripes. This approach has allowed the SDSS photometry to
reach its goals of calibration errors with an rms of 2% in g, r, and i, and 3% in u and z (Ivezi´c et al. 2004), as measured
from repeat scans (see the discussion in Ivezi´c et al. 2007). This is the calibration process that has been used in all
data releases to date. However, it is not ideal for several reasons:
• The u′g ′r′i′z ′ filter system of the PT camera is subtly different from the ugriz system on the 2.5-meter;
• There are persistent problems with the flat-fielding of both the PT and 2.5-meter cameras, especially in u′;
• No use is made of overlap data in the 2.5-meter scans to tie the zeropoints together.
A second approach, termed "ubercalibration" (Padmanabhan et al. 2007) does not use information from the PT
to calibrate individual runs, but rather uses the overlaps between the 2.5-meter imaging runs to tie the photometric
zeropoints of individual runs together and measure the 2.5m flatfields, and to determine the extinction coefficients on
each night. Unlike the standard PT calibrations, ubercalibration explicitly assumes that the photometric calibration
parameters -- a zeropoint for each CCD, and atmospheric extinction linear with airmass -- are constant through a
photometric night.This assumption appears justified, as the resulting calibration has errors of ∼ 1% in g, r, i and
z, and 2% in u, roughly a factor of two below those of the standard processing, as determined from the overlaps
themselves, and from the measurement of the "principal colors" of the stellar locus (see the discussion in Ivezi´c et al.
2004 and the DR3 paper). This scatter is dominated by unmodelled variations in the atmospheric conditions in the
site, including changes in the atmospheric extinction through a night.
The relative calibration of the photometric scans via overlaps does not determine the photometric zeropoints in the
five filters. The zeropoints are constrained in practice by forcing the ubercalibrated photometry of bright stars to
agree in the mean with that calibrated in the standard way (Tucker et al. 2006). Thus this work does not represent an
improvement in the calibration of the SDSS photometry to a true AB system (in which magnitudes can be translated
directly into physical flux units); see the discussion in the DR2 paper, Eisenstein et al. (2006), and Holberg & Bergeron
(2006). Moreover, there are subtle differences between the response of the six filters in each row of the SDSS camera,
especially in z (see the discussion in Ivezi´c et al. 2007); these differences have not been corrected.
Both versions of SDSS photometry are now made available through the CAS in DR6. The PT-calibrated photometry
for each detected object is stored in the database in the same tables and columns as in DR5, and both the offset between
PT and ubercalibration, as well as the ubercalibrated magnitudes, are stored in the UberCal table of the CAS. Database
functions are available to apply these offsets and output ubercalibrated photometry. The distribution of these offsets
is shown in Figures 15 and 16 of Padmanabhan et al. (2007); the improvements are subtle, changing magnitudes of
most individual objects by 0.02 mag or less.
88 At http://photo.astro.princeton.edu .
89 The current ubercalibration has yielded calibrations typically 0.02 mag different from those previously released, but some runs/camera
columns show differences as large as 0.05 mag. The variance within each field is also somewhat reduced by correcting flatfield errors at the
0.01 or 0.02 mag level.
10
Adelman-McCarthy et al.
3.6. The photometry of bright galaxies
Because of scattered light (see the EDR paper), the background sky in the SDSS images is non-uniform on arc-
minute scales. The photometric pipeline determines the median sky value within each 101.4′′ (256 pixel) square on a
grid with 50.7′′ spacing, and bilinearly interpolates this sky value to each pixel. This procedure overestimates the sky
near large extended galaxies and bright stars, and as was already reported in the DR4 paper and Mandelbaum et al.
(2005), causes a systematic decrease in the number density of faint objects near bright galaxies. In addition, it also
strongly affects the photometry of bright galaxies themselves, as has been reported by Lauer et al. (2007), Bernardi
et al. (2007), and Lisker et al. (2007). We have quantified this effect by adding simulated galaxies with exponential
and de Vaucouleurs (1948) profiles to SDSS images, following Blanton et al. (2005a). The simulated galaxies ranged
from apparent magnitude mr = 12 to mr = 19 in half-magnitude steps, with a one-to-one mapping from mr to S´ersic
half-light radius determined using the mean observed relation between these quantities for MAIN sample galaxies
(Strauss et al. 2002) with exponential and de Vaucouleurs profiles. Axis ratios of 0.5 and 1 were used, with random
position angles for the non-circular simulated galaxies. The results in the r band are shown in Figure 3, plotting the
difference between the input magnitude and the model magnitude returned by the SDSS photometric pipeline as a
function of magnitude. Also shown is the fractional error in the scale size re. The biases are significant to r = 16
for late-type galaxies, and to r = 17.5 for early-type galaxies. Hyde & Bernardi (unpublished) fit de Vaucouleurs
models to SDSS images of extended elliptical galaxies, using their own sky subtraction algorithm, which is less likely
to overestimate the sky level near extended sources. Their results, also shown in the figure, are quite consistent with
the simulations.
The scatter in the offset from one realization to another is large enough that we cannot recommend a deterministic
correction for this problem. This scatter depends in part on the position of the simulated galaxy relative to the grid
on which the sky interpolation occurs. We are working on an improved algorithm which will fit the extended profiles
of galaxies explicitly as part of the sky determination, and hope to include the results in a future data release.
4. SPECTROSCOPY
The Sixth Data Release contains a number of improvements and additions to the SDSS spectroscopy. These include
an improved pipeline to extract and calibrate the one-dimensional spectra (§ 4.1), the results of an independent pipeline
to classify objects and measure redshifts (§ 4.2), the results of a pipeline to determine the effective temperatures,
gravities and metallicities of stars (§ 4.3), and improvements to the existing code to measure velocity dispersions
(§ 4.4).
4.1. The extraction and calibration of one-dimensional spectra
The pipeline that extracts, combines, and calibrates the SDSS spectra of individual objects from the two-dimensional
spectrograms ("idlspec2d") was originally designed to obtain meaningful redshifts for galaxies and quasars. However,
there were several ways in which the code was inadequate, especially in light of the stellar focus of the SEGUE
project, and the recognition of the rich stellar data available among the spectra of the main SDSS survey. The
spectrophotometry was tied to the fiber magnitudes of stars, whose relation to the true, PSF magnitudes of stars is
seeing-dependent. In addition, the SEGUE spectroscopy includes "bright plates" which contain substantial numbers of
stars as bright as if iber = 14.2, and scattered light from these stars caused systematic errors in the sky subtraction on
these plates. Finally, there were errors in the wavelength calibration as large as 15 km s−1 on some plates, acceptable
for most extragalactic science, but a real limitation for SEGUE's science goals. These concerns and others have caused
us to substantially revise and improve the idlspec2d pipeline; the results of this improvement are included in DR6.
4.1.1. Spectrophotometry: Flux Scale
The new code has a different spectrophotometric calibration flux scale. The fiber magnitude reported by the
photometric pipeline is the brightness of each object, as measured through a 3′′ diameter aperture corrected to 2′′
seeing to match the entrance aperture of the fibers (see the discussion in the EDR paper). However, the relationship
between the fiber magnitudes of stars and the PSF magnitudes (which, for unresolved objects, is our best determination
of a true, total magnitude) is dependent on seeing; this is made worse because the colors of stars measured via fiber
magnitudes will be sensitive to the different seeing in the different filters (although cases in which the seeing is
dramatically different in the different bands are fairly rare). With this in mind, the pipeline used in DR6 determines
the spectrophotometric calibration on each plate such that the flux of the spectrum of standard stars integrated over
the filter curve matches the PSF magnitude of the stars as measured from their imaging. This calibration is determined
for each of the four cameras (two in each spectrograph) from observations of standard stars. Additional corrections
to handle large-scale astrometric and chromatic terms are measured from isolated stars and galaxies of high S/N, and
are then applied to all the objects on the plate.
The results of this calibration may be seen in Figure 4, which compares synthesized magnitudes from the SDSS
spectra with the PSF and fiber magnitudes in the imaging data, showing results both from the old ("DR5") and new
("DR6") codes. We emphasize that the calibration is not tied to the PSF photometry of each object individually
(otherwise the comparison in Figure 4 would be a tautology); there is a single calibration determined for each camera
in a given plate. This means, for example, that it is meaningful to compare photometry and spectrophotometry of
individual objects to look for variability (e.g., Vanden Berk et al. 2004).
SDSS DR6
11
Fig. 3. -- The effects of sky subtraction errors on the photometry of bright galaxies. Upper panel: The error in the r band model
magnitude of simulated galaxies with an n = 1 (exponential) profile (blue hexagons) and an n = 4 (de Vaucouleurs) profile (red crosses) as
determined by the photometric pipeline, as a function of magnitude. Fifteen galaxies are simulated at each magnitude for each profile. Also
shown are the analogous results from Hyde & Bernardi (unpublished) for three early-type galaxy samples: 54 nearby (z < 0.03) early-type
galaxies from the ENEAR catalog (da Costa et al. 2000) in black; 280 brightest cluster galaxies from the C4 catalog (Miller et al. 2005) in
green; and 9000 early-type galaxies from the Bernardi et al. (2003a) analysis in magenta. Lower panel: The fractional error in the scale
size re as a function of magnitude from the simulations and the Hyde & Bernardi analysis.
The PSF includes light that extends beyond the 3′′ diameter of the filters, and thus the PSF-calibrated spectropho-
tometry is systematically brighter than the old fiber-calibrated photometry by the difference between PSF and fiber
magnitudes, which is roughly 0.35 magnitudes (albeit dependent on seeing). Again, because the PSF photometry
represents an accurate measure of the brightness of stars, this calibration means that the spectrophotometry matches
the PSF photometry for stars to an rms of 4%. This distribution does show an extended tail presumably caused by
blended and variable objects90, but the distribution is substantially more symmetric than for the previous version of
the pipeline. Interestingly, for galaxies, the rms difference between spectroscopic photometry and the fiber magnitudes
is also 4%. The previous code shows a similarly narrow distribution, albeit with larger tails. The distribution of the
difference of the g − r and r − i colors between PSF photometry and as synthesized from the spectrophotometry again
shows a narrow core in both DR5 and DR6, but again with less extensive non-Gaussian outliers with the new code.
Due to errors in the processing step, there are 28 plates, listed in Table 3, that were calibrated using fiber magnitudes
rather than PSF magnitudes. Therefore, objects on these plates have a spectroscopic flux scale systematically lower
by 0.35 mag than the rest of the survey. These will be processed correctly in a subsequent data release.
90 Indeed, the fiber magnitudes include light from overlapping blended objects, thus the tails are less extensive in the fiber magnitude
comparison.
12
Adelman-McCarthy et al.
Fig. 4. -- The distribution of differences between r-band photometry synthesized from SDSS spectra (labelled "SPECTRO"), and PSF
and fiber magnitudes, for stars and galaxies; results are shown for DR6 (left-hand panel) and the previous version of the calibration available
in DR5 (right-hand panel). Only objects with PSF magnitude brighter than 19 are shown. The most important difference is the offset
of 0.35 magnitudes between the two, due to the change in calibration from fiber to PSF photometry. Each panel includes the mean and
standard deviation of the best-fit Gaussian, as well as the number of objects lying beyond 3σ (as a measure of the non-Gaussianity of the
tails). Results are shown for r band, but g and i band results are very similar.
4.1.2. Spectrophotometry: Wavelength Dependence
13
SDSS DR6
TABLE 3
Spectroscopic plates calibrated with fiber magnitudes
plate MJD plate MJD plate MJD plate MJD
269
270
277
284
309
324
336
51910
51909
51908
51943
51666
51666
51999
345
349
353
367
394
403
446
51690
51699
51703
51997
51913
51871
51899
460
492
543
554
556
616
616
51924
51955
52017
52000
51991
52374
52442
683
730
830
872
1394
1414
1453
52524
52466
52293
52339
53108
53135
53084
Note. -- The second column lists the Modified Julian Date
(MJD) on which each plate was observed.
As discussed in the DR2 paper, each plate includes observations of a number of spectrophotometric standards,
typically F subdwarfs. Their observed spectra are fit to and calibrated against the models of Gray & Corbally (1994),
as updated by Gray et al. (2001). We can compare the spectrophotometric calibration between DR5 and DR6 by
plotting the ratio of the summed spectra of these standard stars on each plate as determined by the two versions of
the pipeline. The 0.35 mag overall flux scale between the two calibrations has been taken out by forcing all the curves
through unity at 6200A. The median ratio (as determined from 1278 plates), and the 68.3% and 95.4% outliers, are
shown in Figure 5. The median ratio differs from unity by less than 5% at all wavelengths, but a small fraction of the
plates have differences as large as 30% at the far blue end.
Do these changes represent an improvement scientifically? Figure 4 of the DR2 paper quantified the uncertainties in
the spectrophotometric calibration used at that time by looking at the mean fractional offset between observed spectra
of white dwarfs and best-fit models for them. Figure 6 shows a similar analysis with the old and new reductions.
The curves show the median fractional difference between a sample of 128,000 calibrated luminous red galaxy (LRG,
Eisenstein et al. 2001) spectra, and a model based on averaged observed LRG spectra that is allowed to evolve smoothly
with redshift (see the discussion in § 3 of the DR5 paper). Because the LRGs have a broad range of redshifts, one
expects no feature specific to the LRGs to appear in this plot as a function of observed wavelength, and deviations
from unity are a measure of the small-scale errors in the spectrophotometry. There are systematic oscillations at the
2% level in the DR5 reductions. These wiggles correspond to positions of strong absorption lines in the standard stars,
especially in the vicinity of the 4000A break in the blue. This is now handled by not fitting the instrumental response
to any residual non-telluric features finer than 25-50A, as the response is not expected to vary on those scales. This
reduces the amplitude of the wiggles by a factor of two in the DR6 reductions, especially at the blue end. Redward
of 4500A, 50% of the spectra fall within 3% of the median value; this increases to 7% at 3800A. The features at Ca
K and H (3534 and 3560A) and Na D (5890 and 5896A) are probably due to absorption from the interstellar medium
(although the latter probably has a contribution from sky line residuals). The sky line residuals (marked with the ⊕
symbol) are a function of S/N; a similar analysis with higher S/N quasars shows substantially smaller residuals at the
strong sky lines.
The effect of this improvement in the spectrophotometric calibration becomes clear if we examine the spectra of
individual stars. Figure 7 shows the blue part of the spectrum of an A0 blue horizontal branch star as calibrated with
the old code (dotted) and the new (solid), together with a synthetic spectrum based on the atmospheric parameters
estimated by the SEGUE Stellar Parameter Pipeline (§ 4.3; Teff = 8446 K, log g = 3.15, [Fe/H] = −1.96). The new
reductions are clearly smoother between the absorption lines; the match between the DR6 calibrated spectrum and
the synthetic spectrum is also superior.
4.1.3. Radial velocities
In order to measure the dynamics of the halo of the Milky Way, SEGUE requires stellar radial velocities accurate
to 10 km s−1, significantly more demanding than the original SDSS requirements of 30 km s−1. The previous version
of idlspec2d had systematic errors of 10 -- 15 km s−1 in the wavelength calibration because of a dearth of strong lines
at the blue end of the spectrum in the calibration lamps and in the nighttime sky. The sky-line fits for the blue side
wavelength corrections now use a more robust algorithm allowing less freedom in the fits, and these problems are
largely under control.
We monitor the systematic and random errors in the radial velocities in the SEGUE data by comparing repeat
observations on the bright and faint plates of each SEGUE pointing. The duplicate observations consist of roughly
20 "quality assurance" objects selected at the median magnitude of the SEGUE data, as well as a similar number
of spectroscopic calibration objects that are observed on both plates. The mean difference in the measured radial
velocities between the two observations of the quality assurance objects depends on stellar type, with a standard
deviation of 9 km s−1 for A and F stars and 5 km s−1 for K stars91. The mean radial velocity offset between the
two plates in each pointing, as measured using all the duplicate observations, suggests systematic velocity errors from
plate to plate of only 2 km s−1 rms.
91 Thus the error on a single star is √2 less than these values.
14
Adelman-McCarthy et al.
Fig. 5. -- The ratio of the summed spectra of standard stars on each plate as determined by the DR6 and DR5 versions of spectrophotom-
etry, rescaled to unity at 6200A. The solid line is the median ratio spectrum over 1278 plates, the dotted lines enclose 68.3% of the plates
(corresponding to 1σ for a Gaussian distribution), and the dashed lines enclose 95.4% of the plates (corresponding to 2σ). The distribution
at each wavelength is in fact close to Gaussian.
We have checked the zeropoint of the overall radial velocity scale (as measured using the ELODIE templates in
the specBS code; see the discussion below in § 4.2) by carrying out high-resolution observations of 150 SEGUE stars.
This has revealed a systematic error of 7.3 km s−1 (in the sense that the SpecBS velocities are too low) due to subtly
different algorithms in the line fits to arc and sky lines. This has been fixed in the output files of the SSPP (§ 4.3
below), but has not yet been fixed elsewhere in the CAS.
The improved wavelength calibration leads to smaller sky subtraction residuals for many objects, especially noticeable
in the far red of the spectrum.
4.1.4. Additional outputs
Under good conditions, a typical spectroscopic plate is observed three times in exposures of 15 minutes each; more
exposures are added in poor conditions to reach a target S/N in the spectra. The idlspec2d pipeline stitches together
the resulting individual spectra to determine the final spectrum of a given object. However, for the most accurate
determination of the noise characteristics of the spectra (for example, in detailed analyses of the Lyman α forest of
quasars; see the discussion in McDonald et al. 2006), or to determine whether a specific unusual feature in a spectrum
is real, it is desirable to go back to the uncombined spectra. These uncombined spectra are now made available for
every plate in the so-called spPlate files through the DAS.
The published spectra have had a determination of the spectrum of the foreground sky subtracted from them. The
sky is measured in 32 fibers (64 fibers for the faint SEGUE plates) placed in regions where no object has been detected
SDSS DR6
15
Fig. 6. -- The median ratio of observed flux-calibrated spectra of luminous red elliptical galaxies to their averaged spectra (after taking
evolution into account), for the previous (DR5) and current (DR6) spectroscopic reductions. This quantifies the wavelength dependence
of systematic errors in the spectrophotometric calibration; the amplitude of these features, already small in the previous reductions, have
been reduced further in DR6, especially in the blue. The features at Ca H and K and at Na D are probably due to absorption from the
interstellar medium. The strong features at the sky lines at 5577A and 4358A marked with the ⊕ symbol are related to the S/N of the
spectra; a similar analysis with quasar spectra shows these features to have substantially lower amplitude.
to 5σ in the imaging data, interpolated (both in amplitude and in wavelength, allowing for some undersampling) to
each object exposure, and subtracted. However, it is often useful to see the sky spectrum that has been subtracted from
each object, for example to study the nature of extended foreground emission-line objects in the data (for example,
see Hewett et al. 2003 for the discovery of a 2◦ diameter planetary nebula in the SDSS data). The sky spectrum
subtracted from each object spectrum is now available through both the DAS and the CAS.
4.1.5. The treatment of objects with very strong emission lines
There is a known problem, which is not fixed with the current version of idlspec2d, whereby the code that combines
the individual 15-minute exposures will occasionally mis-interpret the peaks of particularly strong and narrow emission
lines as cosmic rays and remove them. All pixels affected by this have the inverse variance (i.e., the inverse square of
the estimated error at this pixel) set to zero, indicating that the code recognizes that the pixel in question is not valid.
A diagnostic of this problem is unphysical line ratios in the spectra of dwarf starburst galaxies, as the tops of the
strongest lines are artificially clipped. This is a rare problem, affecting less than 1% of galaxies with rest equivalent
width in the Hβ line greater than 25A, but users investigating the properties of galaxies with strong emission lines
should be aware of it. We hope to fix this problem in the next data release.
16
Adelman-McCarthy et al.
Fig. 7. -- The blue part of the spectrum of an A0 blue horizontal branch star, SDSS J004037.41+240906.5, as given by the old (red dotted
curve) and new (black solid curve) versions of idlspec2d. The old curve has been scaled up to reflect the difference in the calibration
of the two reductions. The synthetic spectrum, shown in green, is generated from a model with parameters matching those derived from
the SSPP (Teff = 8500 K, log g = 3.25, [Fe/H] = −2.00). The continuum between the absorption lines is much smoother, and matches
the synthetic spectrum much better for the new reductions than for the old. The synthetic spectrum has been normalized to match the
observed spectrum at 4500 A. Neither the model nor the spectra have been corrected for Galactic reddening (which is E(B − V ) = 0.036
in this line of sight).
4.2. An independent determination of spectral classifications and redshifts
As described in the EDR paper and Subbarao et al. (2002), the spectral classifications and radial velocities available
in the data releases have been based on a code (spectro1d), that cross-correlates the observed spectra with a variety
of templates in Fourier space to determine absorption-line redshifts and fits Gaussians to emission lines to determine
emission-line redshifts. A completely separate code, termed specBS and written by D. Schlegel (in preparation) instead
carries out χ2 fits of the spectra to templates in wavelength space (in the spirit of Glazebrook et al. 1998), allowing
galaxy and quasar spectra to be fit with linear combinations of eigenspectra and low-order polynomials. Stellar radial
velocities are fit both to SDSS-derived stellar templates, and to templates drawn from the high-resolution ELODIE
(Prugniel & Soubiran 2001) library. The spectro1d outputs give the default spectroscopic information available
through the CAS, but the specBS outputs are made available through the CAS for the first time with DR692. While
spectro1d uses manual inspection to correct the redshifts and classifications of a small fraction of its redshifts, specBS
is completely automated.
92 The outputs of specBS have also been made publically available through the NYU Value-Added Galaxy Catalogue; see Blanton et al.
(2005b).
SDSS DR6
17
Redshift warning flags from specBS
TABLE 4
Bit
Name
Comments
0
1
2
3
4
5
6
SKY FIBER
SMALL LAMBDA COVERAGE Because of masked pixels, less than half of the full wavelength range is reliable
Fiber is used to determine sky; there should be no object here.
CHI2 CLOSE
NEGATIVE TEMPLATE
MANY 5SIGMA
CHI2 AT EDGE
NEGATIVE EMLINE
in this spectrum.
The second best-fitting template had a reduced χ2 within 0.01 of the best fit
(common in low S/N spectra).
Synthetic spectrum is negative (only set for stars and QSOs).
More than 5% of pixels lie more than 5 σ from the best-fit template.
χ2 is minimized at the edge of the redshift-fitting region (in this circumstance,
Z ERR is set to −1).
A quasar emission line (C IV, C III], Mg II, Hβ, or Hα) appears in absorption
with more than 3 σ significance due to negative eigenspectra.
Outputs of the specBS pipeline made available in the DR6 CAS.
TABLE 5
Parameter
CLASS
SUBCLASS
Z
Z ERR
RCHI2
DOF
VDISP
VDISP ERR
ZWARNING
ELODIE SPTYPE
ELODIE Z
ELODIE Z ERR
Comments
STAR, GALAXY, or QSO
Stellar subtype, galaxy type (starforming, etc)
Heliocentric redshift
Error in redshift
Value of reduced χ2 for template fit to spectrum
Degrees of freedom in χ2 fit
Velocity Dispersion for galaxies (km s−1)
Error in Velocity Dispersion (km s−1)
Set if the classification or redshift are uncertain; see Table 4
Spectral type of best-fit ELODIE template
Redshift determined from best-fit ELODIE template
Error in redshift determined from best-fit ELODIE template
Tests show that the two pipelines give impressively consistent results. At high S/N, the rms difference between
the redshifts of the two pipelines is of order 7 km s−1 for stars and galaxies, although the spectro1d redshifts are
systematically higher by 12 km s−1 due to differences in the templates. The difference distribution has non-Gaussian
tails, but as a test of catastrophic errors, we find that 98% of all objects with spectra (after excluding the blank sky
fibers) have consistent classification (star, quasar, galaxy) and redshifts agreeing within 300 km s−1 for galaxies and
stars, and 3000 km s−1 for quasars.
Half of the remaining 2% are objects of very low S/N, and the other half are a mixture of a variety of unusual
objects, including BL Lacertae objects (Collinge et al. 2005; their lack of spectral features makes it unsurprising that
the two pipelines come to different conclusions), unusual white dwarfs, including strong magnetic objects and metal-
rich systems (Schmidt et al. 2003; Eisenstein et al. 2006; Dufour et al. 2007), unusual broad absorption line quasars
(Hall et al. 2002), superposed objects, including at least one gravitational lens (Johnston et al. 2003), and so on. Both
pipelines set flags when the classifications or redshifts are uncertain (see Table 4); the majority of these discrepant
cases are flagged as uncertain by both pipelines.
Table 5 lists the outputs from the specBS pipeline included in the CAS for each object.
In addition, the DAS
includes the results of the cross-correlation of each of the templates with each spectrum, as well as Gaussian fits to
the emission lines. These quantities are included in the SSPP table (§ 4.3) in the CAS, and as flat files in the DAS.
4.3. The measurement of stellar atmospheric parameters from the spectra
The SEGUE science goals require accurate determinations of effective temperature, Teff, surface gravity (log g, where
g is in cgs units, cm s−2), and metallicity [Fe/H]), for the stars with spectra (and ugriz photometry) obtained by
SDSS. We have developed the SEGUE Stellar Parameter Pipeline (SSPP), to determine these quantities and measure
77 atomic and molecular line indices for each object. The code and its performance is described in detail by Lee et al.
(2007a). Validation of the sets of parameters based on Galactic open and globular clusters and with high-resolution
spectroscopy obtained for over 150 SDSS/SEGUE stars is discussed by Lee et al. (2007b) and Allende Prieto et al.
(2007). Due to the wide range of parameter space covered by the stars that are observed, a variety of techniques are
used to estimate the atmospheric parameters; a decision tree is implemented to decide which methods or combination
of methods provide optimal measures, based on the colors of the stars and S/N of the spectra.
These methods include:
• Fits of the spectra to synthetic photometry and continuum-corrected spectra based on Kurucz (1993) model
atmospheres (Allende Prieto et al. 2006), or to synthetic spectra computed with the more recent Castelli &
18
Adelman-McCarthy et al.
Kurucz (2003) models (Lee et al. 2007a);
• Measurements of the equivalent widths of various metal-sensitive lines, including the Ca II K line (Beers et al.
1999) and the Ca II infrared triplet (Cenarro et al. 2001);
• Measurements of the equivalent widths of various gravity-sensitive lines such as Ca I λ4227A and the 5175A Mg
Ib/MgH complex (e.g., Morrison et al. 2003);
• Measurements of the autocorrelation function of the spectrum, which is useful for high-metallicity stars (Beers
et al. 1999);
• A neural network technique which takes the observed spectrum as input, trained on previously available param-
eters from the SSPP (Re Fiorentin et al. 2007).
For stars with temperatures between 4500 K and 7500 K and with average S/N per spectral pixel greater than 15,
the typical formal errors returned by the code are σ(Teff ) = 150 K, σ(log g) = 0.25 dex, and σ([Fe/H]) = 0.20 dex.
Comparison with 150 stars with high S/N high resolution spectra (and therefore reliable stellar parameters) validates
these error estimates, at least for those stars with the highest quality SDSS spectra.
The SSPP assumes solar abundance ratios when quoting metallicities, with the caveat that several of the individual
techniques (those that involve the Ca and Mg line strengths) adopt a smoothly increasing [α/Fe] ratio, from 0.0 to
+0.4, as inferred metallicity decreases from solar to [Fe/H] = −1.5. Other techniques, which are based on regions of
the spectra dominated by lines from unresolved Fe-peak elements, do not assume such relationships.
The S/N limit for acceptable estimated stellar parameters varies with each individual method employed by the SSPP.
As a general rule, the SSPP sets a conservative criterion that the average S/N per pixel over the wavelength range
3800-6000A must be greater than 15 for stars with g − r < 0.3, and greater than 10 for stars with g − r ≥ 0.3. Stars
of low S/N do not have their parameters reported by SSPP. Table 5 of Lee et al. (2007a) describes the valid ranges of
effective temperature, g − r color, and S/N for each method used in the SSPP.
The SSPP values are combined with the outputs of specBS (§ 4.2) and are loaded as a single table into the CAS,
with entries for every object with a spectrum.
For the coolest stars, measuring precise values of Teff, log g, and [Fe/H] from spectra dominated by broad molecular
features becomes extremely difficult (e.g., Woolf & Wallerstein 2006). As a result, the SEGUE SSPP does not estimate
atmospheric parameters for stars with Teff < 4500 K, but instead estimates the MK spectral type of each star using
the Hammer spectral typing software developed and described by Covey et al. (2007)93. The Hammer code measures 28
spectral indices, including atomic lines (H, Ca I, Ca II, Na I, Mg I, Fe I, Rb, Cs) and molecular bandheads (G band,
CaH, TiO, VO, CrH) as well as two broad-band color ratios. The best-fit spectral type of each target is assigned by
comparison to the grid of indices measured from more than 1000 spectral type standards derived from spectral libraries
of comparable resolution and coverage (Allen & Strom 1995; Prugniel & Soubiran 2001; Hawley et al. 2002; Bagnulo
et al. 2003; Le Borgne et al. 2003; Valdes et al. 2004; S´anchez-Bl´azquez et al. 2006). These indices, and the best-fit
type from the Hammer code, are available for stars of type F0 and cooler in DR6.
Tests of the accuracy of the Hammer code with degraded (S/N ∼ 5) STELIB (Le Borgne et al. 2003), MILES
(S´anchez-Bl´azquez et al. 2006) and SDSS (Hawley et al. 2002) dwarf template spectra reveal that the Hammer code
assigns spectral types accurate to within ±2 subtypes for K and M stars. The Hammer code can return results for
warmer stars, but as the index set is optimized for cool stars, typical uncertainties are ±4 subtypes for A -- G stars at
S/N ∼ 5; in this temperature regime, SSPP atmospheric parameters are a more reliable indicator of Teff.
Given SEGUE's science goals, we emphasize two limitations to the accuracy of spectral types derived by the Hammer
code:
• The Hammer code uses spectral indices derived from dwarf standards; spectral types assigned to giant stars will
likely have larger, and systematic, uncertainties.
• The Hammer code was developed in the context of SDSS' high latitude spectroscopic program; the use of broad-
band color ratios in the index set will likely make the spectral types estimated by the Hammer code particularly
sensitive to reddening. Spectral types derived in areas of high extinction (i.e., low-latitude SEGUE plates) should
be considered highly uncertain until verified with reddening-insensitive spectral indices.
4.4. Correction of biases in the velocity dispersions
Both specBS and spectro1d measure velocity dispersions (σ) for galaxies. specBS does so, as described above,
by including it as a term in the direct χ2 fit of templates to galaxies. The velocity dispersion in spectro1d was
computed as the average of the Fourier- and direct-fitting methods (Appendix B of Bernardi et al. 2003b; hereafter
B03). However, due to changes in the spectroscopic reductions from the EDR to later releases, a bias appeared in the
recent values available in the CAS. As shown in Appendix A of Bernardi (2007), σ values in the DR5 do not match the
values used by B03. The difference is small but systematic, with spectro1d DR5 larger than B03 at σ ≤ 150 km s−1.
93 The
Hammer
code
has
been made
available
for
community
use:
the
IDL code
can
be downloaded
from
http://www.cfa.harvard.edu/∼kcovey/thehammer .
SDSS DR6
19
Fig. 8. -- Top panels: velocity dispersion measurements from B03 (left), DR5 (middle) and specBS (right) versus the spectro1d DR6
values for the sample of elliptical galaxies used in Bernardi et al. (2003a). Bottom panels: The ratio of DR6 values to the other three
samples (i.e. B03, DR5, and specBS) versus the mean value (e.g. left panel hσi = (σDR6 + σB03)/2). The median value at at each value of
hσi is shown as a solid line; the values including 68% and 95% of the points are given as the dashed and dotted lines, respectively.
A similar bias is seen when comparing spectro1d DR5 with measurements from the literature (using the HyperLeda
database; Paturel et al. 2003). Simulations similar to those in B03 show that the discrepancy results from the fact
that the Fourier-fitting method is biased by ∼ 15% at low dispersions (∼ 100 km s−1), whereas the direct-fitting
method is not. We therefore use only the direct-fitting method in DR6. Figure 8 shows comparisons of the spectro1d
DR6 velocity dispersions with those from B03, DR5 and the specBS measurements. There is good agreement between
spectro1d DR6 and B03 (rms scatter ∼ 7.5%), and between spectro1d DR6 and specBS (rms scatter ∼ 6.5%),
whereas spectro1d DR5 is clearly biased high at σ ≤ 150 km s−1. The agreement between spectro1d DR6 and
specBS is not surprising, since both are now based only on the direct-fitting method. The specBS measurements tend
to be slightly smaller than DR6 at σ ≤ 100 km s−1; specBS is similarly smaller than HyperLeda, whereas DR6 agrees
with HyperLeda at these low dispersions.
Figure 9 shows the distribution of the error on the measured velocity dispersions. The direct-fitting method used
by spectro1d gives slightly larger errors than does the Fourier-fitting method, peaking at ∼ 10%. The figure shows
that this error distribution is consistent with that found by comparing the velocity dispersions of ∼ 300 objects from
DR2 which had been observed more than once.
Finally, HyperLeda reports substantially larger velocity dispersions than SDSS at σ ≥ 250 km s−1. The excellent
agreement between three methods (direct fitting, cross-correlation, and Fourier-fitting) applied to the SDSS spectra at
the high velocity dispersion end gives us confidence in our velocity dispersions (Bernardi 2007), although it is unclear
why the literature values are higher.
4.5. Linking SEGUE imaging and spectroscopy
For the Legacy imaging, there exist simple links between the spectroscopic and imaging data, but these links are not
yet in place between all the SEGUE spectroscopy and imaging. In particular, to obtain BEST or ubercalibrated stellar
photometry of SEGUE spectroscopic objects within CAS, one must perform an "SQL join" command between the
spectroscopic specobjall or sppParams tables in the CAS with the corresponding photometric tables (photoobjall,
seguedr6.photoobjall, or ubercal). Sample queries on how to do this are provided on the SDSS web site. We plan
to simplify this procedure in future data releases.
5. CONCLUSIONS AND THE FUTURE
We have presented the Sixth Data Release of the Sloan Digital Sky Survey. It includes 9583 deg2 of imaging data,
including a contiguous area of 7668 deg2 of the Northern Galactic Cap. The data release includes almost 1.3 million
spectra selected over 7425 deg2 of sky, representing a 20% increment over the previous data release. This data release
includes the first year of data from the SDSS-II, and thus includes extensive low-latitude imaging data, and a great
deal of stellar spectroscopy. New to this data release are:
• 1592 deg2 of imaging data at lower Galactic latitudes, as part of the SEGUE survey, of which 1166 deg2 are in
searchable catalogs in the CAS;
• Revised photometric calibration for the imaging data, with uncertainties of 1% in g, r, i and z, and 2% in u;
• Improved wavelength and flux calibration of spectra;
20
Adelman-McCarthy et al.
Fig. 9. -- Error distribution of the velocity dispersion measurements from spectro1d DR6 (thin black solid line), spectro1d DR5 (dotted
red line), specBS (dashed blue line), and B03 (dotted-dashed green line). The thick solid line was obtained by comparing the velocity
dispersions of ∼ 300 galaxies which had been observed two or more times; it is thus an empirical estimate of the true error.
• Detailed surface temperatures, metallicities, and gravities for stars.
The SDSS-II will end operations in Summer 2008, at which point the Legacy project will have completed spectroscopy
for the entire contiguous area of the Northern Cap region now covered by imaging, and SEGUE will have obtained
spectra for 240,000 stars. The supernova survey (Frieman et al. 2007) has discovered 327 spectroscopically confirmed
SNIa to date in its first two seasons, and has one more season to go.
Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating In-
stitutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space
Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council
for England. The SDSS Web Site is http://www.sdss.org/.
The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participat-
ing Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel,
University of Cambridge, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the
Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear
Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese
Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA),
the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, Ohio State University, University of
Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University
of Washington.
Abazajian, K. et al. 2003, AJ, 126, 2018 (DR1 paper)
Abazajian, K. et al. 2004, AJ, 128, 502 (DR2 paper)
REFERENCES
SDSS DR6
21
Abazajian, K. et al. 2005, AJ, 129, 1755 (DR3 paper)
Adelman-McCarthy, J. K. et al. 2006, ApJS, 162, 38 (DR4 paper)
Adelman-McCarthy, J. K. et al. 2007, ApJS, 172, 634 (DR5
paper)
Allen, L. E., & Strom, K. M. 1995, AJ, 109, 1379
Allende Prieto, C. et al. 2006, ApJ, 636, 804
Allende Prieto, C. et al. 2007, in preparation
Bagnulo, S., Jehin, E., Ledoux, C., Cabanac, R., Melo, C.,
Gilmozzi, R., & the ESO Paranal Science Operations Team
2003, The Messenger, 114, 10
Beers, T. C., Rossi, S., Norris, J. E., Ryan, S. G., & Shefler, T.
1999, ApJ, 506, 892
Bernardi, M. 2007, AJ, 133, 1954
Bernardi, M. et al. 2003a, AJ, 125, 1817
Bernardi, M. et al. 2003b, AJ, 125, 1882 (B03)
Bernardi, M., Hyde, J. B., Sheth, R. K., Miller, C. J., & Nichol,
R. C. 2007, AJ, 133, 1741
Blanton, M.R., Eisenstein, D., Hogg, D.W., Schlegel, D.J., &
Brinkmann, J, 2005a, ApJ, 629, 143
Blanton, M.R. et al. 2005b, AJ, 129, 2562
Blanton, M.R., Lin, H., Lupton, R.H., Maley, F.M., Young, N.,
Zehavi, I., & Loveday, J. 2003, AJ, 125, 2276
Castelli, F. & Kurucz, R. L. 2003, IAU Symposium, 210, A20
Cenarro, A. J., Gorgas J., Cardiel N., Pedraz S., Peletier R.F., &
Vazdekis, A. 2001, MNRAS, 326, 981
Chiu, K., Fan, X., Leggett, S.K., Golimowski, D.A., Zheng, W.,
Geballe, T.R., Schneider, D.P., & Brinkmann, J. 2006, AJ, 131,
2722
Clewley, L., Warren, S.J., Hewett, P.C., Norris, J.E., & Evans,
N.W. 2004, MNRAS, 352, 285
Collinge, M. et al. 2005, AJ, 129, 2542
Covey, K.R. et al. 2007, AJ, in press (arXiv:0707.4473v2)
da Costa, L. N., Bernardi, M., Alonso, M. V., Wegner, G.,
Willmer, C. N. A., Pellegrini, P. S., Rit´e, C., Maia, M. A. G.,
2000, 120, 95
Davenport, J.R.A., Bochanski, J.J., Covey, K.R., Hawley, S.L., &
West, A.A. 2007, preprint
de Vaucouleurs, G. 1948, Annales d'Astrophysique, 11, 247
Dufour, P. et al. 2007, ApJ, 663, 1291
Eisenstein, D.J. et al. 2001, AJ, 122, 2267
Eisenstein, D.J. et al. 2006, AJ, 132, 676
Finkbeiner, D.P. et al. 2004, AJ, 128, 2577
Frieman, J. et al. 2007, AJ, submitted (arXiv:0708.2749)
Fukugita, M., Ichikawa, T., Gunn, J.E., Doi, M., Shimasaku, K.,
& Schneider, D.P. 1996, AJ, 111, 1748
Gates, E. et al. 2004, ApJ, 612, L129
Glazebrook, K., Offer, A.R., & Deeley, K. 1998, ApJ, 492, 98
Gray, R.O. & Corbally, C.J. 1994, AJ, 107, 742
Gray, R.O., Graham, P.W., & Hoyt, S.R. 2001, AJ, 121, 2159
Gunn, J.E. et al. 1998, AJ, 116, 3040
Gunn, J.E. et al. 2006, AJ, 131, 2332
Hall, P.B. et al. 2002, ApJS, 141, 267
Harris, H.C. et al. 2006, AJ, 131, 571
Hawley, S. L. et al. 2002, AJ, 123, 3409
Helmi, A. et al. 2003, ApJ, 586, 195
Hewett, P.C., Irwin, M.J., Skillman, E.D., Foltz, C.B., Willis,
J.P., Warren, S.J., & Walton, N.A. 2003, ApJ, 599, L37
Hogg, D.W., Finkbeiner, D.P., Schlegel, D.J., & Gunn, J.E. 2001,
AJ, 122, 2129
Holberg, J.B. & Bergeron, P. 2006, AJ, 132, 1221
Holtzman, J. et al. 2007, in preparation
Ivezi´c, Z. et al. 2004, Astronomische Nachrichten, 325, 583
Ivezi´c, Z. et al. 2007, AJ, 134, 973
Johnston, D.E. et al. 2003, AJ, 126, 2281
Kurucz, R. L. 1993, private communication
Lauer, T.R. et al. 2007, ApJ, 662, 808
Le Borgne, J.-F., et al. 2003, A&A, 402, 433
Lee, Y.S. et al. 2007a, preprint
Lee, Y.S. et al. 2007b, preprint
Lenz, D., Newberg, H., Rosner, R., Richards, G.T., & Stoughton,
C. 1998, ApJS, 119, 121
L´epine, S, Rich, R.M., Shara, M.M., Cruz, K.L., & Skemer, A.
2007, ApJ, 668, 507
L´epine, S. & Shara, M.M. 2005, AJ, 129, 1483
Lisker, T., Grebel, E.K., Binggeli, B., & Glatt, K. 2007, ApJ, 660,
1186
Lupton, R.H., Gunn, J.E., Ivezi´c, Z., Knapp, G.R., Kent, S., &
Yasuda, N. 2001, in Astronomical Data Analysis Software and
Systems X, edited by F. R. Harnden Jr., F. A. Primini, and H.
E. Payne, ASP Conference Proceedings, 238, 269
Lupton, R.H., Gunn, J.E., & Szalay, A.S. 1999, AJ, 118, 1406
Mandelbaum, R. et al. 2005, MNRAS, 361, 1287
McDonald, P. et al. 2006, ApJS, 163, 80
Miller, C. J. et al. 2005, AJ, 130, 968
Monet, D.G. et al. 2003, AJ, 125, 984
Morrison, H.L. et al. 2001, AJ, 121, 283
Morrison, H. L., Norris, J., Mateo, M., Shectman, S. A.,
Dohm-Palmer, R. C., Helmi, A., & Freeman, K. 2003, AJ, 125,
2502
Munn, J.A., et al. 2004, AJ, 127, 3034
Padmanabhan, N. et al. 2007, ApJ, in press (astro-ph/0703454)
Paturel, G., Petit, C., Prugniel, Ph., Theureau, G., Rousseau, J.,
Brouty, M., Dubois, P., & Cambr´esy, L. 2003, A&A, 412, 45
Pier, J.R., Munn, J.A., Hindsley, R.B., Hennessy, G.S., Kent,
S.M., Lupton, R.H., & Ivezi´c, Z. 2003, AJ, 125, 1559
Prugniel, P., & Soubiran, C. 2001, A&A, 369, 1048
Re Fiorentin, P. et al. 2007, A&A, 467, 1373
Richards, G.T. et al. 2002, AJ, 123, 2945
S´anchez-Bl´azquez, P. et al. 2006, MNRAS, 371, 703
Schlegel, D.J, Finkbeiner, D.P., & Davis, M. 1998, ApJ, 500, 525
(SFD)
Schmidt, G.D. et al. 2003, ApJ, 595, 1101
Schreiber, M.R. & Gansicke, B.T. 2003, A&A, 406, 305
Sirko, E. et al. 2004, AJ, 127, 899
Smith, J.A. et al. 2002, AJ, 123, 2121
Stoughton, C. et al. 2002, AJ, 123, 485 (EDR paper)
Strauss, M.A. et al. 2002, AJ, 124, 1810
Subbarao, M., Frieman, J., Bernardi, M., Loveday, J., Nichol, B.,
Castander, F., & Meiksin, A. 2002, SPIE, 4847, 452
Tucker, D. et al. 2006, AN, 327, 8212
Valdes, F., Gupta, R., Rose, J. A., Singh, H. P., & Bell, D. J.
2004, ApJS, 152, 251
Vanden Berk, D.E. et al. 2004, ApJ, 601, 692
Wilhite, B.C., Vanden Berk, D.E., Kron, R.G., Schneider, D.P.,
Pereyra, N., Brunner, R.J., Richards, G.T., & Brinkmann, J.V.
2005, ApJ, 633, 638
Woolf, V. M. & Wallerstein, G. W. 2006, PASP, 118, 218
York, D.G. et al. 2000, AJ, 120, 1579
|
astro-ph/0412112 | 1 | 0412 | 2004-12-05T18:15:33 | Simulations of Dust in Interacting Galaxies | [
"astro-ph"
] | A new Monte-Carlo radiative-transfer code, Sunrise, is used to study the effects of dust in N-body/hydrodynamic simulations of interacting galaxies. Dust has a profound effect on the appearance of the simulated galaxies. At peak luminosities, about 90% of the bolometric luminosity is absorbed, and the dust obscuration scales with luminosity in such a way that the brightness at UV/visual wavelengths remains roughly constant. A general relationship between the fraction of energy absorbed and the ratio of bolometric luminosity to baryonic mass is found. Comparing to observations, the simulations are found to follow a relation similar to the observed IRX-Beta relation found by Meurer et al (1999) when similar luminosity objects are considered. The highest-luminosity simulated galaxies depart from this relation and occupy the region where local (U)LIRGs are found. This agreement is contingent on the presence of Milky-Way-like dust, while SMC-like dust results in far too red a UV continuum slope to match observations. The simulations are used to study the performance of star-formation indicators in the presence of dust. The far-infrared luminosity is found to be reliable. In contrast, the H-alpha and far-UV luminosity suffer severely from dust attenuation, and dust corrections can only partially remedy the situation. | astro-ph | astro-ph |
Simulations of Dust in Interacting Galaxies
Patrik Jonsson∗, T. J. Cox† and Joel R. Primack∗
∗University of California, Santa Cruz
†Harvard-Smithsonian Center for Astrophysics
Abstract. A new Monte-Carlo radiative-transfer code, Sunrise, is used to study the effects of dust
in N-body/hydrodynamic simulations of interacting galaxies. Dust has a profound effect on the
appearance of the simulated galaxies. At peak luminosities, ∼ 90% of the bolometric luminosity
is absorbed, and the dust obscuration scales with luminosity in such a way that the brightness at
UV/visual wavelengths remains roughly constant. A general relationship between the fraction of
energy absorbed and the ratio of bolometric luminosity to baryonic mass is found.
Comparing to observations, the simulations are found to follow a relation similar to the observed
IRX-b
relation found by Meurer et al. [1] when similar luminosity objects are considered. The
highest-luminosity simulated galaxies depart from this relation and occupy the region where local
(U)LIRGs are found. This agreement is contingent on the presence of Milky-Way-like dust, while
SMC-like dust results in far too red a UV continuum slope to match observations.
The simulations are used to study the performance of star-formation indicators in the presence of
dust. The far-infrared luminosity is found to be reliable. In contrast, the Ha and far-UV luminosity
suffer severely from dust attenuation, and dust corrections can only partially remedy the situation.
INTRODUCTION
Galaxy mergers are an important ingredient in the hierarchical picture of galaxy for-
mation. They transform disks to spheroids, and may have been responsible for forming
a majority of the stars in the Universe [2]. Interacting galaxies are also triggering the
most luminous starbursts in the Universe, in the Luminous and Ultraluminous Infrared
Galaxies. Numerical simulations have been used to study interacting galaxies and the
resulting starbursts [3, 4], but these simulations alone cannot predict what these objects
would look like to observers, as the spectacular bursts of star formation in (U)LIRGS
are completely hidden by interstellar dust. Calculating the effect of dust in galaxies is
complicated. The mixed geometry of stars and dust makes the dust effects geometry-
dependent and nontrivial to deduce. Because of this, a full radiative-transfer model is
necessary to calculate these effects realistically. Previous studies of dust in galaxies have,
with few exceptions [5, 6], not used information from hydrodynamic simulations. This
study applies a new Monte-Carlo radiative-transfer code, Sunrise, to the outputs from a
comprehensive suite of N-body/hydrodynamic simulations of merging galaxies [7].
The N-body simulations of merging galaxies consist of a comprehensive suite of
hydrodynamic simulations using the GADGET N-body/SPH code. The simulations
study the effects of merger mass ratio, encounter orbit, and progenitor galaxy structure,
as well as different star-formation and feedback prescriptions, on the ensuing starburst
and the structure of the merger remnant [7]. The merging galaxies have been modeled
to closely resemble observed spiral galaxies in the local Universe. Examples of results
from this study include the discovery of a shock-driven gas outflow from some merging
systems [8] and an improved measure of how the intensity of the starburst scales with
merger mass ratio in minor mergers. Our simulations also show how a disk reforms after
the merger event, as in simulations by Springel and Hernquist [9].
For mergers with mass ratios smaller than 1:5, there is little induced star formation.
Also, major mergers of smaller progenitor galaxies have fundamentally different star-
formation properties than larger ones, with larger and more prolonged enhancements of
star formation due to the merger event. Our suite of merging galaxy simulations is by far
the largest performed so far.
RADIATIVE-TRANSFER MODEL
In order to calculate the effects of dust in these simulations, a new Monte-Carlo
radiative-transfer code, Sunrise, was developed [10]. The geometry of stars and gas in
the N-body simulations and the star-formation history of the system are used as inputs to
the radiative-transfer calculation. To make calculations with such complicated geome-
tries feasible, Sunrise features an adaptive-mesh refinement grid and is shared-memory
parallel. For the stellar emission, SEDs from the Starburst99 population synthesis model
[11] are used. The dust model is taken from Weingartner and Draine [12], and dust is
assumed to trace metals in the simulations. Currently, the infrared dust emission is not
calculated self-consistently. Instead, the infrared templates of Devriendt et al. [13] are
used. A self-consistent dust emission calculation is planned for the future. In addition to
using N-body simulations as inputs, the code can be used to solve problems specified in
other ways. As a service to the community, Sunrise is being released to the public under
the GNU General Public License1.
SIMULATION RESULTS
Radiative-transfer calculations of over 20 simulated disk-galaxy major mergers have
been completed, resulting in over 11,000 simulated images and spectra [10]. Example
images are shown in Figure 1, and movies of entire merger simulations are available on
the Internet2. A smaller number of simulations of the isolated progenitor galaxies have
also been done.
The radiative-transfer calculations show that dust has a profound effect on the ap-
pearance of the simulated systems. At the most luminous phase of the merger, ∼ 90%
of the bolometric luminosity is absorbed by dust. The dust attenuation scales with lu-
minosity in such a way that the brightness at UV/visual wavelengths remains roughly
constant throughout the merger event, even though the bolometric luminosity of the sys-
tem increases by a factor of 4 due to the merger-driven starburst. A general relationship
1 The Sunrise web site is http://sunrise.familjenjonsson.org.
2 http://sunrise.familjenjonsson.org/thesis
FIGURE 1.
Images of a simulated late-stage galaxy merger showing, from left to right, GALEX FUV
band, SDSS r band and infrared dust emission. Because the radiative-transfer model currently does not
calculate a spatially dependent infrared SED, the IR image corresponds to bolometric dust luminosity.
The images cover 100 kpc. While the UV/visual images show a peculiar, obviously dusty, galaxy, the IR
image shows that the cores of the progenitor galaxies are still distinct and vigorously star forming. Images
like these can be compared to observations of interacting systems.
)
0
0
6
1
L
/
S
A
R
I
,
I
R
F
L
(
g
o
l
4
4
3
3
2
2
1
1
0
0
-1
-1
Simulations
MHC fit
MHC
(U)LIRGs
-2
-2
-1
-1
0
0
1
1
2
2
3
3
)
0
0
6
1
L
/
S
A
R
I
,
I
R
F
L
(
g
o
l
4
4
3
3
2
2
1
1
0
0
-1
-1
Simulations
MHC fit
MHC
(U)LIRGs
-2
-2
-1
-1
0
0
1
1
2
2
3
3
FIGURE 2. The IRX-b relation of the simulations (shaded region), compared to the results from Meurer
et al. [1, crosses] and Goldader et al. [14, diamonds/triangles]. On the left, only simulated galaxies
with bolometric luminosity Lbol < 2 · 1011 L⊙ have been included. This low-luminosity sample agrees
fairly well with the MHC correlation. On the right, only the highest-luminosity simulated galaxies, with
Lbol > 7 · 1011 L⊙ have been included. These points depart completely from the MHC galaxies and instead
occupy the region of (U)LIRGs from the Goldader et al. [14] sample.
between the fraction of energy absorbed and the ratio of bolometric luminosity to bary-
onic mass is found to hold in galaxies with metallicities > 0.7Z⊙ over a factor of 100 in
mass.
The accuracy with which the simulations describe observed starburst galaxies is eval-
uated by comparing them to observations by Meurer et al. [1] and Heckman et al. [15].
The simulations are found to follow a relation similar to the IRX-b
relation found by
Meurer et al. [1] when similar luminosity objects are considered. The highest-luminosity
simulated galaxies depart from this relation and occupy the region where local (U)LIRGs
b
b
are found. These results are shown in Figure 2. Comparing to the Heckman et al. [15]
sample, the simulations are found to obey the same relations between UV luminosity,
UV color, IR luminosity, absolute blue magnitude and metallicity as the observations.
This agreement is contingent on the presence of a realistic mass-metallicity relation, and
Milky-Way-like dust. In contrast with earlier studies [16], SMC-like dust results in far
too red a UV continuum slope to match observations. On the whole, the agreement be-
tween the simulated and observed galaxies is impressive considering that the simulations
have not been fit to agree with the observations, and we conclude that the simulations
provide a realistic replication of the real universe.
The simulations are then used to study the performance of star-formation indicators
in the presence of dust. The far-infrared luminosity is found to be a reliable tracer of star
formation, as long as the star-formation rate is larger than about 1 M⊙/ yr. In contrast, the
Ha and far-ultraviolet luminosities suffer severely from dust attenuation, as expected.
Published dust corrections based on the Balmer line ratios [17] or the ultraviolet spectral
slope [18] only partially remedy the situation, still underestimating the star-formation
rate by up to an order of magnitude.
ACKNOWLEDGMENTS
PJ has been supported by a University Collaborative Grant through the Institute for
Geophysics and Planetary Physics (IGPP). This work used resources of the National
Energy Research Scientific Computing Center (NERSC), which is supported by the
Office of Science of the U.S. Department of Energy.
REFERENCES
Somerville, R. S., Primack, J. R., and Faber, S. M., MNRAS, 320, 504 -- 528 (2001).
1. Meurer, G. R., Heckman, T. M., and Calzetti, D., ApJ, 521, 64 -- 80 (1999).
2.
3. Mihos, J. C., and Hernquist, L., ApJ, 464, 641 (1996).
4.
5. Bekki, K., and Shioya, Y., ApJ, 542, 201 -- 215 (2000).
6. Bekki, K., and Shioya, Y., A&A, 362, 97 -- 104 (2000).
7. Cox, T. J., Star Formation and Feedback in Simulations of Interacting Galaxies, Ph.D. thesis, UC
Springel, V., MNRAS, 312, 859 -- 879 (2000).
Santa Cruz (2004), http://physics.ucsc.edu/~tj/work/thesis/thesis.pdf.
8. Cox, T. J., Primack, J., Jonsson, P., and Somerville, R. S., ApJ, 607, L87 -- L90 (2004).
9.
10. Jonsson, P., Simulations of Dust in Interacting Galaxies, Ph.D. thesis, University of California, Santa
Springel, V., and Hernquist, L., astro-ph/0411379 (2004).
Cruz (2004), http://sunrise.familjenjonsson.org/thesis.
11. Leitherer, C., Schaerer, D., Goldader, J. D., Delgado, R. M. G., Robert, C., Kune, D. F., de Mello,
D. F., Devost, D., and Heckman, T. M., ApJS, 123, 3 -- 40 (1999).
12. Weingartner, J. C., and Draine, B. T., ApJ, 548, 296 -- 309 (2001).
13. Devriendt, J. E. G., Guiderdoni, B., and Sadat, R., A&A, 350, 381 -- 398 (1999).
14. Goldader, J. D., Meurer, G., Heckman, T. M., Seibert, M., Sanders, D. B., Calzetti, D., and Steidel,
C. C., ApJ, 568, 651 -- 678 (2002).
15. Heckman, T. M., Robert, C., Leitherer, C., Garnett, D. R., and van der Rydt, F., ApJ, 503, 646 (1998).
16. Gordon, K. D., Calzetti, D., and Witt, A. N., ApJ, 487, 625 (1997).
17. Calzetti, D., Kinney, A. L., and Storchi-Bergmann, T., ApJ, 429, 582 -- 601 (1994).
18. Bell, E. F., and Kennicutt, R. C., ApJ, 548, 681 -- 693 (2001).
|
astro-ph/9912430 | 1 | 9912 | 1999-12-20T20:18:54 | The luminosity function of galactic ultra-compact HII regions and the IMF for massive stars | [
"astro-ph"
] | The population of newly formed massive stars, while still embedded in their parent molecular clouds, is studied on the galactic disk scale. We analyse the luminosity function of IRAS point-like sources, with far-infrared (FIR) colours of ultra-compact HII regions, that have been detected in the CS(2-1) line - a tracer of high density molecular gas. The FIR luminosities of 555 massive star forming regions (MSFRs), 413 of which lie within the solar circle, are inferred from their fluxes in the four IRAS bands and from their kinematic distances, derived using the CS(2-1) velocity profiles. The luminosity function (LF) for the UCHII region candidates shows a peak well above the completeness limit, and is different within and outside the solar circle (96% confidence level). While within the solar circle the LF has a maximum for 2E5 Lo, outside the solar circle the maximum is at 5E4 Lo. We model the LF using three free parameters: -alpha, the exponent for the initial mass function (IMF) expressed in log(M/Mo); -beta, the exponent for a power law distribution in N*, the number of stars per MSFR; and N*max, an upper limit for N*. While alpha has a value of \~ 2.0 throughout the Galaxy, beta changes from ~ 0.5 inside the solar circle to \~ 0.7 outside, with a maximum for the number of stars per MSFR of ~650 and \~450 (with 1 <M/Mo< 120). While the IMF appears not to vary, the average number of stars per MSFR within the solar circle is higher than for the outer Galaxy. | astro-ph | astro-ph |
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
08.12.3, 08.06.02, 09.08.1, 09.13.2, 13.09.3, 13.19.3
ASTRONOMY
AND
ASTROPHYSICS
May 17, 2021
The luminosity function of galactic ultra-compact H ii
regions and the IMF for massive stars ⋆
S. Casassus1,2, L. Bronfman1, J. May1, and L. -- A. Nyman3,4
1 Departamento de Astronom´ıa, Universidad de Chile, Casilla 36-D, Santiago, Chile
2 Astrophysics, University of Oxford, Keble Road, Oxford OX1 3RH, UK
3 SEST, ESO-La Silla, Casilla 19001, Santiago 19, Chile
4 Onsala Space Observatory, S -- 439 92 Sweden
received; accepted
Abstract. The population of newly formed massive stars,
while still embedded in their parent molecular clouds, is
studied on the galactic disk scale. We analyse the lu-
minosity function of IRAS point-like sources, with far-
infrared (FIR) colours of ultra-compact H ii regions, that
have been detected in the CS(2 -- 1) line - a tracer of high
density molecular gas. The FIR luminosities of 555 mas-
sive star forming regions (MSFRs), 413 of which lie within
the solar circle, are inferred from their fluxes in the four
IRAS bands and from their kinematic distances, derived
using the CS(2 -- 1) velocity profiles. The luminosity func-
tion (LF) for the UCH ii region candidates shows a peak
well above the completeness limit, and is different within
and outside the solar circle (96% confidence level). While
within the solar circle the LF has a maximum for 2 105 L⊙,
outside the solar circle the maximum is at 5 104 L⊙. We
model the LF using three free parameters: −α, the ex-
ponent for the initial mass function (IMF) expressed in
log(M/M⊙); −β, the exponent for a power law distribu-
tion in N ⋆, the number of stars per MSFR; and N ⋆
max, an
upper limit for N ⋆. While α has a value of ∼ 2.0 through-
out the Galaxy, β changes from ∼ 0.5 inside the solar circle
to ∼ 0.7 outside, with a maximum for the number of stars
per MSFR of ∼650 and ∼450 (with 1 ≤ M/M⊙ ≤ 120).
While the IMF appears not to vary, the average number
of stars per MSFR within the solar circle is higher than
for the outer Galaxy.
Key words: stars:luminosity function, mass function --
stars: formation -- ISM: H ii regions -- ISM: molecules --
infrared: ISM: continuum -- radio lines: ISM
1. Introduction
A new tool is now available to probe the population of re-
cently formed massive stars, while still embedded in their
Send offprint requests to: L. Bronfman
⋆ based partly on results collected at the European Southern
Observatory, La Silla, Chile
parent clouds. Such stars are surrounded by a compact
H ii region, with an ionization front working outwards into
the cloud. Regions undergoing massive star formation con-
tain one or more ultra-compact H ii (UCH ii) regions, and
possibly more evolved H ii regions. Bronfman et al. (1996,
BNM) completed a survey in CS(2 -- 1) towards IRAS point
sources satisfying the Wood & Churchwell (1989a) far-
infrared (FIR) colour criteria for UCH ii regions. The CS
molecule is a tracer of high density molecular gas; a CS(2 --
1) detection strengthens the UCH ii region identification
and provides kinematic information. BNM detected 843
sources (hereafter IRAS/CS sources), whose azimuthally
averaged galactic distribution is presented in Bronfman
et al. (2000, BCMN). In this work we construct the lumi-
nosity function (LF) for the IRAS/CS sources, and show
that it presents significant variations with galactocentric
radius. We interpret the shape of the IRAS/CS sources LF
and its variations in terms of an ensemble of massive star
forming regions. The stellar content of IRAS/CS sources is
characterised by the number of stars and their mass spec-
trum which, because of the youth of the systems, is taken
to represent closely their initial mass function (IMF).
The use of young tracers to derive the IMF minimises
the dependence on modelling, which afflicts most previ-
ous studies. The standard approach to determining the
high mass end of the IMF has been through O and B star
counts in the solar neighbourhood (e.g. Miller & Scalo
1979, Lequeux 1979). The average mass spectrum of newly
formed stars is taken as a power law, and the value of the
exponent characterises the IMF. These approaches require
assumptions on the star formation history, and have the
drawback of not probing the whole galactic disk. Garmany
et al. (1982) addressed the question of large scale varia-
tions in the IMF exponent, and they favour a decrease
with galactocentric radius (although their result has been
reinterpreted, e.g. Massey 1998). A different tool was used
by V´azquez & Feinstein (1989), who linked the variations
in the open cluster luminosity function (Burki 1977) to
variations in the IMF index.
2
S. Casassus et al.: The luminosity function of galactic UCH ii regions
Young objects such as H ii regions have also been used
to study massive stars in the galactic context. McKee &
Williams (1997, MW97) characterised the population of
newly formed massive stars through the luminosity func-
tion of OB associations. The ionising luminosity absorbed
by the gas surrounding OB associations can be traced
through the radio flux of H ii regions, and the frequency
distribution in luminosity of H ii regions is in turn a func-
tion of the number of exciting stars and their masses. How-
ever, MW97 a-priori fixed the shape of the IMF and the
distribution for the number of exciting stars. This method
is also strongly model dependent due to the lack of com-
plete information on galactic H ii regions, and the fact that
the population of H ii regions is not homogeneous in age.
Comer´on & Torra (1996, CT96) also analysed the galactic
disk distribution of newly formed massive stars, through
a sample of IRAS point sources with colours of UCH ii
regions. But they did not use kinematic information to
derive the galactic distribution of the UCH ii regions, and
their work is based on a position-independent LF.
The aims of this work are to present the FIR LF of
massive star forming regions still embedded in their parent
clouds; to investigate the LF large scale variations; and to
analyse the LF in terms of the embedded stars' mass spec-
trum. In Sect. 2 we construct the LF of IRAS/CS sources
for different sectors in the galactic disk, and show that it
presents significant variations with galactocentric radius.
In Sect. 3 we analyse the stellar population underlying
the IRAS/CS sample through a simple model based on a
synthetic ensemble of massive star forming regions (MS-
FRs). A search in parameter space is conducted in Sect.
4. The results of this analysis and the questions of the
number and the mass spectrum of stars per MSFR will be
addressed in Sect. 5. In particular, we will show that the
mass spectrum of newly formed massive stars is constant
across the galactic disk, although the average number of
stars per MSFR is lower outside the solar circle. Section
6 is a brief estimate of the fraction of lifetime O stars
spend in the embedded phase. In Sect. 7 we summarise
our conclusions.
2. FIR luminosity function of UCH ii regions
The velocity information provided by the CS(2 -- 1) obser-
vations from BNM allows, through the adoption of a ro-
tation curve, to derive their galactocentric distances, and
hence their heliocentric distances and luminosities. The
galactic disk is assumed to be in circular motion, with
R◦ = 8.5 kpc and V⊙ = 220 km s−1. For the region of the
Galaxy outside the solar circle (outer Galaxy) the proce-
dure is a simple coordinate transformation from galactic
longitude, latitude and velocity (l, b, Vlsr) to galactocen-
tric radius, height over the plane and azimuth (R, z, θ).
But such a transformation is bivalued for the region within
the solar circle (inner Galaxy). There is a heliocentric dis-
tance ambiguity such that, unless a source lies just on
the subcentral point (the tangent point to a galactocen-
tric ring for a given longitude), there are two points along
the line of sight, at the same distance on both sides of the
subcentral point, that have the same line of sight velocity.
The method we used to resolve the distance ambiguity
is described in BCMN. It is a statistical method that con-
sists in weighting the near and far distances with a normal
distribution in height over the galactic plane. Each source
is assigned an effective luminosity, which is the weighted
average of the near and far luminosities. The centroid and
(R), are
width of the vertical distribution, Z◦(R) and Z 1
determined through an iterative process for galactocentric
bins 0.1 R◦ wide. A consistency check for this method can
be found below in this Section.
2
The kinematic distances are not reliable in the direc-
tion of the galactic centre, and also when the line of sight
velocities are of the same order as the non-circular veloc-
ity components. Therefore, we excluded from the present
analysis all sources within ±10◦ of the galactic centre, and
within ±5◦ of the galactic anti-centre, as well as sources
with Vlsr ≤ 10 km s−1. The resulting range in galactocen-
tric radius, where the disk is properly sampled, excludes
the solar circle. We restricted the analysis to sources with
0.3 ≤ R/R◦ ≤ 0.9 and 1.1 ≤ R/R◦ ≤ 1.6.
The sources close to the subcentral points are an im-
portant consistency check of our method to resolve the
distance ambiguity within the solar circle. The kinematic
distance of a source at the subcentral point and in pure
circular motion about the galactic centre is uniquely deter-
mined. The subcentral sources sample was defined as the
subset with line of sight velocities no more than 10 km s−1
different in absolute value from the terminal velocity (the
maximum velocity expected for a given longitude in the
case of circular rotation).
We estimate the far infrared flux of an IRAS/CS source
by summing over the four IRAS bands,
FIRAS =
4
X
j=1
ν Fν (j),
(1)
where Fν (j) are the IRAS band flux densities, as listed in
the IRAS Point Source Catalog (1985). In order to test
this approximation we compared with the total fluxes re-
ported for 53 UCH ii regions by Wood & Churchell (1989b,
WC89, their Tables 17 and 18). Figure 1a shows the ra-
tio of the FIRAS fluxes (obtained by Eq. 1) to the total
fluxes of WC89 (which are integrated up to 100 µm), as
a function of FIRAS. Equation 1 overestimates the WC89
fluxes by about a constant 20%, but WC89 did not in-
clude a correction for the 100 µm -- 1 mm flux, which they
estimate could be as high as 50%. We thus expect Eq. 1
to be a good estimate of the total luminosity of IRAS/CS
sources within 30% (50% -- 20%).
Since we will average the luminosity function of UCH ii
regions over large areas of the galactic disk, it is impor-
tant to estimate the minimum luminosity above which
S. Casassus et al.: The luminosity function of galactic UCH ii regions
3
Fig. 1. a) The ratio of the fluxes published by WC89 to
FIRAS, as a function of FIRAS in 10−9 W m−2. b) The
number of sources (left hand scale) as a function of FIRAS
in 10−9 W m−2. The proportion of sources with an upper
limit in the 100 µm band is shown in triangles (right hand
scale)
the disk is properly sampled. The lowest flux FIRAS in
the sample, FIRAS = 6 10−13 W m−2, corresponds to
log(Lmin/L⊙) = 3.6 at a distance of 15 kpc. Figure 1b
shows a histogram of the total number of sources in our
sample as a function of FIRAS (without the velocity fil-
ter Vlsr ≤ 10 km s−1). The triangles show the fraction
of sources with only upper limits in the 100 µm band
(right hand scale). Could a significant number of sources
be missed by IRAS near the minimum detected flux? A
lower FIRAS sensitivity would be hinted at by an increased
fraction of upper limits in the reported IRAS fluxes, which
is not the case. However, the higher far-IR background to-
wards the central regions of the Galaxy results in a com-
pleteness limit of log(Lmin/L⊙) = 4.5 at 8.5 kpc, within
−10 < l < 10, −0.3 < b < 0.3. But over a broader
longitude range, log(Lmin/L⊙) = 4.0 at 8.5 kpc, within
−60 < l < 60, −0.3 < b < 0.3. As the luminosity functions
of the subcentral sources (which are all within 8.5 kpc of
the Sun) is in good agreement with that of the whole in-
ner Galaxy (see below), we take log(Lmin/L⊙) = 4 as the
completeness limit of the IRAS/CS sample1.
The LF of the whole sample of IRAS/CS sources, our
main observational result, appears to be significantly dif-
ferent inside and outside the solar circle; in Fig. 2 we dis-
tinguish between R < R◦ and R > R◦. The LFs cover a
very wide range in luminosity, over three orders of magni-
tude, which allows using logarithmic luminosity bins cor-
responding to a factor of 300%. Within the solar circle the
LF obtained using the effective luminosities is confirmed
to be a good estimate of the actual LF through its close
similarity with the LF of the sources near the subcen-
tral points, shown in dotted line2. For comparison, plac-
ing all the sources at the 'near' or 'far' kinematic distance
1 The CS(2 -- 1) detection requirement has been checked
(BCMN) not to introduce additional biases within 8.5 kpc
2 A total of 57 sources were used to calculate this LF. Sources
with galactocentric radii larger than 0.8 R◦ were excluded: our
definition of the subcentral source sample would otherwise in-
clude most sources in that region, because the non-circular
Fig. 2. The luminosity function for galactic UCH ii re-
gions, from the IRAS/CS sample. The whole disk was di-
vided at the solar circle, the upper and lower plots corre-
spond to the inner and outer Galaxy LFs, computed with
413 and 142 sources respectively. In the upper plot the
inner Galaxy LF derived from the effective luminosities is
shown in solid line, while the thick dotted line is the LF
for sources near the subcentral points (57 sources). The
distributions in these plots are normalised so that the ar-
eas under the histograms is one over log(L/L⊙)>4. Shot
noise gives 1-σ error bars on the subcentral LF of ∼25%
changes the peak of the LF as a function of logarithmic
luminosity from 4.25 to 5.75, while the LF obtained using
the effective luminosities peaks at 5.25. The good match
with the subcentral source sample LF lends strength to a
comparison of the luminosity functions between the outer
and inner Galaxy, based on the effective luminosities. We
will refer to the luminosity functions for the whole inner
and outer Galaxy by LFin and LFout.
The dominant source of uncertainty in the LF is shot
noise. The errors in the galactic disk surface FIR luminos-
ity amount to about 10% upwards, 20% downwards (Fig.
3 in BCMN). The fractional error on the FIR surface lu-
minosity represents the typical fractional error on the lu-
minosity of one source. These errors stem from the IRAS
100 µm band flux uncertainty, coupled with the kinematic
distance uncertainty due to non-circular motions of about
velocity components dominate the spread about the terminal
velocity
4
S. Casassus et al.: The luminosity function of galactic UCH ii regions
5 km s−1. Adding in quadrature the 30% uncertainty re-
lated to the use of Eq. 1, we have an average error on the
luminosity of a source of at most 36%. Compared to the
300% width of the luminosity bins, a 36% uncertainty is
negligible, apart from a slight smoothing effect without
practical consequence.
The differences in the LFs inside and outside the solar
circle are statistically significant. As a statistic for the
difference between the inner and outer LFs, we used a χ2
test which has the following expression in this context,
χ2 = X
i
(LFin
i − LFout
i
)2
LFout
i
LFin
i
(Nin∆ log L)2 +
(Nout∆ log L)2
,
(2)
where we sum over the bins above the luminosity limit of
104 L⊙. The result is χ2 = 11.4, or that LFin and LFout
are different at a significance level of 95.6% (with a χ2
distribution for 5 degrees of freedom, corresponding to the
number of bins with non-zero counts above the luminos-
ity limit). For comparison, the same test applied to the
northern and southern3 LFs inside the solar circle gives
a probability of 45% that the distributions are different,
so they are comparable relative to the differences between
the LFs inside and outside the solar circle. Another ap-
plication of this statistical test gives that LFin and the
subcentral source LF are the same at a confidence level of
87%.
We emphasize the presence of a peak in the LF of
IRAS/CS sources, well above the completeness limit. The
strongest evidence in that sense can be found in the LFs
for the outer Galaxy and for the subcentral sources, where
the Lmin completeness limits are lowest. The shape for the
LF we report is quite different from a power law functional
form (e.g. as used in CT96).
3. Synthetic fits to the IRAS/CS luminosity
function
One immediate consequence of the shape of the IRAS/CS
luminosity function is that these sources are better under-
stood as clusters of stars rather than in the framework of
one dominant star per source. A single exciting star would
result in a power law LF: an IMF index α (see below), and
mass-luminosity relation L ∝ M γ, give a probability dis-
tribution in luminosity p(L) ∝ L−
γ .
α+γ
We used a simple model to examine whether the vari-
ations in the LFs of IRAS/CS sources from inside to
outside the solar circle can be traced to the underlying
young stellar population. We proceed to describe a Monte
Carlo analysis for the ensemble of massive star forming
regions in the galactic disk. The luminosity of a MSFR
is the sum of the luminosities of each star, given by the
mass-luminosity relationship. We used a polynomial fit to
3 We refer to the sector with longitudes less than 180◦ as
the northern Galaxy, while longitudes greater than 180◦ cor-
respond to the southern Galaxy
the mass-luminosity relation of the tracks presented in
Schaller et al. (1992) for Z=0.02, at the first time step
they list,
log(L/L⊙) = −0.127 + 4.656 log(M/M⊙)
−0.764 log2(M/M⊙).
(3)
The metallicity dependence of the mass-luminosity re-
lation was neglected, as the Z=0.001 tracks in Schaller
et al. (1992) have luminosities within 10% of Eq. 3 for
M > 7 M⊙.
⋆
The synthetic population of MSFRs was generated in
the following way. The number of stars in a given MSFR,
N⋆, is generated randomly within the range 1 ≤ N⋆ ≤
N max
, and subject to a power law probability distribu-
tion with exponent −β, p(N⋆) ∝ N −β
. A discussion of
the model sensitivity on N max
will be found in Sect. 4.
Given N⋆, the total luminosity of a MSFR is calculated
by summing the individual luminosities of each star, using
the mass-luminosity relationship. The mass of each star is
generated randomly, within the range 1 ≤ M/M⊙ ≤ 120
and satisfying the IMF distribution,
⋆
⋆
p(M ) ∝ M −(1+α);
(4)
in this notation the Salpeter (1955) IMF corresponds to
α = 1.35. Thus the luminosity of each MSFR is randomly
generated with LMSFR = PN⋆
i=1 L(Mi), and an ensemble of
5000 MSFRs provides a population large enough to com-
pute the synthetic luminosity function (that this is one
order of magnitude larger than the number considered in
the observed LF does not affect the results, only helps to
reduce the random fluctuations).
An important simplification in this approach is that
the ensemble of MSFRs is assumed to be homogeneous
in age (see Sect. 6). Furthermore, it should be mentioned
before discussing the results of the model that the mass-
luminosity relation remains mainly theoretical for mas-
sive stars. Although Burkholder et al. (1997) give obser-
vational evidence that support the massive star models up
to 25 M⊙, the upper mass limit we used is 120 M⊙, where
to our knowledge no direct observational information ex-
ists to back the theoretical models.
4. Search in parameter space and the role of
N max
⋆
⋆
A first broad search for the parameters α and β requires
specifying N max
. It could be thought firsthand that any
⋆
value of N max
high enough to simulate infinity would do,
but we tried N max
⋆ = 10000 and did not obtain any fit,
over the range 1.5 < α < 2.7 , 0 < β < 2. We distin-
guished three cases, fixing the maximum number of stars
per MSFR, N max
, to 500, 1000 and 2000. Figure 3 shows
the χ2 cumulative probability for the goodness of fit as a
function of parameter space. In the observed IRAS/CS
LF, we have five bins with non-zero counts above the
⋆
S. Casassus et al.: The luminosity function of galactic UCH ii regions
5
completeness limit of 104 L⊙. The number of degrees of
freedom is thus 3, which corresponds to the five bins in
comparison less two free parameters (α and β). The fit to
LFout is a lot more noisy, a consequence of the reduced
number of IRAS/CS sources used to compute the LF.
⋆
⋆
It is apparent that for any value of N max
, α is similar
inside and outside the solar circle, while the acceptable
values for β are markedly different. It can also be noticed
from Fig. 3 that the best fit β are rather independent of
N max
, in contrast with the behaviour of the best fit IMF
index α. For N max
⋆ = 1000, the IMF index in the inner
Galaxy, at 50% confidence level for the goodness of fit
(with two free parameters), corresponds to 2.05 ≤ α ≤
2.15. This is very close to the result for the outer Galaxy,
2.1 ≤ α ≤ 2.3. The case where N max
⋆ = 500 seems to give
better results for R > R◦, and constrains 1.9 < α < 2.15
at 50% confidence. This range of values, ∆α = 0.1 for
R < R◦ and ∆α = 0.15 for R > R◦, will be used as an
indication of the uncertainty level in the best fit α.
As the distribution of N⋆ seems to be a bounded power
law, there exists a maximum for the number of stars born
in a MSFR within a finite mass range. It should be kept in
mind that the total number of stars could in fact depend
on the stellar masses and the star formation history of a
MSFR; we expect the parameter N⋆ to synthesise more
complex processes. Observational constraints to fix N max
are difficult to find, because stellar censuses are available
only for much larger regions like OB associations, open
clusters, or longer lived H ii regions.
⋆
5. The IMF index and possible large scale
variations
After finding reasonable values for N max
search to include all three parameters, and obtain
⋆
, we extend the
(α, β, N max
(α, β, N max
⋆
⋆
) = (1.988, 0.49, 646), for R < R◦, and
(5)
) = (1.991, 0.73, 450), for R > R◦.
Thus the best fitting sets of parameters have α = 2.0.
Values for the significance of the fits are ∼ 60%, with two
degrees of freedom (5 bins in comparison less 3 free param-
eters). Figure 4 shows the models that best fit the LFs.
Due to the Monte Carlo approach, secondary χ2 minima
are found about the true minimum. Multiple minimiza-
tion runs with various initial guesses fluctuate about the
above values: from (1.98, 0.57, 708) to (2.03, 0.40, 748) for
R < R◦.
⋆
Keeping N max
fixed at the best fit value, with the ∆α
uncertainties quoted in the previous section (i.e. above a
50% confidence level in a 2-D slice), our best values for α
are 1.95 < α < 2.05 in the inner Galaxy, and 1.9 < α <
2.15 in the outer Galaxy.
As an indication of the uncertainty level related to
increasing the bolometric fluxes to
the use of Eq. 1,
2 FIRAS results in (α, β, N max
) = (1.72, 0.29, 457) for
R < R◦. Only very low significance fits were found
⋆
Fig. 3. Search in (α, β) parameter space for the best fit
. Plotted is the χ2 cu-
model, with three guesses for N max
mulative probability for the goodness of fit (with 2 free
parameters), with contours at 10%, 30%, 50% and 70%.
The upper two plots were computed with N max
⋆ = 2000,
the middle two with N max
⋆ = 1000, and the bottom two
with N max
⋆ = 500
⋆
using either 5 FIRAS, or 0.5 FIRAS. A factor of 2 in
the mass-luminosity relation L(M ) gives (α, β, N max
) =
(2.005, 0.48, 700), i.e. no significant change. A steep IMF
index seems to be a robust result of our analysis.
⋆
The IMF index α for the inner and outer Galaxy seems
to be the same, although β changes significantly. In other
words, although the average mass of newly formed stars
is the same inside and outside the solar circle, the average
number of stars per star forming region and in a finite
mass range is lower in the outer Galaxy. The best fits
of the parameter β imply that the expectation value for
the number of stars per MSFR with 1 ≤ M/M⊙ ≤ 120,
< N >, decreases from 225 for R < R◦ to 120 for R >
6
S. Casassus et al.: The luminosity function of galactic UCH ii regions
IMF, the exact value of the IMF index is still an open
issue. Our contribution to this debate is that there cer-
tainly are differences in the physical processes governing
star formation in the outer spiral arms and the molecular
ring. In our simple approach, and understanding that we
restricted our study to regions of massive star formation
embedded in dense molecular cores and with at least one
UCH ii region, it seems that although the IMF index is
constant, the average number of stars born per region is
lower outside the solar circle. We favour a rather steep
IMF index, close to the values from Brown (1998).
6. Lifetime of UCH ii regions
UCH ii regions are fairly short lived in comparison with
giant H ii regions. Wood & Churchwell (1989a) estimated
the lifetimes of UCH ii regions by comparing the number of
O stars in the solar neighbourhood (Conti et al. 1983) and
the number of IRAS point sources that match the IRAS
colours of UCH ii regions. They conclude that 10% to 20%
of an O star's main sequence lifetime is spent embedded
in a molecular cloud, in the UCH ii phase. This fraction
has been reevaluated to only 0.5% by CT96, and our es-
timate is ≤ 2% (see below). Under the assumption that
the luminosity of a MSFR is dominated by the most mas-
sive star, CT96 show that if there is a variation of UCH ii
lifetime with stellar mass, it is not as significant as the un-
certainties in the IMF. The crossing time at 10 km s−1 is
104 yrs for a typical UCH ii region (i.e. G34.3+0.2, which
at a distance of 3.7 kpc is about 5 1017 cm). Thus the mar-
gin between the dynamical timescale and the time O stars
spend in the embedded phase of < 105 yrs is narrow, and
can be affected by many parameters other than the mass of
the most massive star, such as clumpiness of the molecular
clouds and relative motions between the exciting star and
the molecular clumps. It is unlikely that the lifetimes of
UCH ii regions depend strongly on their stellar contents.
From our synthetic population of MSFRs, we find that
the average number of O stars per MSFR, with luminosi-
ties in excess of L(20 M⊙), is 0.55 for R < R◦ and 0.29
for R > R◦ - we will consider 0.5 O stars per IRAS/CS
source. Within 2.5 kpc of the Sun, Conti et al. (1983) re-
port 436 O stars with M > 20 M⊙. We have 15 IRAS/CS
sources above the luminosity limit L(20 M⊙) (counting all
sources, outside the cuts in longitude described in Sect.
2 but including the sources with Vlsr ≤ 10 km s−1).
The average fraction of lifetime spent in the embedded
phase would thus be 0.5 × 15/(436 + 0.5 × 15) = 0.017.
Alternatively, counting the IRAS/CS sources with galac-
tocentric radii within 0.9 to 1.1 R◦, we have 40 sources
more luminous than L(20 M⊙), distributed over an area of
80.5 kpc2 (BCMN). This would give 9.8 IRAS/CS sources
within 2.5 kpc of the Sun. The average fraction of life-
time O stars spend in the embedded phase would thus be
0.5 × 9.8/(436 + 0.5 × 9.8) = 0.011. This latter approach
has the advantage of not suffering as much from the pe-
Fig. 4. Best fit models (dotted lines) to the inner and
outer Galaxy IRAS/CS LFs (solid lines), above the com-
pleteness limit log(L/L⊙) = 4. In ordinates is the LF nor-
malised over log(L/L⊙) ≥ 4, as a function of log(L/L⊙)
R◦. The decrease in < N > is not an effect related to
the area covered by the 100 µm IRAS beam: due to the
lower FIR background towards the outer Galaxy, the IRAS
point sources are detected to greater distances than in the
inner Galaxy (the average distance to IRAS/CS sources
is 6.92 kpc for the outer Galaxy, and 5.32 kpc in terms of
the effective distances for the inner Galaxy).
Is there a gradient in the IMF index with galactocen-
tric radius? How does our value for the IMF index compare
with other estimates? The gradient proposed by Garmany
et al. (1982) has been re-interpreted as a result of contami-
nation from field O stars (Massey et al. 1995a), which seem
to derive from a very steep IMF. The values Massey et al.
(1995b) and Massey (1998) quote for α in the Milky Way
and the Magellanic Clouds OB associations are 1.1 ± 0.1
and 1.3 ± 0.1, which argues against a metallicity depen-
dence of α, at least over 10 -120 M⊙. But constraining the
analysis to nearby OB associations gives steeper values,
Claudius & Grosbøl (1980) give α = 1.9 over 2.2 -10 M⊙,
and Brown (1998) gives α=1.9 for the Upper Scorpius
subgroup of Sco OB2, over 3 -16 M⊙. Brown (1998) used
Hipparcos data to ascertain membership (see de Zeeuw et
al. 1999), which is crucial in the coeval approximation;
the very large OB associations in Massey et al. probably
have a complex star formation history. Although Massey
et al. establish a good case for a universal massive star
S. Casassus et al.: The luminosity function of galactic UCH ii regions
7
studentship. L.B., S.C., and J.M. acknowledge support from
FONDECYT-Chile grant 8970017, and from a C´atedra Presi-
dencial en Ciencias 1997.
References
Bronfman, L., Casassus, S., May, J., Nyman L.-A., 2000, A&A
(submitted)(BCMN)
Bronfman, L., Nyman, L.-A., May, J., 1996, A&AS 115, 81
(BNM)
Brown A.G.A., 1998, PASPC 142, 45
Burkholder V., Massey P., Morrell N., 1997, ApJ 490, 328
Burki G., 1977, A&A 57, 135
Claudius M., Grosbøl P.J., 1980, A&A 87, 339
Comer´on F., Torra J., 1996, A&A 314, 776 (CT96)
Conti P.S., Garmany C.D., de Loore C., Vanbeveren D., 1983,
ApJ 274, 302
deZeeuw P.T., Hoogerwerf R., Bruijne J.H., Brown A.G.A.,
Blaauw A., 1999, AJ 117, 354
Garmany C.D., Conti P.S., Chiosi C. 1982, ApJ 263, 777
IRAS Catalogs and Atlases, Explanatory Supplement, 1985, ed.
C.A. Beichman, G. Neugebauer, H.J. Habing, P.E. Clegg,
T.J. Chester (Government Printing Office, Washington,
DC)
IRAS Point Source Catalog, 1985, Joint IRAS Science Working
Group (Government Printing Office, Washington, DC)
Lequeux J., 1979, A&A 80, 35
Massey P., 1998, PASPC 142, 17
Massey P., Johnson K.E., DeGioia-Eastwood K., Garmany
C.D., 1995a, ApJ 454, 161
Massey P., Lang C.C., DeGioia-Eastwood K., Garmany C.D.,
1995b, ApJ 438, 188
McKee C.F., Williams J.P., 1997, ApJ 476, 144 (MW97)
Miller G.E., Scalo J.M., 1979, ApJSS 41, 513
Salpeter E.E., 1955, ApJ 121, 161
Schaller G., Schaerer D., Meynet G., Maeder A., 1992, A&AS
96, 269
V´azquez R.A., Feinstein A., 1989, Rev. Mex. Astron. Astrofis.
17, 3
Wood D.O.S., Churchwell E., 1989a, ApJ 340, 265
Wood D.O.S., Churchwell E., 1989b, ApJS 69, 831 (WC89)
culiar velocities of local sources. But some O stars in the
transition between the embedded UCH ii phase and the
field stars could have been missed by Conti et al. (1983).
Therefore we estimate that O stars spend ≤ 2% of their
lifetime in the embedded phase.
7. Conclusions
We have constructed the FIR luminosity function for
IRAS point sources with colours of UCH ii regions and
CS(2 -- 1) detections, for different sectors in the galactic
disk. The LFs are reliable above a luminosity of 104 L⊙,
and extend to ∼ 106 L⊙ in the high luminosity end, with
a peak at ∼ 105 L⊙. We analysed the trends in the LF of
IRAS/CS sources in terms of an ensemble of massive star
forming regions, and two free parameters suffice to provide
a remarkably good fit to the shape of the LF. The fits re-
quired a maximum for the number of stars born in a given
MSFR with 1 ≤ M/M⊙ ≤ 120. The best results were ob-
tained setting N max
⋆ = 646 for R < R◦, and N max
⋆ = 450
for R > R◦. A few conclusions can be summarised:
1. The LFs inside and outside the solar circle are different
at 96% confidence level. The LF within the solar circle,
built with 413 sources, peaks at 2 105L⊙, while the LF
outside the solar circle, built with 142 sources, peaks
at 5 104 L⊙.
2. The IMF index we obtain is α = 2 and constant with
galactocentric radius. At 50% cumulative probability
for the goodness of fit, and keeping N max
fixed at the
best fit value, 1.95 < α < 2.05 in the inner Galaxy,
and 1.9 < α < 2.15 in the outer Galaxy.
⋆
3. A power law distribution for the number of stars per
MSFR has an exponent −β ∼ −0.49 in the inner
Galaxy. But in the outer Galaxy the best fit model cor-
responds to −β ∼ −0.73. Thus, the expectation value
for the number of stars per MSFR with 1 ≤ M/M⊙ ≤
120 decreases from 225 for R < R◦ to 120 for R > R◦.
The results of our analysis show that the observed lu-
minosity functions for UCH ii regions can be traced to the
underlying young population. The differences within and
outside the solar circle reflect a decrease in the average
number of stars per massive star forming region towards
the outer Galaxy, rather than a steeper IMF.
Acknowledgements. This article benefited from the suggestions
and encouraging comments of the referee, James Lequeux. We
are also grateful to Rodrigo Soto and Mark Seaborne for help-
ful discussions. The staff at the SEST and OSO telescopes
kindly assisted us within the course of the observations. The
Swedish-ESO Submillimetre Telescope is operated jointly by
ESO and the Swedish National Facility for Radioastronomy,
Onsala Space Observatory, at Chalmers University of Tech-
nology. The Onsala 20m telescope is operated by the Swedish
National Facility for Radioastronomy, Onsala Space Observa-
tory, at Chalmers University of Technology. S.C. acknowledges
support from Fundaci´on Andes and PPARC through a Gemini
|
astro-ph/9803177 | 3 | 9803 | 1998-06-23T10:53:28 | Quantifying uncertainties in primordial nucleosynthesis without Monte Carlo simulations | [
"astro-ph",
"hep-ph",
"nucl-th"
] | We present a simple method for determining the (correlated) uncertainties of the light element abundances expected from big bang nucleosynthesis, which avoids the need for lengthy Monte Carlo simulations. Our approach helps to clarify the role of the different nuclear reactions contributing to a particular elemental abundance and makes it easy to implement energy-independent changes in the measured reaction rates. As an application, we demonstrate how this method simplifies the statistical estimation of the nucleon-to-photon ratio through comparison of the standard BBN predictions with the observationally inferred abundances. | astro-ph | astro-ph | INFNFE-02-98
BARI-TH/297-98
OUTP-98-14-P
astro-ph/9803177
Quantifying uncertainties in primordial nucleosynthesis
without Monte Carlo simulations
G. Fiorentini a, E. Lisi b, S. Sarkar c, and F. L. Villante a
aDipartimento di Fisica and Sezione INFN di Ferrara, Via del Paradiso 12, I-44100 Ferrara, Italy
bDipartimento di Fisica and Sezione INFN di Bari, Via Amendola 173, I-70126 Bari, Italy
cTheoretical Physics, University of Oxford, 1 Keble Road, Oxford OX1 3NP, UK
Abstract
We present a simple method for determining the (correlated) uncertainties of
the light element abundances expected from big bang nucleosynthesis, which
avoids the need for lengthy Monte Carlo simulations. Our approach helps to
clarify the role of the different nuclear reactions contributing to a particu-
lar elemental abundance and makes it easy to implement energy-independent
changes in the measured reaction rates. As an application, we demonstrate
how this method simplifies the statistical estimation of the nucleon-to-photon
ratio through comparison of the standard BBN predictions with the observa-
tionally inferred abundances.
26.35.+c, 98.80.Ft
8
9
9
1
n
u
J
3
2
3
v
7
7
1
3
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Typeset using REVTEX
1
I. INTRODUCTION
Big bang nucleosynthesis is entering the precision era [1]. On the one hand, there has been
major progress in the observational determination of the abundances of the light elements
D [2,3], 3He [4,5], 4He [6,7], and 7Li [8,9], although the increasing precision has highlighted
discrepancies between different measurements (see Refs. [10 -- 12] for recent assessments).
Secondly, we have a sound analytical understanding of the physical processes involved [13,14]
and the standard BBN computer code [15,16] which incorporates this physics is robust and
can be easily altered to accomodate changes in the input parameters, e.g. nuclear reaction
rates [17]. The comparison of increasingly accurate observationally inferred and theoretical
abundances will further constrain the values of fundamental parameters, such as the nucleon
density parameter (see, e.g., Ref. [18]) or extra degrees of freedom related to possible new
physics beyond the Standard Model (see, e.g., Ref. [19]). It goes without saying that error
evaluation represents an essential part of such comparisons.
Because of the complex interplay between different nuclear reactions, it is not straight-
forward to assess the effect on a particular elemental yield of the uncertainties in the exper-
imentally determined reaction rates. The authors of Ref. [20] first employed Monte Carlo
methods to sample the error distributions of the relevant reaction cross-sections which were
then used as inputs to the standard BBN computer code. This enables well-defined confi-
dence levels to be attached to the theoretically predicted abundances, e.g. the abundance
range within which say 95% of the computed values fall correspond to 95% C.L. limits on the
expected abundance. It was later realized that error correlations are also relevant, and can
be estimated with the same technique [21,22]. The Monte Carlo (MC) approach has since
become the standard tool for comparing theory and data [17,23 -- 26]. However, although it
can include refinements such as asymmetric or temperature-dependent reaction rate uncer-
tainties [17], it requires lengthy calculations which need to be repeated each time (any of)
the input parameters are changed or updated. Since we may expect continued improvement
in the determination of the relevant parameters, it is desirable to have a faster method for
error evaluation and comparison with observations.
In this work we propose a simple method for estimation of the BBN abundance un-
certainties and their correlations which requires little computational effort. The method,
based on linear error propagation, is described in Sec. II. A concrete application is given in
Sec. III, where theory and observations are compared using simple χ2 statistics to obtain
the best-fit value of the nucleon-to-photon ratio. In Sec. IV we study with this method the
relative importance of different nuclear reactions in determining the synthesized abundances.
Conclusions and perspectives for further work are presented in Sec. V.
II. PROPAGATING INPUT CROSS SECTION UNCERTAINTIES TO OUTPUT
ELEMENTAL ABUNDANCES
A. Notation and input
The four relevant element abundances Yi considered in this work are defined in Table I.
(Note that the abundance of 4He is conventionally quoted as a mass fraction, while the
2
abundances of D, 3He and 7Li are ratios by number.) In BBN calculations, the Yi's depend
both on model parameters (the nucleon-to-photon ratio η, the number of neutrino families
Nν, etc.) and on a network of nuclear reactions Rk:
Yi = Yi(η, Nν, . . . ; {Rk}) .
(1)
The most important Rk's are listed, numbered as in Ref. [16], in the first two columns
of Table II, while our default inputs for the rates Rk and their 1σ uncertainties ±∆ Rk
are given in the third and fourth columns. The numerical values are given in ratio to the
reference reaction rates compiled in Table 1 of Ref. [17]; we have chosen identical values
(i.e., Rk = 1) except for R1, the neutron decay rate, where we adopt the most recent world
average for the neutron lifetime of τn = 886.7 ± 1.9 s [27,28], as compared to the value of
τn = 888.54 ± 3.73 s used in Ref. [17]. The fractional uncertainties ±∆Rk/Rk (k 6= 1) have
also been taken from Ref. [17] (see their Table 2), assuming conservatively the largest value
for the temperature-dependent errors ∆ R7 and ∆ R10.1
B. The method
For simplicity we consider only the standard BBN case (i.e., Nν = 3, etc.), so that η is
the only model parameter being varied in the calculation of the abundances Yi and of their
uncertainties σi:
Yi = Yi(η) ± σi(η) .
(2)
Our method can however be easily generalized to nonstandard cases.
For a relatively small change δ Rk of the input rate Rk (Rk → Rk + δ Rk), the corre-
sponding deviation δ Yi of the i-th elemental abundance (Yi → Yi + δ Yi), as given by linear
propagation, reads
δ Yi(η) = Yi(η)Xk
λik(η)
δ Rk
Rk
,
(3)
where the functions λik(η) represent the logarithmic derivatives of Yi with respect to Rk:
λik(η) =
∂ ln Yi(η)
∂ ln Rk(η)
.
(4)
In general, the deviations δ Yi in Eq. (3) are correlated, since they all originate from
the same set of reaction rate shifts {δ Rk}. The global information is contained in the error
matrix (also called covariance matrix) [29], which is a generalization of the "error vector" δ Yi
in Eq. (3). In particular, the abundance error matrix σ2
ij(η) obtained by linearly propagating
the input ±1σ reaction rate uncertainties ±∆ Rk to the output abundances Yi reads:
1Our method for error propagation requires that the ∆ Rk/Rk's be constant (i.e., temperature
independent). We will comment on this point at the end of this Section.
3
σ2
ij(η) = Yi(η)Yj(η)Xk
λik(η)λjk(η)(cid:18)∆ Rk
Rk (cid:19)
2
.
(5)
This matrix completely defines the abundance uncertainties. In particular, the 1σ abundance
errors σi of Eq. (2) are given by the square roots of the diagonal elements,
while the error correlations ρij can be derived from Eqs. (5,6) through the standard definition
σi(η) = qσ2
ii(η) ,
(6)
ρij(η) =
σ2
ij(η)
σi(η) σj(η)
.
(7)
Thus Eqs.(2 -- 7) represent all that is required to calculate the errors in the predicted abun-
dances and their correlations.
Note that the relevant physics is contained entirely in the central values Yi and in their
logarithmic derivatives λik, which have to be evaluated just once with a BBN numerical
code, thus dramatically reducing the required computing time.2 We have made a further
check of the linearity of the error propagation by calculating the logarithmic derivatives with
increments equal to ∆ Rk (default) and 2 ∆ Rk, obtaining practically the same functions λik
in either case. This means that doubling the error on Rk also doubles the corresponding
error component of Yi, i.e. the error propagation is indeed linear.
We think it useful to present the results of this exercise in the form of tables so all
calculations that will now follow can be done on a pocket calculator. Tables III and IV show
the coefficients of polynomial fits to Yi and λik, respectively, for η in the usually considered
range 10−10 − 10−9.3 The abundances Yi(η) with their associated ±2 standard deviation
error bands calculated through Eq. (6) are shown in Fig. 1. The functions λik(η) are shown
in Fig. 2; note that some of these vary strongly (and even change sign) with η, indicating
that the physical dependence of the Yi's on the Rk's may be quite subtle. Although some
general features of this dependence have been addressed in Ref. [14], further work is needed to
interpret the functional form of the λik's in Fig. 2. We intend to address this issue elsewhere.
Finally, Fig. 3 displays the fractional uncertainties σi/Yi and their correlations ρij as derived
from Eqs. (5 -- 7). Notice that, in general, the error correlations are non-negligible and should
be properly taken into account in statistical analyses, as first emphasized in Ref. [21].
2The logarithmic derivatives are numerically defined as λik(η) = [Yi(η, Rk + ∆ Rk) − Yi(η, Rk −
∆ Rk)]Rk/2Yi(η, Rk)∆ Rk, at given η. We find negligible difference between left and right
derivatives.
3Small corrections to the helium abundance Y4 due to Coulomb, radiative and finite temperature
effects, finite nucleon mass effects and differential neutrino heating, have been incorporated accord-
ing to the prescription given in Ref. [19]. We have used the BBN code [16] with the lowest possible
settings of the time steps in the (2nd order) Runge-Kutta routine, which allows rapid convergence
to within 0.01% of the true value [30]. We understand that our results are in good agreement with
a recent independent computation of Y4 using a new BBN computer code [31].
4
In summary, the recipe for evaluating the BBN uncertainties ±σi affecting the Yi's for a
given value of η is:
(i) Determine the abundances Yi(η) and their logarithmic derivatives λik(η) using Tables
III and IV, respectively;
(ii) If the central values of the reaction rates Rk are updated (Rk → Rk +δ Rk) with respect
to those reported in Table II, then update also the central values of the abundances
(Yi → Yi + δ Yi) through Eq. (3);
(iii) For given reaction rate uncertainties ∆ Rk (e.g., from Table III), compute the abun-
dance errors σi and their correlations ρij using Eqs. (5 -- 7).
C. Comparison with MC estimates and remarks
Our approach is based on the linear propagation of errors originating from many inde-
pendent sources (i.e. the Rk's). One can expect that this method will work reasonably well,
both because the input fractional uncertainties ∆ Rk/Rk are relatively small, and because
the final output uncertainties σi affecting the abundances Yi are "regularized" by the central
limit theorem. Indeed, our ±2σ bands in Fig. 1 compare well with the MC-estimated bands
of Refs. [17,21,23 -- 26], with small relative differences which depend, in part on different in-
put Rk ± ∆ Rk's, and that are not larger than the spread among the various MC estimates
themselves.
In order to be more quantitative, we compare in Fig. 4 (upper panel) the MC evaluation
of the fractional uncertainties σi/Yi as derived from Ref. [17]4 with our analytic estimate
(using, for this exercise, the same input parameters). There is good agreement between
these two totally independent estimates. In Fig. 4 (lower panel) we also show a comparison
with the only MC evaluation of ρij we are aware of (viz., Ref. [22]), obtaining again good
agreement with our calculation when the same input Rk ± ∆ Rk are used. We conclude the
discussion of Fig. 4 by noting that the uncertainty σ(2+3) and the correlations ρ(2+3)j related
to the often-used combination of abundances Y(2+3) = Y2 + Y3 = (D + 3He)/H are given,
within our approach, by:
σ2
(2+3) = σ2
3 + 2ρ23σ2σ3 ,
ρ(2+3)jσ(2+3)σj = ρ2jσ2σj + ρ3jσ3σj .
2 + σ2
(8)
(9)
There are, of course, some refined features of the MC approach that cannot be addressed
with our method, such as asymmetric or temperature-dependent uncertainties ∆ Rk [17].
However we consider these refinements not essential for practical applications. In a sense,
the possible asymmetry between "upper" and "lower" errors is where one wants it to be. For
4The MC values of σi/Yi have been read off the (small) panels of Fig. 27 in Ref. [17] and, therefore,
may be subject to small transcription errors.
5
instance, if one assumes a priori symmetric errors in the astrophysical S-factors, then asym-
metric errors are induced in the thermally averaged reaction rates R ∼ hσvi; conversely, the
requirement of a priori symmetric ∆ Rk errors requires that the input S-factor uncertainties
are readjusted, as discussed in Ref. [17]. Although the authors of Ref. [17] have adopted the
latter option ( + ∆ Rk = − ∆ Rk), the former option or others are equally acceptable,
and would clearly produce different outputs for the MC estimate of the abundance errors.
For instance, the upper and lower errors of Y7 appear to be rather symmetrical in the MC
calculation of Ref. [17], while they are noticeably asymmetrical in Ref. [22].
Concerning the temperature-dependent [17] uncertainties ∆ R7 and ∆ R10, which affect
mainly the estimate of Y7, our conservative choice in Table II proves to be successful for the
estimate of σ7/Y7 (see Fig. 4). This seems to indicate that the uncertainties at low temper-
atures (which are larger than those at high temperatures [17]) dominate in the estimate of
these errors. In any case, since practically all reaction rate uncertainties ∆ Rk contribute to
the final value of σ7, temperature-dependent refinements in the propagation of just two of
these (∆ R7 and ∆ R10) do not appear to be decisive for the estimate of the global error σ7.
In conclusion, we have shown that our simple analytic method for error evaluation repre-
sents an useful alternative to lengthy and computationally expensive MC simulations. Both
the magnitude and the correlations of the total errors affecting the theoretical abundances
are reproduced with good accuracy. We therefore advocate the use of this method for BBN
analyses as an alternative to MC simulations.5
III. DETERMINING THE LIKELY NUCLEON-TO-PHOTON RATIO
The comparison of the predicted primordial abundances Yi(η) ± σi(η) with their observa-
tionally inferred values Y i ± σi through a statistical test allows extraction of the likelihood
range for the fundamental parameter η. So far, this has been done either through fit-by-eye
(see, e.g., Ref. [17]) or by Monte Carlo-based maximum likelihood methods (see, e.g., Refs.
[25,26]). In this Section we show how limits on η can be simply extracted using χ2 statistics
based on the method described in the previous Section.
Assuming that the errors σi in the determinations of different abundances Y i are uncor-
related, the experimental squared error matrix σ2
ij is simply
σ2
ij = δijσiσj ,
(10)
where δij is Kronecker's delta. The total (experimental + theoretical) error matrix S2
then obtained by summing the matrices in Eqs. (5,10):
ij is
S2
ij(η) = σ2
ij(η) + σ2
ij .
(11)
5 A parallel situation holds in the field of solar neutrino physics, where the correlated uncertainties
of the neutrino fluxes predicted by solar models have been estimated through both Monte Carlo
simulations [32] and linear propagation of input errors [33]. The latter technique has proved more
popular because of its ease of use.
6
Its inverse defines the weight matrix Wij(η):
Wij(η) = [S2
ij(η)]−1.
(12)
The χ2 statistic associated with the difference between theoretical (Yi) and observational
(Yi) light element abundance determinations is then [29]:
χ2(η) = Xij
[Yi(η) − Yi] · Wij(η) · [Yj(η) − Yj] .
(13)
Minimization of the χ2 gives the most probable value of η, while the intervals defined by
χ2 = χ2
min + ∆ χ2 give the likely ranges of η at the confidence level set by ∆ χ2 (for one
degree of freedom, η).6
In order to illustrate this, we estimate η using recent observational data for the three light
element abundances (Y2, Y4, Y7) whose primordial origin is most secure. It is well known that
the observationally inferred values Y2 and Y4 are still controversial, and the conflict between
different determinations has driven a lively debate on the status of BBN (see Refs. [18,34]
and references therein). In this paper we do not enter into this debate but rather apply our
method to two possible (although mutually incompatible) selections of measurements which
we name data set "A" and data set "B":
Y2 = 1.9 ± 0.4 × 10−4 ,
Y4 = 0.234 ± 0.0054 ,
Y7 = 1.6 ± 0.36 × 10−10 ;
Data set A :
Y2 = 3.40 ± 0.25 × 10−5 ,
Data set B :
Y4 = 0.243 ± 0.003 ,
Y7 = 1.73 ± 0.12 × 10−10 .
(14)
(15)
The data set "A" is used in Ref. [26], the authors of which make a detailed MC-based
fit to η, thus enabling comparison with our method. Note that their adopted value of
the primordial deuterium abundance from observations of high redshift quasar absorption
systems is consistent with another recent observation, Y2 = (2.15 ± 0.35) × 10−4 [3], but in
conflict with the significantly smaller value reported in Ref. [2], which we adopt for data set
"B". Similarly the primordial helium mass fraction inferred from observations of metal-poor
blue compact galaxies which we adopt for data set "B" is from Ref. [7] in which it is argued
that previous analyses leading to the smaller value of Y2 used in data set "A" underestimate
the true abundance (although this is disputed in Ref. [6].) Finally the estimates for the
primordial lithium abundance in both data sets are based on observations of pop II stars,
with the slightly higher value [9] of Y7 in data set "B" taken from an updated analysis.
Readers who prefer to adopt different combinations of these, or indeed other, estimates for
6We remind the reader that ∆ χ2 = 1, 2.71, 3.84, and 6.64 correspond to confidence levels (C.L.'s)
of 68%, 90%, 95%, and 99%, respectively.
7
the primordial abundances are invited to perform their own fit to η by following the simple
prescription given here.
Before performing the χ2 fit, it is useful to get an idea of what one should expect by
comparing the data with the theoretical predictions at various values of η. Fig. 5 shows
the theoretical predictions for the abundances Y2, Y4, and Y7 in the three possible planes
(Yi, Yj), for representative values of x ≡ log10(η/10−10). The corresponding 1σ error ellipses
show clearly the size and the correlation of the "theoretical" errors. The observational data
sets "A" and "B" are also indicated on the figure, as crosses with 1σ error bars. Clearly the
former prefers η ∼ 2 × 10−10 while the latter favours η ∼ (4 -- 5) × 10−10.
A more precise estimate of the likely range of η is obtained, as anticipated, through
a χ2 fit. The results are shown in Fig. 6. The value of χ2
min is almost zero for the fit
to data set "A", indicating very good agreement between theory and observations, while
it is somewhat larger for data set "B". Note that the characteristic minimum in the 7Li
curve at η ≈ 2.6 × 10−10 (see bottom panel of Fig. 1) allows its measured abundance to be
compatible with both high D/low 4He (data set "A") or low D/high 4He (data set "B").
The 95% C.L. ranges allowed by each of the two data sets, as obtained by cutting the curves
at ∆χ2 = χ2 − χ2
min = 3.84, are:
Data set A : η = 1.78+0.54
Data set B : η = 5.13+0.72
−0.34 × 10−10 ,
−0.66 × 10−10 .
(16)
(17)
The range for case "A" agrees very well with the 95% C.L. range estimated in Ref. [26]
with the same inputs but with a different method (Monte Carlo + maximum likelihood). Of
course, the incompatibility between the above two ranges of η reflects the incompatibility
between the input abundance data within their stated errors.
IV. ROLE OF DIFFERENT REACTIONS IN LIGHT ELEMENT
NUCLEOSYNTHESIS
The role of the different nuclear reactions rates listed in Table II in the synthesis of
the light elements can be studied by "perturbing" the values of the input reaction rates
and observing their effect on the predicted abundances. More precisely, one can study the
contribution to the total uncertainty σi of Yi induced by a +1σ shift of Rk:
Rk → Rk + ∆ Rk =⇒ Yi → Yi + δ Yi .
(18)
Within our approach, this can be done very easily using Eq. (3), with the ∆ Rk's from
Table II. Of course, the results depend on the value chosen for η. To illustrate various
trends, we choose the best-fit values η = 1.78 × 10−10 and η = 5.13 × 10−10, corresponding
to data sets "A" and "B" respectively.
Figs. 7 and 8 show the deviations δ Yi (normalized to the total error σi) induced by
+1σ shifts in the Rk's, plotted in the same set of planes as used for Fig. 5. The 1σ error
ellipses shown in these figures are obtained by combining the deviation vectors δYi/σi in
an uncorrelated manner. Several interesting conclusions can be drawn from this exercise.
As expected, the uncertainty in the weak interaction rate R1 has the greatest impact on
8
Y4 for the high value of η (Fig. 8), since essentially all neutrons end up being bound in
4He. However at the lower value of η (Fig. 7), the uncertainty in R2 -- the "deuterium
bottleneck" -- plays an equally important role as R1 in determining Y4 because nuclear
burning is less complete here than at high η. Similarly with reference to the reaction rates
R7, R10 − R12 which synthesize 7Li, at low η it is the competition between R7 and R11
which largely determines Y7, while at high η it is the competition between R10 and R12. The
anticorrelation between Y4 and Y2 is driven mainly by R2 at low η and, to a lesser extent,
by R4 and R5, while the reverse is the case at high η. The anticorrelation between Y4 and
Y7 at low η is also basically driven by R2, while the correlation at high η is due to both
R2 and R4. Thus we have a direct visual basis for assessing in what direction the output
abundances Yi are pulled by possible changes in the input cross sections Rk.
V. CONCLUSIONS
We have shown that a simple method based on linear error propagation allows us to
quantify the uncertainties associated with the elemental abundances expected from big bang
nucleosynthesis, in excellent agreement with the results obtained from Monte Carlo simu-
lations. This method makes transparent which nuclear reaction rate is mainly responsible
for the uncertainty in the abundance of a given element. If determinations of the primor-
dial abundances improve to the point where the observational errors become smaller than
the theoretical uncertainties (say for 7Li), this will enable attention to be focussed on the
particular reaction rate whose value needs to be experimentally better known.
We have also demonstrated that for standard BBN, our method enables the use of simple
χ2 statistics to obtain the best-fit value of η from the comparison of theory and observations.
At present there are conflicting claims regarding the primordial abundances of, particularly,
D and 4He, and different choices of input data sets imply values of η differing by a factor of
∼ 3. However this quantity can also be determined through measurements of the angular
anisotropy of the cosmic microwave background (CMB) on small angular scales. Within a
decade the forthcoming all-sky surveyors MAP and PLANCK are expected to pinpoint the
nucleon density to within ∼ 5% [35]. Such measurements probe the acoustic oscillations
of the coupled photon-matter plasma at the (re)combination epoch and will thus provide
an independent check of BBN, assuming η did not change significantly between the two
epochs.7 Nevertheless precise measurements of light element abundances, particularly 4He,
are still crucial because they provide a unique probe of physical conditions, in particular
the expansion rate at the BBN epoch. To illustrate, if η was determined by the CMB
measurements to be ≈ 2 × 10−10 (consistent with data set "A"), but the abundance of 4He
was established to be actually closer to its higher value of ≈ 24% in data set "B", this
7New physics beyond the Standard Model can change η, e.g. by increasing the photon number
through massive particle decay [36] or, more exotically, by decreasing the photon number through
photon mixing with a shadow sector [37]. However such possibilities are strongly constrained by
the absence of distortions in the Planck spectrum of the CMB [38] and also, in the latter case, by
the absence of Sakharov oscillations in the power spectrum of large-scale structure [39].
9
would be a strong indication that the expansion rate during BBN was higher than in the
standard case with Nν = 3 neutrinos. Although the number of SU(2) doublet neutrinos is
indeed 3, there are many light particles expected in extensions of the Standard Model, e.g.
singlet neutrinos, which can speed up the expansion rate during nucleosynthesis [19]. The
generalization of our method to such non-standard cases is straightforward and we intend to
present these results in a future publication [40]. It is clear that BBN analyses will continue
to be important in this regard for both particle physics and cosmology.
ACKNOWLEDGMENTS
E.L. thanks the Department of Physics at Oxford University for hospitality during the
early stages of this work. S.S. and F.V. thank the organizers and participants in the In-
ternational Workshop on Synthesis of Light Nuclei in the Early Universe held at ECT,
Trento in June 1997 for useful discussions. We thank Geza Gyuk and Rob Lopez for very
helpful feedback. This work was supported by the EC Theoretical Astroparticle Network
CHRX-CT93-0120 (DG12 COMA).
10
TABLE I. The four light elemental abundances Yi considered in this work. Alternative symbols
used in the literature are indicated in parentheses.
TABLES
Symbol
Y4
Y2
Y3
Y7
... or
(Yp)
(y2p)
(y3p)
(y7p)
Definition
4He mass fraction
D/H (by number)
3He/H (by number)
7Li/H (by number)
TABLE II. The BBN reaction rates Rk and their 1σ uncertainties ±∆ Rk adopted in this work.
The numbering follows Ref. [16] while the reference "unit" values (Rk ≡ 1) correspond to the rates
in Ref. [17].
k
1
2
3
4
5
6
7
8
9
10
11
12
Reaction
n → p e ¯νe
p(n,γ)d
d(p,γ)3He
d(d,n)3He
d(d,p)t
t(d,n)4He
t(α,γ)7Li
3He(n,p)t
3He(d,p)4He
3He(α,γ)7Be
7Li(p,α)4He
7Be(n,p)7Li
Rk
0.9979
1
1
1
1
1
1
1
1
1
1
1
±∆ Rk
±0.0021
±0.07
±0.10
±0.10
±0.10
±0.08
±0.26
±0.10
±0.08
±0.16
±0.08
±0.09
TABLE III. Polynomial fit
abundances,
Yi = a0 + a1x + a2x2 + a3x3 + a4x4 + a5x5, with x ≡ log10(η/10−10) in the range 0 -- 1. The
abundances were obtained using the BBN computer code [16] with the input Rk's as in Table II.
The value of Y4 has been corrected using the prescription of Ref. [19]. The accuracy of the fit is
better than 1/25 of the total theoretical uncertainty for each Yi.
central value
elemental
to
the
of
the
Y2 × 103
Y3 × 105
Y4 × 101
Y7 × 109
a0
+0.4808
+3.4308
+2.2305
+0.5369
a1
−1.8112
−6.1701
+0.5479
−2.8036
a2
+3.2564
+8.1311
−0.6050
+7.6983
a3
−3.3525
−9.7612
+0.6261
−12.571
a4
+1.8834
+7.7018
−0.3713
+12.085
a5
−0.4458
−2.5244
+0.0949
−3.8632
11
Polynomial
fits
to
the
TABLE IV.
=
a0 + a1x + a2x2 + a3x3 + a4x4 + a5x5, as functions of x ≡ log10(η/10−10) ∈ [0, 1] (see Fig. 2).
In most cases, polynomials of degree < 5 provide sufficently accurate fits. Only non-negligible
logarithmic derivatives are tabulated.
logarithmic
derivatives,
λik(x)
λ2k
λ3k
λ4k
λ7k
k
1
2
3
4
5
8
9
1
2
3
4
5
6
8
9
1
2
4
5
1
2
3
4
5
6
7
8
9
10
11
12
a0
+0.7130
−0.7025
−0.0189
−0.4228
−0.4138
−0.0073
−0.0011
+0.0940
+0.0981
+0.0610
+0.3050
−0.5118
−0.0327
−0.5580
−0.1080
+0.8138
+0.0610
+0.0082
+0.0075
+1.9638
−0.9214
+0.0500
+0.1734
+0.1837
−0.9877
+0.9644
+0.0529
−0.0764
+0.0690
−1.4095
−0.0043
a1
−0.7964
+0.2611
−0.1879
−0.1698
−0.1477
−0.0003
−0.0348
+0.1892
+0.9948
+0.1640
+0.0805
+0.1274
+0.0829
−0.0287
−0.5089
−0.1465
−0.1962
−0.0058
−0.0058
+1.8520
−1.6472
−1.0433
−2.7428
−0.0875
−0.8168
+0.6888
+0.6318
+0.3017
−1.5360
−0.3543
−0.3779
a3
+0.0914
−0.8934
−0.6806
+0.0247
+0.0010
−0.0416
−0.0511
−0.0199
+3.4108
−0.2605
+0.1555
−0.2106
+0.0362
−0.8735
+0.8163
0
−0.1049
0
0
+27.542
+54.059
−2.4441
−10.736
−0.1350
−4.4380
+4.0284
+2.5754
+1.9311
−10.168
−4.7956
−42.390
a4
0
0
0
0
0
0
0
0
−1.2845
0
0
0
0
0
0
0
0
0
0
−11.041
−112.80
0
+2.5263
0
0
0
0
0
+2.9807
0
a5
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
+56.426
0
0
0
0
0
0
0
0
0
+61.401
−26.778
a2
+0.4577
+1.2008
+0.2502
+0.0207
+0.1103
+0.0801
+0.0592
−0.1484
−3.1667
+0.5368
−0.4208
+0.4081
−0.0939
+1.3574
−1.0157
+0.0408
+0.2416
+0.0034
+0.0034
−19.721
+5.6187
+4.1384
+11.209
+0.0158
+6.1555
−5.6151
−3.3995
−2.9632
+9.5808
+6.4780
+7.7358
12
REFERENCES
[1] D. N. Schramm and M. S. Turner, Rev. Mod. Phys. (Colloquia) 70, 303 (1998).
[2] S. Burles and D. Tytler, astro-ph/9712265, in Primordial Nuclei and their Galactic
Evolution, Proc. ISSI Workshop (Bern, 1997) edited by N. Prantzos, M. Tosi, and R.
von Steiger (Kluwer, Dordrecht), to appear.
[3] J. K. Webb, R. F. Carswell, K. M. Lanzetta, R. Ferlet, M. Lemoine, and A. Vidal-
Madjar, astro-ph/9710089, in Structure and Evolution of the Intergalactic Medium from
QSO Absorption Line Systems, Proceedings of the 13th IAP Workshop (Paris, 1997),
edited by P. Petitjean and S. Charlot (Nouvelles Fronti`eres, Paris), to appear.
[4] D. S. Balser, T. M. Bania, R. T. Rood and T. M. Wilson, Astrophys. J. 430, 667 (1994).
[5] G. Gloeckler and J. Geiss, Nature (London) 381, 210 (1996).
[6] K. A. Olive, G. Steigman and E. D. Skillman, Astrophys. J. 483, 788 (1997).
[7] Y. I. Izotov, T. X. Thuan, and V. A. Lipovetsky, Astrophys. J. Suppl. 108, 1 (1997).
[8] S. G. Ryan, T. C. Beers, C. P. Deliyannis and J. A. Thorburn, Astrophys. J. 458, 543
(1996).
[9] P. Bonifacio and P. Molaro, Mon. Not. Roy. Astron. Soc. 285, 847 (1997).
[10] P. Molaro, astro-ph/9709245, in From Quantum Fluctuations to Cosmological Struc-
tures, Proc. 1st Moroccan International Astronomy School (Casablanca, 1996), edited
by D. Valls-Gabaud et al., ASP Conf. Ser. 126, p. 103.
[11] C. Hogan, astro-ph/9712031, in Primordial Nuclei and their Galactic Evolution, Proc.
ISSI Workshop (Bern, 1997) edited by N. Prantzos, M. Tosi, and R. von Steiger (Kluwer,
Dordrecht), to appear.
[12] S. A. Levshakov, W. H. Kegel, and F. Takahara, astro-ph/9712136, in Toward the first
light of HDS , Proc. SUBARU HDS Workshop (Tokyo, 1997), to appear.
[13] J. A. Bernstein, L. S. Brown, and G. Feinberg, Rev. Mod. Phys. 61, 25 (1989).
[14] R. Esmailzadeh, G. D. Starkman, and S. Dimopoulos, Astrophys. J. 378, 504 (1991).
[15] R. V. Wagoner, Astrophys. J. 179, 343 (1973).
[16] L. Kawano, Report No. Fermilab-Pub-88/34-A, 1988 (unpublished); Report No.
Fermilab-Pub-92/04-A, 1992 (unpublished).
[17] M. S. Smith, L. H. Kawano, and R. A. Malaney, Astrophys. J. Suppl. 85, 219 (1993).
[18] K. A. Olive, astro-ph/9712160, in Proc. TAUP'97, 5th International Workshop on Top-
ics in Astroparticle and Underground Physics (Laboratori Nazionali del Gran Sasso,
Assergi, 1997) to appear.
[19] S. Sarkar, Rep. Prog. Phys. 59 (1996), 1493.
[20] L. M. Krauss and P. Romanelli, Astrophys. J. 358, 47 (1990).
[21] P. J. Kernan and L. M. Krauss, Phys. Rev. Lett. 72, 3309 (1994).
[22] P. J. Kernan and L. M. Krauss, Case Western Reserve University report No. CWRU-
P2-94, astro-ph/9402010. This report, published in an abridged version as Ref. [21],
contains additional numerical results relevant for comparison with our work.
[23] L. M. Krauss and P. J. Kernan, Astrophys. J. Lett. 432, 79 (1994); Phys. Lett. B 347,
347 (1995).
[24] C. J. Copi, D. N. Schramm, and M. S. Turner, Science 267, 192 (1995); Phys. Rev.
Lett. 75, 3981 (1995).
[25] N. Hata, R. J. Scherrer, G. Steigman, D. Thomas, T. P. Walker, S. Bludman, and P.
13
Langacker, Phys. Rev. Lett. 75, 3977 (1995); N. Hata, R. J. Scherrer, G. Steigman, D.
Thomas, and T. P. Walker, Astrophys. J. 458, 637 (1996); N. Hata, G. Steigman, S.
Bludman, and P. Langacker, Phys. Rev. D 55, 540 (1997).
[26] K. A. Olive and D. Thomas, Astropart. Phys. 7, 27 (1997). See also B. D. Fields and
K. A. Olive, Phys. Lett. B 368, 103 (1996); B. D. Fields, K. Kainulainen, K. A. Olive,
and D. Thomas, New Astron. 1, 77 (1996).
[27] R. M. Barnett et al., Phys. Rev. D 54, 1 (1996), and 1997 off-year partial up-
date for the 1998 edition available on the Particle Data Group WWW pages (URL:
http//pdg.lbl.gov/). This update includes the revised Penning-trap measurement of
Ref. [28].
[28] τn = 889.2 ± 4.8 s, J. Byrne, et al., Europhys. Lett. 33, 187 (1996). Here the authors
revise and replace their previous estimate (τn = 893.5 ± 5.3 s) which appeared in J.
Byrne et al., Phys. Rev. Lett. 65, 289 (1990).
[29] W. T. Eadie, D. Drijard, F. E. James, M. Roos, and B. Sadoulet, Statistical Methods
in Experimental Physics (North Holland, Amsterdam and London, 1971).
[30] P. J. Kernan, PhD thesis, Ohio State University (1993).
[31] R. Lopez, private communication.
[32] J. N. Bahcall and R. N. Ulrich, Rev. Mod. Phys. 60, 297 (1988); J. N. Bahcall and W.
C. Haxton, Phys. Rev. D 40, 931 (1989); X. Shi and D. N. Schramm, Phys. Lett. B
283, 305 (1992); L. M. Krauss, E. Gates, and M. White, Phys. Lett. B 299, 94 (1993).
[33] G. L. Fogli and E. Lisi, Astropart. Phys. 3, 185 (1995); N. Hata and P. Langacker, Phys.
Rev. D 50, 632 (1994); E. Gates, L. M. Krauss, and M. White, Phys. Rev. D 51, 2631
(1995).
[34] P. J. Kernan and S. Sarkar, Phys. Rev. D 54, R3681 (1996); S. Sarkar, astro-ph/9611232,
in Aspects of Dark Matter in Astro- and Particle Physics, edited by H. V. Klapdor-
Kleingrothaus and Y. Ramachers (World Scientific, 1997) p. 235.
[35] L. Knox, Phys. Rev. D 52 4307 (1995); G. Jungman, M. Kamionkowski, A. Kosowski,
and D. N. Spergel, Phys. Rev. D 54 1332 (1996); M. Zaldarriaga, D. N. Spergel and U.
Seljak, Astrophys. J. 488 1 (1997); J. R. Bond, G. Efstathiou and M. Tegmark, Mon.
Not. R. Astron. Soc. 291, L33 (1997).
[36] K. Sato and M. Kobayashi, Prog. Theor. Phys. 58, 1775 (1977); D. A. Dicus, E. W.
Kolb, V. L. Teplitz and R. V. Wagoner, Phys. Rev. D 17, 1529 (1978).
[37] J. G. Bartlett and L. Hall, Phys. Rev. Lett. 66, 541 (1991).
[38] J. Ellis, G. B. Gelmini, D. V. Nanopoulos, J. Lopez and S. Sarkar, Nucl. Phys. B 373,
399 (1992); W. Hu and J. Silk, Phys. Rev. Lett. 70, 2661 (1993).
[39] M. Birkel and S. Sarkar, Phys. Lett. B 408, 59 (1997). See also D. J. Eisenstein, W.
Hu, J. Silk and A. Szalay, Astrophys. J. Lett. 494, L1 (1998).
[40] G. Fiorentini, E. Lisi, S. Sarkar, and F. L. Villante, in preparation.
14
FIGURES
FIG. 1. Primordial abundances Yi (solid lines) and their ±2σ bands (dashed lines), as functions
of the nucleon-to-photon ratio η.
15
∂ ln Yi(h )
∂ ln Rk
FIG. 2. Logarithmic derivatives λik of the abundances Yi with respect to the reaction rates Rk,
as functions of η.
16
FIG. 3. Fractional abundance uncertainties σi/Yi (upper panel) and their correlations ρij (lower
panel), as functions of η.
17
FIG. 4. Monte Carlo estimates of σi/Yi (SKM '93 [17], dots) and ρij (KK '94 [22], dots),
compared with our analytic evaluation (solid lines), using the same inputs.
18
FIG. 5. Standard BBN predictions (dotted lines) in the 2-dimensional planes defined by the
abundances Y2, Y4, and Y7, as functions of x ≡ log10(η/10−10). The theoretical uncertainties are
depicted as 1σ error ellipses at x = 0, 0.1, 0.2, . . . , 1. The crosses indicate the two observational
data sets (with 1σ errors).
19
FIG. 6. Our χ2 fit to the data sets A and B, including observational and (correlated) theoretical
errors.
20
FIG. 7. Individual contributions of different reaction rates Rk to the uncertainties in Y2, Y4,
and Y7, normalized to the corresponding total errors σ2, σ4, and σ7, for η = 1.78 × 10−10. Each
arrow corresponds to the shift δ Yi induced by a +1σ shift of Rk. Some small error components
have not been plotted.
21
FIG. 8. Same as Fig. 7, but for η = 5.13 × 10−10.
22
|
astro-ph/0409282 | 5 | 0409 | 2006-03-09T20:31:11 | On the theory of canonical perturbations and its application to Earth rotation | [
"astro-ph",
"math.DS",
"physics.geo-ph"
] | Both orbital and rotational dynamics employ the method of variation of parameters. We express, in a non-perturbed setting, the coordinates (Cartesian, in the orbital case, or Eulerian in the rotation case) via the time and six adjustable constants called elements (orbital elements or rotational elements). If, under disturbance, we use this expression as ansatz and endow the "constants" with time dependence, then the perturbed velocity (Cartesian or angular) will consist of a partial derivative with respect to time and a so-called convective term, one that includes the time derivatives of the variable "constants." Out of sheer convenience, the so-called Lagrange constraint is often imposed. It nullifies the convective term and, thereby, guarantees that the functional dependence of the velocity upon the time and "constants" stays, under perturbation, the same as it used to be in the undisturbed setting.
When the dynamical equations, written in terms of the "constants," are demanded to be symplectic (and the "constants" make conjugated pairs $ Q, P$), these "constants" are called Delaunay elements, in the orbital case, or Andoyer elements, in the rotational case. The Andoyer and Delaunay sets of elements share a feature not visible with a naked eye: in certain cases, the standard equations render these elements non-osculating.
Hence, even though the Andoyer variables in the Kinoshita-Souchay theory are introduced in a precessing frame of the Earth orbit, they nevertheless return the angular velocity relative to an inertial frame. | astro-ph | astro-ph |
T alk at the conf erence "Journ´ees 2004: Syst`emes de r´ef ´erence spatio−temporels,"
l′Observatoire de P aris, 20 − 22 septembre 2004.
On the theory of canonical perturbations and its applica-
tion to Earth rotation. A source of inaccuracy in the
calculation of the angular velocity
US Naval Observatory, Washington DC 20392 USA,
e-mail: me @ usno.navy.mil
Michael Efroimsky
November 21, 2018
Abstract
Both orbital and rotational dynamics employ the method of variation of parameters. We
express, in a non-perturbed setting, the coordinates (Cartesian, in the orbital case, or Eule-
rian in the rotation case) via the time and six adjustable constants called elements (orbital
elements or rotational elements).
If, under disturbance, we use this expression as ansatz
and endow the "constants" with time dependence, then the perturbed velocity (Cartesian or
angular) will consist of a partial derivative with respect to time and a so-called convective
term, one that includes the time derivatives of the variable "constants." Out of sheer conve-
nience, the so-called Lagrange constraint is often imposed. It nullifies the convective term
and, thereby, guarantees that the functional dependence of the velocity upon the time and
"constants" stays, under perturbation, the same as it used to be in the undisturbed setting.
The variable "constants" obeying this condition are called osculating elements. Otherwise,
they are simply called orbital or rotational elements.
When the dynamical equations, written in terms of the "constants," are demanded to
be symplectic (and the "constants" make conjugated pairs Q, P ), these "constants" are
called Delaunay elements, in the orbital case, or Andoyer elements, in the rotational case.
The Andoyer and Delaunay sets of elements share a feature not readily apparent: in certain
cases, the standard equations render these elements non-osculating.
In orbital mechanics, the elements, calculated via the standard planetary equations, come
out non-osculating when perturbations depend on velocities. This complication often arises
but seldom gets noticed. To keep elements osculating under such perturbations, extra terms
must enter the equations, terms that will not be parts of the disturbing function (Efroimsky
& Goldreich 2003, 2004). In the case of parametrisation through the Kepler elements, this will
merely complicate the equations. In the case of Delaunay parametrisation, these extra terms
will not only complicate the Delaunay equations, but will also destroy their canonicity. Under
velocity-dependent disturbances, the osculation and canonicity conditions are incompatible.
Similarly, in rotational dynamics, the Andoyer elements come out non-osculating when
the perturbation depends upon the angular velocity of the top. Since a switch to a non-
inertial frame is an angular-velocity-dependent perturbation, then amendment of the dy-
namical equations by only adding extra terms to the Hamiltonian makes these equations
1
render non-osculating Andoyer elements. To make them osculating, extra terms must be
added to the dynamical equations (and then these equations will no longer be symplectic).
Calculations in terms of non-osculating variables are mathematically valid, but their
physical interpretation is problematic. Non-osculating orbital elements parametrise instan-
taneous conics not tangent to the orbit. Their inclination, the non-osculating i, may differ
much from the physical inclination of the orbit, given by the osculating i . Similarly, in the
case of rotation, non-osculating Andoyer variables do correctly describe a perturbed spin but
lack simple physical meaning. The customary expressions for the spin-axis' orientation, in
terms of the Andoyer elements, will no longer be valid, if the elements are non-osculating.
These expressions, though, will stay valid for osculating elements, but then the (correct)
dynamical equations for such elements will no longer be canonical -- circumstance ignored in
the Kinoshita-Souchay (KS) theory which tacitly employs non-osculating variables. While
the loss of osculation will not influence the predictions for the figure axis of the planet, it
considerably effects the predictions for the orientation of the instantaneous axis of rotation.
1 Kepler and Euler
In orbital dynamics, a Keplerian conic, emerging as an undisturbed two-body orbit, is regarded as
a sort of "elementary motion," so that all the other available motions are conveniently considered
as distortions of such conics, distortions implemented through endowing the orbital constants
Cj with their own time dependence. Points of the orbit can be contributed by the "elementary
curves" either in a non-osculating fashion, as in Fig. 1, or in the osculating way, as in Fig. 2.
The disturbances, causing the evolution of the motion from one instantaneous conic to another,
are the primary's oblateness, the gravitational pull of other bodies, the atmospheric and radiation-
caused drag, and the non-inertiality of the reference system.
Similarly, in rotational dynamics, a complex spin can be presented as a sequence of configu-
rations borrowed from a family of some elementary rotations. The easiest possibility here will be
to employ in this role the Eulerian cones, i.e., the loci of the rotational axis, corresponding to
non-perturbed spin states. These are the simple motions exhibited by an undeformable free top
with no torques acting thereupon.1 Then, to implement a perturbed motion, we shall have to go
from one Eulerian cone to another, just as in Fig. 1 and 2 we go from one Keplerian ellipse to
another. Hence, similar to those pictures, a smooth "walk" over the instantaneous Eulerian cones
may be osculating or non-osculating.
The physical torques, the actual triaxiality of the top, and the non-inertial nature of the refer-
ence frame will then be regarded as perturbations causing the "walk." The latter two perturbations
depend not only upon the rotator's orientation but also upon its angular velocity.
2 Delaunay and Andoyer
In orbital dynamics, we can express the Lagrangian of the reduced two-body problem via the
spherical coordinates qj = { r , ϕ , θ } , then calculate their conjugated momenta pj and the
Hamiltonian H(q, p) , and then carry out the Hamilton-Jacobi procedure (Plummer 1918), to
arrive to the Delaunay variables
{ Q1 , Q2 , Q3 ; P1 , P2 , P3 } ≡ { L , G , H ; l , g , h} =
1 Here one opportunity will be to employ in the role of "elementary" motions the non-circular Eulerian cones
described by the actual triaxial top, when this top is unforced. Another opportunity will be to use, as "elementary"
motions, the circular Eulerian cones described by a dynamically symmetrical top (and to treat its actual triaxiality
as another perturbation). The main result of our paper will be invariant under this choice.
2
{ √µa , qµa (1 − e2) , qµa (1 − e2) cos i
where µ denotes the reduced mass.
; − Mo , − ω , − Ω} ,
Similarly, in rotational dynamics one can define a spin state of a top by means of the three
Euler angles qj = ψ , θ , ϕ and their canonical momenta pj , and then perform a canonical trans-
formation to the Andoyer elements L , G , H , l , g , h . A minor technicality is that, historically,
these variables were introduced by Andoyer (1923) in a manner slightly different from the set of
canonical constants: while, for a free rotator, the three Andoyer variables G , H , h are constants,
the other three, L , l , g do evolve in time (for the Andoyer Hamiltonian of a free top is not zero,
(1)
The perturbed trajectory is a set of points belonging to a sequence of
Fig.1.
confocal instantaneous ellipses. The ellipses are not supposed to be tangent, nor
even coplanar to the orbit at the intersection point. As a result, the physical velocity
~r (tangent to the trajectory at each of its points) differs from the Keplerian velocity
~g (tangent to the ellipse). To parametrise the depicted sequence of non-osculating
ellipses, and to single it out of all the other such sequences, it is suitable to employ
the difference between ~r and ~g , expressed as a function of time and six (non-
osculating) orbital elements:
~Φ(t , C1 , . . . , C6) = ~r(t , C1 , . . . , C6) − ~g(t , C1 , . . . , C6) .
Since
~r =
∂~r
∂t
+
6
X
j=1
∂Cj
∂t
Cj = ~g +
6
X
j=1
∂Cj
∂t
Cj
,
then the difference ~Φ is simply the convective term P (∂~r/∂Cj) Cj which
emerges whenever the instantaneous ellipses are being gradually altered by the per-
turbation (and the orbital elements become time-dependent).
In the literature,
~Φ(t , C1 , . . . , C6) is called the gauge function or gauge velocity or, simply, gauge.
Fig.2. The perturbed trajectory is represented through a sequence of confocal
instantaneous ellipses which are tangent to the trajectory at the intersection points,
i.e., are osculating. Now, the physical velocity ~r (which is tangent to the trajectory)
will coincide with the Keplerian velocity ~g (which is tangent to the ellipse), so that
their difference ~Φ(t C1 , . . . , C6) vanishes everywhere:
~Φ(t , C1 , . . . , C6) ≡ ~r(t , C1 , . . . , C6) − ~g(t , C1 , . . . , C6) =
6
X
j=1
∂Cj
∂t
Cj = 0 .
This equality, called Lagrange constraint or Lagrange gauge, is the necessary and
sufficient condition of osculation.
3
but a function of l , L and G ). This way, to make our analogy complete, we may carry out
one more canonical transformation, from the Andoyer variables { L , G , H , l , g , h} to "almost
Andoyer" variables { Lo , G , H , lo , go , h} , where Lo , lo and go are the initial values of L , l
and g . The latter set consists only of the constants of integration; the corresponding Hamiltonian
becomes nil. Therefore, these constants are the true analogues of the Delaunay variables (while
the conventional Andoyer set is analogous to the Delaunay set with M used instead of Mo .). The
main result obtained below for the modified Andoyer set { Lo , G , H , lo , go , h} can be easily
modified for the regular Andoyer set of variables { L , G , H , l , g , h} . (Efroimsky 2005b)
To summarise this section, in both cases we start out with
q =
∂H(o)
∂p
,
p = −
∂H(o)
∂q
.
(2)
q and p being the coordinates and their conjugated momenta, in the orbital case, or the Euler
angles and their momenta, in the rotation case. Then we switch, via a canonical transformation
q = f (Q , P , t)
p = χ(Q , P , t) ,
to
Q =
∂H∗
∂P
= 0 ,
P = −
∂H∗
∂Q
= 0 , H∗ = 0 ,
(3)
(4)
This scheme relies on the fact that, for an unperturbed Keplerian orbit (and, similarly, for an
where Q and P denote the set of Delaunay elements, in the orbital case, or the (modified, as
explained above) Andoyer set { Lo , G , H , lo , go , h} , in the case of rigid-body rotation.
undisturbed Eulerian cone) its six-constant parametrisation may be chosen so that:
1.
zero Hamiltonian: H∗(Q, P ) = 0 ;
2.
the dynamical equations (2).
the parameters are constants and, at the same time, are canonical variables { Q , P } with a
for constant Q and P , the transformation equations (3) are mathematically equivalent to
3 When do the elements come out non-osculating?
3.1 General-type motion
Under perturbation, the "constants" Q , P begin to evolve so that, after their substitution into
q = f ( Q(t) , P (t) , t )
p = χ( Q(t) , P (t) , t )
(5)
(f and χ being the same functions as in (3) ), the resulting motion obeys the disturbed equations
q =
∂ (cid:16)H(o) + ∆H(cid:17)
∂p
,
p = −
∂ (cid:16)H(o) + ∆H(cid:17)
∂q
.
We also want our "constants" Q and P to remain canonical and to obey
Q =
∂ (H∗ + ∆H∗)
∂P
,
P = −
∂ (H∗ + ∆H∗)
∂Q
(6)
(7)
4
where
H∗ = 0
and ∆H∗ (Q , P t) = ∆H ( q(Q, P, t) , p(Q, P, t) , t )
.
(8)
Above all, an optimist will expect that the perturbed "constants" Cj ≡ Q1 , Q2 , Q3 , P1 , P2 , P3
(the Delaunay elements, in the orbital case, or the modified Andoyer elements, in the rotation
case) will remain osculating. This means that the perturbed velocity will be expressed by the
same function of Cj(t) and t as the unperturbed one used to. Let us check to what extent this
optimism is justified. The perturbed velocity reads
where
q = g + Φ
g(C(t), t) ≡
∂q(C(t), t)
∂t
is the functional expression for the unperturbed velocity; and
Φ(C(t), t) ≡
6
X
j=1
∂q(C(t), t)
∂Cj
Cj(t)
(9)
(10)
(11)
is the convective term. Since we chose the "constants" Cj to make canonical pairs (Q, P ) obeying
(7 - 8), with vanishing H∗ , then insertion of (7) into (22) will result in
∂∆H(q, p)
Qn(t) +
Pn(t) =
Φ =
(12)
.
3
X
n=1
∂q
∂Qn
3
X
n=1
∂q
∂Pn
∂p
So the canonicity demand is incompatible with osculation. In other words, whenever a momentum-
dependent perturbation is present, we still can use the ansatz (5) for calculation of the coordinates
and momenta, but can no longer use (14) for calculating the velocities. Instead, we must use (13).
Application of this machinery to the case of orbital motion is depicted on Fig.1. Here the constants
Cj = (Qn, Pn) parametrise instantaneous ellipses which, for nonzero Φ , are not tangent to the
trajectory. (For more details see Efroimsky & Goldreich (2003).) In the case of orbital motion,
the situation will be similar, except that, instead of the instantaneous Keplerian conics, one will
deal with instantaneous Eulerian cones (i.e., with the loci of the rotational axis, corresponding to
non-perturbed spin states).
3.2 Orbital motion
In the orbital-motion case, osculation means the following. Let the unperturbed position be given,
in some fixed Cartesian frame, by vector function ~f :
~r = ~f (C1 , . . . , C6 , t)
,
~r ≡ { x , y , z } .
Employing this functional ansatz also under disturbance, we get the perturbed velocity as
~r = ~g (C1(t) , . . . , C6(t) , t) + ~Φ (C1(t) , . . . , C6(t) , t)
where
∂ ~f
∂t
~g ≡
and
~Φ ≡
∞
X
j=1
∂ ~f
∂Cj
Cj
.
5
(13)
(14)
(15)
The osculation condition is a convenient (but totally arbitrary!) demand that the perturbed
velocity ~r has the same functional dependence upon t and Cj as the unperturbed velocity ~g :
~r (C1(t) , . . . , C6(t) , t) = ~f (C1(t) , . . . , C6(t) , t)
~r (C1(t) , . . . , C6(t) , t) = ~g (C1(t) , . . . , C6(t) , t)
,
.
or, equivalently, that the so-called Lagrange constraint is satisfied:
∞
X
j=1
∂ ~f
∂Cj
Cj = ~Φ (C1(t) , . . . , C6(t) , t) where ~Φ = 0 .
(16)
(17)
Fulfilment of these expectations, however, should in no way be taken for granted, because the
Lagrange constraint (17) and the canonicity demand (7 - 8) are now two independent conditions
whose compatibility is not guaranteed. As shown in Efroimsky (2002a,b), this problem has gauge
freedom, which means that any arbitrary choice of the gauge function ~Φ (C1(t) , . . . , C6(t) , t)
will render, after substitution into (13 - 14), the same values for ~r and ~r as were rendered by
Lagrange's choice (17).2 As can be seen from (12), the assumption, that the "constants" Q
and P are canonical, fixes the non-Lagrange gauge
∞
X
j=1
∂ ~f
∂Cj
Cj = ~Φ (C1(t) , . . . , C6(t) , t) where ~Φ =
∂∆H
∂~p
.
(18)
It is easy to show (Efroimsky & Goldreich 2003; Efroimsky 2005b) that this same non-Lagrange
gauge simultaneously guarantees fulfilment of the momentum-osculation condition:
~r (C1(t) , . . . , C6(t) , t) = ~f (C1(t) , . . . , C6(t) , t)
~p (C1(t) , . . . , C6(t) , t) = ~g (C1(t) , . . . , C6(t) , t)
,
.
(19)
Any gauge different from (18), will prohibit the canonicity of the elements.
In particular, for
momentum-dependent ∆H , the choice of osculation condition ~Φ = 0 would violate canonicity.
For example, an attempt of a Hamiltonian description of orbits about a precessing oblate
primary will bring up the following predicament. On the one hand, it is most natural and conve-
nient to define the Delaunay elements in a co-precessing (equatorial) coordinate system. On the
other hand, these elements will not be osculating in the frame wherein they were introduced, and
therefore their physical interpretation will be difficult, if at all possible. Indeed, instantaneous
ellipses on Fig.1 may cross the trajectory at whatever angles (and may be even perpendicular
thereto). Thence, their orbital elements will not describe the real orientation or shape of the
physical trajectory (Efroimsky & Goldreich 2004; Efroimsky 2005a).
For the first time, non-osculating elements obeying (22) implicitly emerged in (Goldreich 1965)
and then in Brumberg et al (1971), though their exact definition in terms of gauge freedom was
not yet known at that time. Both authors noticed that these elements were not osculating.
Brumberg (1992) called them "contact elements." The osculating and contact variables coincide
when the disturbance is velocity-independent. Otherwise, they differ already in the first order of
the velocity-dependent perturbation. Luckily, in some situations their secular parts differ only in
the second order (Efroimsky 2005a), a fortunate circumstance anticipated yet by Goldreich (1965).
2 Physically, this simply means,
~r on Fig.1 can be decomposed into ~g and ~Φ in a continuous variety of ways.
Mathematically, this freedom reflects a more general construction that emerges in the ODE theory. (Newman &
Efroimsky 2003)
6
3.3 Rotational motion
In rotational dynamics, the situation of an axially symmetric unsupported top at each instant of
time is fully defined by the three Euler angles qn = θ , φ , ψ and their derivatives qn = θ ,
ψ.
The time dependence of these six quantities can be calculated from three dynamical equations of
the second order and will, therefore, depend upon the time and six integration constants:
φ ,
qn = fn (S1 , . . . , S6 , t)
qn = gn (S1 , . . . , S6 , t)
,
,
(20)
the functions gn and fn being interconnected via gn ≡ ∂fn/∂t , for n = 1 , 2 , 3 = ψ , θ , φ.
Under disturbance, the motion will be altered:
qn = fn (S1(t) , . . . , S6(t) , t)
,
qn = gn (S1(t) , . . . , S6(t) , t) + Φn (S1(t) , . . . , S6(t) , t)
,
where
Φn (S1(t) , . . . , S6(t) , t) ≡
6
X
j=1
∂fn
∂Sj
Sj
.
(21)
(22)
Now choose the "constants" Sj to make canonical pairs (Q, P ) obeying (7 - 8), with H∗ being
zero for (Q, P ) = (Lo , G , H , lo , go , h) . Then insertion of (7) into (22) will result in
Φn (S1(t) , . . . , S6(t) , t) ≡ X
∂fn
∂Q
Q + X
∂fn
∂P
P =
∂∆H(q, p)
∂pn
,
(23)
so that the canonicity demand (7 - 8) violates the gauge freedom in a non-Lagrange fashion. This
is merely a particular case of (12).
This yields two consequences. One is that, in the canonical formalism, calculation of the
angular velocities via the elements must be performed not through the second equation of (20)
but through the second equation of (21), with (23) substituted therein. This means, for example,
that in Kinoshita (1977) expressions (2.6) and (6.24 - 6.27) render not the angular velocity of
the Earth relative to the precessing frame (wherein the Andoyer variables were defined) but
the angular velocity relative to an inertial frame. (For an extended explanation of this fact see
Efroimsky 2005b.) This, however, is not a drawback of the Kinoshita-Souchay theory but rather
its advantage, because it is the angular velocity relative to an inertial frame that is directly
measurable. (Schreiber et al 2004)
The second consequence is that, if we wish to make our Andoyer variables osculating (so that
the second equation of (20) could be used), the price to be payed for this repair will be the loss of
canonicity. (Angular-velocity-dependent disturbances cannot be accounted for by merely amend-
ing the Hamiltonian!) The osculating elements will obey non-canonical dynamical equations.
To draw to a close, we would add that, under some special circumstances, the secular parts
of contact elements may coincide in the first order with those of their osculating counterparts.3
Whether this will be the case for the Earth or Mars remains to be investigated. This matter
will be crucial for examining the validity of the presently available computations of the history of
Mars' obliquity.4
3 In regard to orbital motions, this possibility was anticipated yet in 1965 by Peter Goldreich. As demonstrated
by Efroimsky (2005a), this is true for constant rate of frame precession (but not for variable precession).
4 The pioneer study on this topic was conducted by Ward (1973) in a direct manner and was, therefore, exempt
from the problems associated with the loss of osculation. However, some of his successors chose the canonical
formalism and exploited the Hamiltonian borrowed from the Kinoshita theory.
7
4 Conclusions
In this talk we have explained why the Hamiltonian theory of Earth rotation renders non-osculating
Andoyer elements. We have also explained how this defect of the theory should be mended.
In attitude mechanics, osculation loss has the same consequences as in the theory of orbits:
while this defect of the theory has no influence upon the theory's predictions for the figure axis
of the planet, it considerably effects the predictions for the orientation of the instantaneous axis
of rotation.
In our paper Efroimsky (2005b) we shall demonstrate that, even though the Andoyer variables
in the Kinoshita-Souchay theory are introduced in the precessing frame of the Earth orbit, they
return the angular velocity not relative to that frame, but relative to an inertial one. This is not
a drawback of this theory but rather its advantage, because it is the angular velocity relative to
an inertial frame that is directly measurable at present. (Schreiber et al 2004)
References
[1] Andoyer, H. 1923. Cours de M´ecanique C´eleste, Paris. Gauthier-Villars 1923.
[2] Brumberg, V. A., L. S. Evdokimova, & N. G. Kochina. 1971. "Analytical Methods for the
Orbits of Artificial Satellites of the Moon." Celestial Mechanics, Vol. 3, pp. 197 - 221.
[3] Brumberg, V.A. 1992. Essential Relativistic Celestial Mechanics. Adam Hilger, Bristol.
[4] Efroimsky, Michael. 2002a. "Equations for the orbital elements. Hidden symmetry." Preprint
No 1844 of the Institute of Mathematics and its Applications, University of Minnesota
http://www.ima.umn.edu/preprints/feb02/feb02.html
[5] Efroimsky, Michael. 2002b. "The Implicit Gauge Symmetry Emerging in the N-body Problem
of Celestial Mechanics."
astro-ph/0212245
[6] Efroimsky, M., & P. Goldreich. 2003. "Gauge Symmetry of the N-body Problem in the
Hamilton-Jacobi Approach." Journal of Mathematical Physics, Vol. 44, pp. 5958 - 5977
astro-ph/0305344
[7] Efroimsky, M., & Goldreich, P. 2004. "Gauge Freedom in the N-body Problem of Celestial
Mechanics." Astronomy & Astrophysics, Vol. 415, pp. 1187 - 1199
astro-ph/0307130
[8] Efroimsky, M. 2005a. "Long-term evolution of orbits about a precessing oblate planet. The
case of uniform precession." Celestial Mechanics and Dynamical Astronomy, Vol.91, pp.75-108
astro-ph/0408168
[9] Efroimsky, M. 2005b. "The theory of canonical perturbations applied to attitude dynamics
and to the Earth rotation."
astro-ph/0506427
[10] Goldreich, P. 1965. "Inclination of satellite orbits about an oblate precessing planet." The
Astronomical Journal, Vol. 70, pp. 5 - 9.
[11] Kinoshita, H. 1977. "Theory of the Rotation of the Rigid Earth." Celestial Mechanics, Vol.
15, pp. 277 - 326
8
[12] Newman, W., & M. Efroimsky. 2003. "The Method of Variation of Constants and Multiple
Time Scales in Orbital Mechanics." Chaos, Vol. 13, pp. 476 - 485.
[13] Plummer, H. C. 1918. An Introductory Treatise on Dynamical Astronomy. Cambridge Uni-
versity Press, UK.
[14] Schreiber, K. U.; Velikoseltsev, A.; Rothacher, M.; Klugel, T.; Stedman, G. E.; & Wiltshire,
D. L. 2004. "Direct Measurement of Diurnal Polar Motion by Ring Laser Gyroscopes." Journal
of Geophysical Research, Vol. 109, p. B06405
[15] Ward, W. 1973. "Large-scale Variations in the Obliquity of Mars." Science. Vol. 181, pp. 260
- 262.
9
|
astro-ph/0407187 | 1 | 0407 | 2004-07-09T05:56:07 | Does regenerated emission change the high-energy signal from gamma-ray burst afterglows? | [
"astro-ph"
] | We study regenerated high-energy emission from the gamma-ray burst (GRB) afterglows, and compare its flux with the direct component from the same afterglow. When the intrinsic emission spectrum extends to TeV region, these very high-energy photons are significantly absorbed by the cosmic infrared background (CIB) radiation field, creating electron/positron pairs; since these pairs are highly energetic, they can scatter the cosmic microwave background radiation up to GeV energies, which may change the intrinsic afterglow light curve in the GeV region. Using the theoretical modeling given in literature and the reasonable choice of relevant parameters, we calculate the expected light curve due to the regeneration mechanism. As the result, we find that the regenerated emission could only slightly change the original light curve, even if we take a rather large value for the CIB density, independently of the density profile of surrounding medium, i.e., constant or wind-like profile. This ensures us the reliable estimation of the intrinsic GRB parameters when the high-energy observation is accessible, regardless of a large amount of uncertainty concerning the CIB density as well as extragalactic magnetic field strength. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 6 (2004)
Printed 6 November 2018
(MN LATEX style file v2.2)
Does regenerated emission change the high-energy signal
from gamma-ray burst afterglows?
Shin'ichiro Ando⋆
Department of Physics, School of Science, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
Submitted 9 June 2004; accepted 8 July 2004
ABSTRACT
We study regenerated high-energy emission from the gamma-ray burst (GRB) after-
glows, and compare its flux with the direct component from the same afterglow. When
the intrinsic emission spectrum extends to TeV region, these very high-energy pho-
tons are significantly absorbed by the cosmic infrared background (CIB) radiation
field, creating electron/positron pairs; since these pairs are highly energetic, they can
scatter the cosmic microwave background radiation up to GeV energies, which may
change the intrinsic afterglow light curve in the GeV region. Using the theoretical
modeling given in literature and the reasonable choice of relevant parameters, we cal-
culate the expected light curve due to the regeneration mechanism. As the result, we
find that the regenerated emission could only slightly change the original light curve,
even if we take a rather large value for the CIB density, independently of the density
profile of surrounding medium, i.e., constant or wind-like profile. This ensures us the
reliable estimation of the intrinsic GRB parameters when the high-energy observation
is accessible, regardless of a large amount of uncertainty concerning the CIB density
as well as extragalactic magnetic field strength.
Key words: gamma-rays: bursts -- diffuse radiation -- magnetic fields.
1
INTRODUCTION
Gamma-ray bursts (GRBs) are known to be highly energetic
astrophysical objects located at cosmological distance. Ac-
cumulated data of many GRBs strongly support relativistic
fireball scenario, in which γ-rays up to MeV are attributed
to internal shocks due to collisions between fireball shells,
while their transient component, afterglow, from radio to
X-rays is attributed to external shocks due to the interac-
tion of the fireball with external medium. In addition to
such signals, very high-energy photons that range from a
few tens of MeV to GeV have been detected (Hurley et al.
1994), and further the detection of an excess of TeV pho-
tons from GRB 970417a has been claimed with a chance
probability ∼ 1.5 × 10−3 (Atkins et al. 2000). Although the
statistics of these high-energy signals are not sufficient yet,
planned future satellites or detectors will promisingly enable
us to discuss high-energy emission mechanisms of GRBs.
Several emission mechanisms of GeV -- TeV photons are
proposed, such as synchrotron self-inverse Compton (IC)
emission of the electrons (M´esz´aros, Rees & Papathanassiou
1994; Waxman
1998;
Wei & Lu 1998, 2000; Dermer, Bottcher & Chiang 2000a;
Dermer, Chiang & Mitman 2000b; Panaitescu & Kumar
1997; Panaitescu & M´esz´aros
⋆ E-mail: [email protected]
Sari & Esin
2001;
the
proton-synchrotron
emission
Zhang & M´esz´aros
(Vietri
2001)
2000;
1997;
and
as well
as
Bottcher & Dermer
components
some
(Bottcher & Dermer
could
be valid for internal shocks, external forward shocks, or
external reverse shocks of GRBs.
1998). These mechanisms
1998),
emission
1998; Totani
other
hadron-related
Regardless of the emission mechanism, high-energy pho-
tons above ∼ 100 GeV are expected to be attenuated via the
γγ → e+e− process. Target photons with which the initial
high-energy photons interact are the GRB emission itself or
the cosmic infrared background (CIB). As for the latter, it is
suggested that such very high-energy photons may largely be
absorbed during its propagation, if the GRB location is suffi-
ciently cosmological as z & 1 (Stecker, de Jager & Salamon
1992; MacMinn & Primack 1996; Madau & Phinney 1996;
Malkan & Stecker 1998; Salamon & Stecker 1998). There-
fore, the detection of TeV photons from cosmological
GRBs will be very difficult. However, since created elec-
tron/positron pairs due to the interaction with CIB pho-
tons are very energetic, they can IC scatter on the most nu-
merous cosmic microwave background (CMB) photons, giv-
ing rise to a delayed secondary MeV -- GeV emission (Plaga
1995; Cheng & Cheng 1996; Dai & Lu 2002; Wang et al.
2004; Razzaque, M´esz´aros & Zhang 2004). These regener-
ated emissions would, therefore, be indirect evidence of
2
S. Ando
the intrinsic TeV emission as well as a probe of the CIB
radiation field, which is not satisfactorily constrained. It
has been argued that the internal shocks (Dai & Lu 2002;
Razzaque et al. 2004) as well as the prompt phase of the
external shocks (Wang et al. 2004) are possibly responsible
for these delayed MeV -- GeV emission, and it would be dis-
tinguishable from a different delayed GeV component due
to the direct IC emission from the afterglow.
Future detection of the direct MeV -- GeV emission pre-
dicted from the afterglow, would be a probe of an emission
mechanism of the high-energy region as well as physical pa-
rameters of the fireball. However, the primary emission is
possibly modified when we consider the regenerated light
due to an absorption of TeV photons by CIB, and esti-
mating its flux is a nontrivial problem; if the regenerated
light significantly changes the high-energy emission profile,
it gives a quite large amount of uncertainty on the GRB
physics since the CIB background as well as extragalactic
magnetic field strength, both of which are not satisfactorily
constrained yet, alters the expected signal. In this paper,
therefore, we investigate the effects of the delayed emission
on the primary afterglow light curve in the GeV range; as a
source of the regenerated GeV emission, we consider an af-
terglow phase itself, which is well described by an external
shock model, while the other phases have already been inves-
tigated in several papers (Dai & Lu 2002; Wang et al. 2004;
Razzaque et al. 2004). Among several mechanisms that pre-
dict an afterglow spectrum extending to TeV range, we
adopt the IC scattering of the synchrotron photons and fol-
low the formulation given by Zhang & M´esz´aros (2001) and
Sari & Esin (2001); this is because the high-energy emis-
sion up to TeV region is most likely realized due to the IC
mechanism (Zhang & M´esz´aros 2001), and the required val-
ues for relevant parameters appear to be realized in many
GRBs (Panaitescu & Kumar 2002). We show that the regen-
erated GeV emission, due to the TeV absorption and follow-
ing CMB scattering, can only slightly changes the detected
signals. Therefore, we conclude that the GeV light curves
obtained by the future detectors such as the Gamma-Ray
Large Area Space Telescope (GLAST) surely give us intrin-
sic information concerning the GRB fireball, not affected
by the uncertainty of the CIB as well as the extragalactic
magnetic fields.
This paper is organized as follows. In § 2, we briefly sum-
marize the formulation of the prompt high-energy emission
given in literature, and then in § 3, we describe the regen-
eration mechanism from the prompt high-energy emission
and give formulation for that. The result of the numerical
calculation using a reasonable parameter set is presented in
§ 4, and finally, we discuss that result and give conclusions
in § 5.
2 HIGH-ENERGY RADIATION FROM
AFTERGLOWS
As for evolution of the fireball and radiation spectrum, we
follow the formulation given in Zhang & M´esz´aros (2001)
and Sari & Esin (2001), and refer the reader to the liter-
ature for a detailed discussion; here we briefly summarize
necessary information.
The spectrum of the afterglow synchrotron emission
i.e., the self-absorption frequency νa,
(Sari, Piran & Narayan 1998) has breaks at several frequen-
cies,
injection fre-
quency νm corresponding to the minimum electron Lorentz
factor γm, cooling frequency νc corresponding to the electron
Lorentz factor γc for which the radiative timescale equals
the dynamical time, and cutoff frequency νu corresponding
to γu above which electrons cannot be accelerated. Break
frequencies relevant for this study and peak flux are given
by
52 t−3/2
νm = 2.9 × 1016 Hz ǫ2
νc = 3.1 × 1013 Hz (1 + Ye)−2ǫ−3/2
B E 1/2
eǫ1/2
h
(1 + z)1/2,
52 n−1
−1/2
E
B
× t−1/2
h
(1 + z)−1/2,
νu = 2.3 × 1022 Hz (1 + Ye)−1E 1/8
52 n−1/8t−3/8
h
× (1 + z)−5/8,
Fν,max = 29 mJy ǫ1/2
B E52n1/2D−2
L,28(1 + z),
(1)
(2)
(3)
(4)
where z is the redshift of the GRB, E52 is the fireball en-
ergy per unit solid angle in units of 1052 ergs sr−1, n the
external medium density in units of cm−3, ǫe and ǫB rep-
resent the fraction of the kinetic energy going to the elec-
trons and magnetic fields, respectively, th is the observer
time measured in hours, and DL,28 is the burst luminos-
ity distance measured in units of 1028 cm. The Compton
parameter Ye is given as a ratio between the luminosities
due to IC and synchrotron radiation, and can be repre-
sented by Ye = LIC/Lsyn = [−1 + (1 + 4ηǫe/ǫB)1/2]/2,
where η = min[1, (γm/γc)p−2] with p representing the spec-
tral index of injected electrons (Panaitescu & Kumar 2000;
Sari & Esin 2001).
m ≃ γ2
uνu, ν IC
c ≃ γ2
u = min(γ2
KN), where ν IC
The IC spectrum due to the scattering on the syn-
chrotron seed photons can be very hard if the Compton
parameter Ye is sufficiently large. The typical break fre-
quencies that characterize the IC spectrum are ν IC
mνm
and ν IC
c νc. The cutoff frequency in the IC compo-
nent is defined by ν IC
KN is
the Klein-Nishina limit, above which the IC cross section is
suppressed. Sari & Esin (2001) explicitly gave analytic ex-
pressions for the IC spectrum, and they pointed out that
the power-law approximation is no longer accurate at high-
frequency region, on which we focus in this paper. Therefore,
we use their analytic expressions shown in Appendix A of
Sari & Esin (2001), on the contrary to Zhang & M´esz´aros
(2001), in which the authors used an power-law expression
for simplicity. High-energy photons reaching to TeV due to
the IC scatterings are absorbed by the soft photons in the
fireball and create the electron/positron pairs. We follow the
treatment of Zhang & M´esz´aros (2001) for this intrinsic ab-
sorption (see also, Lithwick & Sari 2001; Coppi & Blandford
1990; Bottcher & Schlickeiser 1997; Dermer et al. 2000b).
Until this point, we described the emission property
in the case that the density profile of surrounding mat-
ter is uniform, i.e., n does not depend on the radius. We
also consider the case of the wind density profile, which is
possibly the case because the GRB progenitors can eject
envelope as a stellar wind; assuming a constant speed of
the wind, the density profile becomes n(r) = Ar−2, where
A is a constant independent of radius r. We normalize
this constant A as A = 3.0 × 1035A∗ cm−1 where A∗ =
( M /10−5M⊙ yr−1)/(v/103 km s−1) as in Chevalier & Li
(2000) for a Wolf-Rayet star. The relevant frequencies and
the peak flux are then be represented by
52 t−3/2
νm = 2.8 × 1016 Hz ǫ2
νc = 2.4 × 1011 Hz (1 + Ye)−2ǫ−3/2
B E 1/2
eǫ1/2
h
B
(1 + z)1/2,
E 1/2
52 A−2
∗
× t1/2
h (1 + z)−3/2,
νu = 1.3 × 1022 Hz (1 + Ye)−1E 1/4
52 A−1/4
∗
t−1/4
h
Fν,max = 0.33 Jy ǫ1/2
52 A∗t−1/2
h D−2
L,28(1 + z)3/2,
× (1 + z)−3/4,
B E 1/2
(5)
(6)
(7)
(8)
following the discussion given in Zhang & M´esz´aros (2001),
which is applied to the case of wind profile. Both the syn-
chrotron and IC spectra at some fixed time are obtained by
using the same procedure already given above, but the time
evolution of these spectra changes since the dynamics of an
expanding jet differs from the case of constant medium. Al-
though we do not give full representation of the spectral evo-
lution, the reader is referred to Panaitescu & Kumar (2000)
for analytic treatment including the wind-like structure.
3
INTERACTION WITH COSMIC INFRARED
BACKGROUND AND REGENERATED
HIGH-ENERGY EMISSION
at
redshift
typical GRB locations
z = 1,
For
Salamon & Stecker (1998) indicated that the optical depth
due to the CIB radiation field reaches ∼ 10 when the en-
ergy of prompt emission is higher than 300 GeV. We assume
that the electron and positron of the e± pair share 1/2 the
photon energy, i.e., γe = ǫγ /2me. With this assumption, the
created electron/positron spectrum can be described by
d2Ne
dtpdγe
= 2
dν
dγe
Fν (tp, ǫγ )
ǫγ
=
2
hγe
Fν (tp, 2meγe),
(9)
where ǫγ = hν, tp represents the observed time of the
prompt emission provided that there is no absorption
by the CIB, and Fν is the flux of prompt photons in-
cluding the intrinsic absorption. Following the result of
Salamon & Stecker (1998), we assume that the high-energy
photons with ǫγ > 300 GeV are completely attenuated,
creating e± pairs with γe > 3 × 105. The pair creations
typically occur at the distance Rpair = (0.26σT nIR)−1 ≃
5.8 × 1024 cm−3(nIR/1 cm−3)−1, where nIR is the CIB num-
ber density; this length scale is much less than the distance
from the observer to the GRB, DL, and hence, the attenu-
ation of the primary photons can be regarded as quite local
phenomenon.
The secondary electron/positron pairs then IC scat-
ter the CMB photons up to GeV energy scale, and pairs
cool on a timescale tIC = 3mec/(4γeσT ucmb(z)) ≈ 7.3 ×
1013(γe/106)−1(1 + z)−4 s in the local rest frame, where
ucmb(z) represents the CMB energy density at redshift z.
The IC spectrum from an electron (positron) with the
Lorentz factor γe, scattering on a CMB photon whose
energy is ǫcmb, d3Nγ /dǫcmbdt′
ddEγ , is explicitly given by
Blumenthal & Gould (1970) to be
d3Nγ
dǫcmbdt′
ddEγ
=
πr2
0c
2γ2
e
ncmb(ǫcmb, z)
ǫ2
cmb
(cid:18)2Eγ ln
Eγ
e ǫcmb
+Eγ + 4γ2
e ǫcmb −
(10)
4γ2
E2
γ
e ǫcmb(cid:19) ,
2γ2
Regenerated emission from GRB afterglows
3
where ncmb(ǫcmb, z) is the number spectrum of the CMB at
redshift z, t′
d represents the time of the delayed emission in
the local rest frame, measured from the onset of the e± pair
generation, and Eγ the energy of the delayed γ-ray. Since the
observed time of the delayed emission t can be represented
by t = tp + td, where td is the observed time of the delayed
emission measured from the pair generation, the flux of the
regenerated γ-ray is obtained by
Fν(t, Eγ) = Z t
0
where
d3Ndelayed IC
dtpdtddEγ
×(cid:18)
d3Nγ
dǫcmbdt′
dtp Eγ
d3Ndelayed IC
dtpdtddEγ (cid:12)(cid:12)(cid:12)(cid:12)td=t−tp
= Z dǫcmbZ dγe(cid:18) d2Ne
dtpdγe(cid:19)
ddEγ(cid:19) tIC(γe)
e−td/∆t(γe)
∆t(γe)
,
,
(11)
(12)
which can be calculated with the previously evaluated spec-
tra (eqs. [9] and [10]). Here, the lower bound of the inte-
gration over γe is max[3 × 105, (Eγ /ǫcmb)1/2/2]. In equation
(12), (d3Nγ /dǫcmbdt′
ddEγ)tIC shows a total number of the
IC photons per unit CMB energy per unit IC photon en-
ergy, emitted until the parent electron with γe cools. The
last part of the same equation e−td/∆t/∆t represents the
time profile of the delayed γ-ray emission, and ∆t(γe) is the
typical observed duration of the IC photons from the elec-
tron (positron) with Lorentz factor γe. The typical timescale
of the delayed emission measured in the observer frame is
given by ∆t(γe) = max(∆tIC, ∆tA, ∆tB), where ∆tIC = (1+
z)tIC/2γ2
e c is
the angular spreading time; and ∆tB = (1+z)tICθ2
B/2 is the
delay time due to magnetic deflection. The deflection angle
θB is given by θB ≈ 1.3 × 10−5(γe/106)−2BIG,−20, where
BIG,−20 represents the extragalactic magnetic field strength
in units of 10−20 G.
e is the IC cooling time; ∆tA = (1 + z)Rpair/2γ2
4 RESULTS
We calculated the expected high-energy signal in the GeV
range due to the prompt and regenerated afterglow emis-
sions using equation (11) as well as the formulation given
by Zhang & M´esz´aros (2001) and Sari & Esin (2001). In the
following discussion, we fix several relevant parameters used
in our calculation as follows: E52 = 10, n = 1, A∗ = 1, z = 1,
and BIG,−20 = 1.
ν1
R ν2
As for the parameters ǫe and ǫB, we first fix them
at 0.5 and 0.01, respectively. Figure 1(a) shows a fluence
Fν dν t as a function of observed time t, where hν1 = 400
MeV and hν2 = 200 GeV; the lower three curves represent
the fluence of the regenerated emission in the case of nIR = 1
(dashed curve), 0.1 (solid curve), and 0.01 cm−3 (dot-dashed
curve). Total fluence from the prompt and delayed afterglow
components are shown as upper three (almost degenerate)
curves using the same line type according to the CIB density.
The fluence threshold for the GLAST satellite is roughly
∼ 4 × 10−7(t/105 s)1/2 ergs cm−2 for a long integration
time regime (exposure time t & 105 s) and ∼ 4 × 10−7 ergs
cm−2 for a short integration time, following the criterion
that at least 5 photons are collected (Gehrels & Michelson
4
S. Ando
ν1
Figure 1. (a) FluenceR ν2
Fν dν t as a function of observed time t,
integrated over 400 MeV to 200 GeV. Lower three curves show the
fluence of the regenerated emission in the case of nIR = 1 (dashed
curve), 0.1 (solid curve), and 0.01 cm−3 (dot-dashed curve). Up-
per three (almost degenerate) curves show the total fluence. The
values for the relevant parameters are: E52 = 10, n = 1, ǫe =
0.5, ǫB = 0.01, z = 1, and BIG,−20 = 1. The sensitivity curve of
the GLAST satellite is also shown. (b) Flux ratio of the delayed
and prompt afterglow emission, integrated over the same energy
range.
Figure 2. The same as figure 1, but for various sets of (ǫe, ǫB)
with fixed value of nIR = 0.1 cm−3. Total fluence alone is shown
in the upper panel (a).
1999; Zhang & M´esz´aros 2001); this sensitivity curve is also
plotted in the same figure as a thin solid line.
Figure 1(b) shows a ratio of the regenerated and prompt
flux integrated over the same energy range. From these fig-
ures, it is found that the contribution from the regenerated
GeV emission due to the absorption by the CIB photons
peaks around 10 -- 104 s after the onset of the afterglow, ac-
cording to the CIB density. This is because the time delay
occurs mainly by the angular spreading at the location of
the absorption, ∆tA ∝ n−1
IR . The regenerated emission is
expected to only slightly change the afterglow light curve
even when we adopt a rather large value of nIR; for nIR = 1
cm−3, its contribution reaches ∼ 20% of the prompt emis-
sion around 10 s, but the total fluence around that time is
far below the detection threshold.
By fixing nIR to be 0.1 cm−3, we then investigated the
dependence of the GeV light-curve on ǫe and ǫB; the result
is shown in figure 2. Curves in figure 2(a) indicate the to-
tal fluence evaluated using various sets of (ǫe, ǫB), and the
ratio of regenerated and prompt emission is shown in figure
2(b). As we expect, rather large values of ǫe are favourable
for possible detection by the GLAST, because they make
the spectrum extend to high-energy region owing to the IC
scattering. In the case of the small ǫe, on the other hand,
the flux in the GeV region is not as strong as the case of
large ǫe, as already discussed in several past papers (e.g.,
Zhang & M´esz´aros 2001). Regardless of the detectability of
the total emission, it is easily found that the regenerated
emission is very weak compared with the prompt one, at
Figure 3. The same as figure 2, but for the wind-like profile of
surrounding medium with A∗ = 1.
most ∼ 30% if the value of ǫB is large, as shown in figure
2(b).
The result of the same calculation is shown for the case
of wind-like profile of surrounding matter, n(r) ∝ r−2, in fig-
ure 3 for various values of (ǫe, ǫB); other parameters are the
same as figure 2 except for A∗ = 1. We can confirm that the
GeV light-curves as well as their parameter dependence are
basically similar to the case of constant density profile. The
fraction of the regenerated flux to the prompt one shown in
figure 3, on the other hand, behaves somewhat differently;
it increases as the time pasts. However, it reaches only less
than 40% at 106 s after the onset of the afterglow, after
which the detection threshold for the fluence glows as t1/2
and the detection itself becomes more and more difficult.
Furthermore, favourable model with large ǫe gives smaller
contribution from the regenerated emission than the models
with small ǫe, which is not favoured from the viewpoint of
detectability. In consequence, the regenerated emission gives
only very slight correction to the prompt afterglow emission
in the GeV region, and further, it is found that this charac-
teristic is considerably independent of the relevant parame-
ters such as nIR, ǫe, and ǫB as well as the density profile of
the surrounding medium.
5 DISCUSSION AND CONCLUSIONS
In recent years, it is suggested that the delayed GeV emis-
sion, due to the absorption of the TeV photons by the
CIB radiation field and the following IC scattering on the
CMB photons, may be detected by the future high-energy
detectors such as the GLAST. As a source of the orig-
inal TeV emissions, the internal shocks (Dai & Lu 2002;
Razzaque et al. 2004) as well as the initial phase of the exter-
nal shocks (Wang et al. 2004) have been considered. These
authors claim that the delayed emission due to the regener-
ation process during propagation can be distinguished from
the direct GeV component due to the IC emission in the
afterglow phase.
The emission from the afterglows is also expected to
extend to TeV region with reasonable choices of relevant
parameters, and then the afterglow phase itself can also
be a source of the regenerated emission. Since an estima-
tion of this regenerated emission has not been performed
yet, and further, whether its intensity is above the detec-
tion threshold is nontrivial question, we investigated in this
paper the evolution of the regenerated light from the after-
glows, and discussed its detectability. As an original high-
energy emission model that extends to TeV region, we have
used the modeling of the synchrotron self-IC mechanism by
Zhang & M´esz´aros (2001) and Sari & Esin (2001), and also
used reasonable choices of relevant parameters.
As the result of calculation using formalism summarized
in § 3 and the constant density profile as well as ǫe = 0.5
and ǫB = 0.01, we found that the contribution of the re-
generation of GeV photons could give a correction at most
∼ 20% even if rather large value for the CIB density (nIR = 1
cm−3) is used as shown in figure 1. Although the CIB den-
sity around z = 1 is unknown, observations suggest that
the local CIB flux at 2.2 µm is of the order of 10 nW m−2
sr−1 (Wright & Johnson Wright & Johnson; Wright 2004),
which corresponds to nIR = 0.45 × 10−2 cm−3. Theoretical
model by Salamon & Stecker (1998) indicates that the co-
moving density of the CIB photons does not change largely
(i.e., less than factor of ∼ 3) from z = 1 to 0, and there-
fore the proper density of the CIB might be estimated to be
around 0.1 cm−3, although there remains a fair amount of
ambiguity. Our calculation suggests that even if we take a
fairly large value for the CIB density, the regenerated emis-
sion cannot change the shape of the intrinsic light curve, and
further, its intensity is far below the detection threshold of
the GLAST satellite.
Regenerated emission from GRB afterglows
5
In addition to the cosmic CIB density, extragalactic
magnetic field strength may affect the results of our calcula-
tion via the value of ∆tB appearing in equation (12) if it is
larger than 10−20 G that we used throughout the above dis-
cussions. The strength of extragalactic magnetic fields has
not been determined thus far. Faraday rotation measures im-
ply an upper limit of ∼ 10−9 G for a field with 1 Mpc correla-
tion length (see Kronberg 1994 for a review). Other methods
were proposed to probe fields in the range 10−10 to 10−21
G (Lee, Olinto & Sigl 1995; Plaga 1995; Guetta & Granot
2003). To interpret the observed microgauss magnetic fields
in galaxies and X-ray clusters, the seed fields required in
dynamo theories could be as low as 10−20 G (Kulsrud et al.
1997; Kulsrud 1999). Theoretical calculations of primordial
magnetic fields show that these fields could be of order 10−20
G or even as low as 10−29 G, generated during the cosmo-
logical QCD or electroweak phase transition, respectively
(Sigl, Olinto & Jedamzik 1997). Hence, although we used
the value of 10−20 G as the extragalactic magnetic field
strength, it is accompanied by a quite large amount of uncer-
tainty and it may affect the results give above. From figure
1, it is found that when BIG = 10−20 G, the time delay
of the regenerated emission is mainly dominated by angu-
lar spreading, i.e., ∆t = ∆tA, because the delayed compo-
nent changes according to the CIB density nIR. For further
smaller values of BIG than 10−20 G, therefore, our conclu-
sion does not change. On the other hand, if its value is suffi-
ciently large such that the condition ∆tB > ∆tA is satisfied
for the majority of possible γe, it further suppresses the re-
generated emission since its flux is inversely proportional to
∆tB as clearly shown in equation (12). In consequence, even
if the value of extragalactic magnetic fields differs from our
reference value, our central conclusion that the regenerated
emission is negligibly weak compared with the prompt one
does not change.
We also performed the same calculation but by fo-
cusing on dependence on the relevant parameters (ǫe, ǫB)
that strongly affect the high-energy emission mechanism.
We showed that although the total fluence in the GeV re-
gion considerably depends on the values of (ǫe, ǫB), but the
fraction of the regenerated emission to the prompt one is
always small for any choices of parameter sets (figure 2).
This characteristic holds in the case of the wind-like profile
of surrounding medium as shown in figure 3. All of these
facts given above enable us to probe a high-energy emission
mechanism in the GRB fireballs when data of the GeV pho-
tons are accumulated, because the expected signal would
be almost completely free of a large amount of uncertainty
concerning the CIB density as well as extragalactic mag-
netic field; they never affect the afterglow emission itself for
any choices of the relevant parameters (ǫe, ǫB), whichever
(constant or wind-like) profile of the surrounding medium is
truly realized.
ACKNOWLEDGMENTS
This work was supported by a Grant-in-Aid for JSPS Fel-
lows.
6
S. Ando
REFERENCES
Atkins R., et al., 2000, ApJ, 533, L119
Blumenthal G. R., Gould R. J., 1970, Rev. Mod. Phys., 42,
237
Bottcher M., Dermer C. D., 1998, ApJ, 499, L131
Bottcher M., Schlickeiser R., 1997, A&A, 325, 866
Cheng L. X., Cheng K. S., 1996, ApJ, 459, L79
Chevalier R. A., Li Z. Y., 2000, ApJ, 520, L29
Coppi P. S., Blandford R. D., 1990, MNRAS, 245, 453
Dai Z. G., Lu T., 2002, ApJ, 580, 1013
Dermer C. D., Bottcher M., Chiang J., 2000a, ApJ, 537,
255
Dermer C. D., Chiang J., Mitman K., 2000b, ApJ, 537, 785
Gehrels N., Michelson P., 1999, Astropart. Phys., 11, 277
Guetta D., Granot J., 2003, ApJ, 585, 885
Hurley K., et al., 1994, Nature, 371, 652
Kronberg P. P., 1994, Rep. Prog. Phys., 57, 325
Kulsrud R., 1999, Ann. Rev. A&A, 37, 37
Kulsrud R., Cowley S. C., Gruzinov A. V., Sudan R. N.,
1997, Phys. Rep., 283, 213
Lee S., Olinto A. V., Sigl G., 1995, ApJ, 455, L21
Lithwick Y., Sari R., 2001, ApJ, 555, 540
MacMinn D., Primack J. R., 1996, Space Sci. Rev., 75, 413
Madau P., Phinney E. S., 1996, ApJ, 456, 124
Malkan M. A., Stecker F. W., 1998, ApJ, 496, 13
M´esz´aros P., Rees M. J., Papathanassiou H., 1994, ApJ,
432, 181
Panaitescu A., Kumar P., 2000, ApJ, 543, 66
Panaitescu A., Kumar P., 2002, ApJ, 571, 779
Panaitescu A., M´esz´aros P., 1998, ApJ, 501, 772
Plaga R., 1995, Nature, 374, 430
Razzaque S., M´esz´aros P., Zhang B., , 2004, preprint (astro-
ph/0404076)
Salamon M. H., Stecker F. W., 1998, ApJ, 493, 547
Sari R., Esin A., 2001, ApJ, 548, 787
Sari R., Piran T., Narayan R., 1998, ApJ, 497, L17
Sigl G., Olinto A. V., Jedamzik K., 1997, Phys. Rev. D, 55,
4582
Stecker F. W., de Jager O. C., Salamon F. W., 1992, ApJ,
390, L49
Totani T., 1998, ApJ, 502, L13
Vietri M., 1997, Phys. Rev. Lett., 78, 4328
Wang X. Y., Cheng K. S., Dai Z. G., Lu T., 2004, ApJ,
604, 306
Waxman E., 1997, ApJ, 485, L5
Wei D. M., Lu T., 1998, ApJ, 505, 252
Wei D. M., Lu T., 2000, A&A, 360, L13
Wright E., 2004, New. Astron. Rev., 48, 465
Wright E., Johnson B., 2001, preprint (astro-ph/0107205)
Zhang B., M´esz´aros P., 2001, ApJ, 559, 110
|
astro-ph/0212185 | 1 | 0212 | 2002-12-07T17:42:28 | AGB Stars as Tracers of Star Formation Histories: Implications for GAIA Photometry and Spectroscopy | [
"astro-ph"
] | We argue that tracing star formation histories with ESA's space mission GAIA using main sequence turn-off (MSTO) point dating will mainly be effective in cases of mild interstellar extinction (E(B-V)<0.5). For higher reddenings the MSTO approach will be severely limited both in terms of traceable ages (t<0.5 Gyr at 8 kpc; E(B-V)=1.0) and/or distances (d<2 kpc for t=15 Gyr; E(B-V)=1.0), since the MSTO will be located at magnitudes too faint for GAIA. AGB stars may alternatively provide precise population ages with GAIA for a wide range of ages and metallicities, with traceable distances of up to 250 kpc at E(B-V)=0 (d=15 kpc if E(B-V)=2.0). It is essential however that effective temperatures precise to 0.01 dex, metallicities to 0.2 dex, and E(B-V) to 0.03 are derived for individual stars, in order to obtain their ages precise to 0.2 dex. This task is quite challenging for GAIA photometry and spectroscopy, though preliminary tests show that comparable precisions may be achieved with GAIA medium band photometry. | astro-ph | astro-ph |
GAIA Spectroscopy, Science and Technology
ASP Conference Series, Vol. XXX, 2002
U.Munari ed.
AGB Stars as Tracers of Star Formation Histories:
Implications for GAIA Photometry and Spectroscopy
A. Kucinskas
Lund Observatory, Box 43, SE-221 00 Lund, Sweden
Institute of Theoretical Physics & Astronomy, Gostauto 12, Vilnius
2600, Lithuania
L. Lindegren
Lund Observatory, Box 43, SE-221 00 Lund, Sweden
T. Tanab´e
Institute of Astronomy, The University of Tokyo, Tokyo, 181-015, Japan
V. Vansevicius
Institute of Physics, Gostauto 12, Vilnius 2600, Lithuania
Abstract. We argue that tracing star formation histories with GAIA
using main sequence turn-off (MSTO) point dating will mainly be effec-
tive in cases of very mild interstellar extinction (EB−V < 0.5). For higher
reddenings the MSTO approach will be severely limited both in terms of
traceable ages (t < 0.5 Gyr at 8 kpc; EB−V = 1.0) and/or distances
(d = 2 kpc if t ≤ 15 Gyr; EB−V = 1.0), since the MSTO will be located
at magnitudes too faint for GAIA. AGB stars may alternatively provide
precise population ages with GAIA for a wide range of ages and metal-
licities, with traceable distances of up to 250 kpc at EB−V = 0 (15 kpc if
EB−V = 2.0). It is essential however that effective temperatures, metal-
licities, and reddenings of individual stars are derived with the precision
of σ(log Teff ) ∼ 0.01, σ([M/H]) ∼ 0.2, and σ(EB−V ) ∼ 0.03, to obtain
σ(log t) ∼ 0.15. This task is quite challenging for GAIA photometry and
spectroscopy, though preliminary tests show that comparable precisions
may be achieved with GAIA medium band photometry.
1.
Introduction
Out of the numerous methods available for assessing ages of stellar populations
only a few will be practically applicable for use with GAIA, because of obvious
limitations related to the detection limits of GAIA etc. (e.g., Cacciari 2002).
Though the main sequence turn-off (MSTO) point approach is perhaps the most
reliable and accurate of all dating methods available today, it will be severely
distance-limited when used with GAIA.
1
2
Kucinskas, Lindegren, Tanab´e, Vansevicius
10
E(B-V)=0
r
y
G
,
X
A
M
t
1
E(B-V)=0.5
E(B-V)=1.0
10
E(B-V)=0
c
p
k
,
X
A
M
d
3
E(B-V)=0.5
E(B-V)=1.0
0.1
0.5
0.0
-0.5
-1.0
-1.5
-2.0
1
0.5
0.0
-0.5
-1.0
-1.5
-2.0
[Fe/H]
[Fe/H]
Figure 1.
Left: maximum age for which MSTO point can be dated
at 8 kpc with GAIA, for different [Fe/H] and EB−V . Right: maximum
distance at which stellar population up to 15 Gyr old can be dated
using MSTO with GAIA, for different [Fe/H] and EB−V .
The extent of these limitations is illustrated in Fig. 1. The left panel shows
the maximum age obtainable with GAIA employing MSTO fitting for stellar
population placed at a distance of 8 kpc (assuming that the MSTO point is
located at V ∼ 18.5, while the detection limit of GAIA is V = 20). The right
panel shows the maximum distance at which ages up to 15 Gyr may be quantified
(in both cases, the MSTO point on the isochrones of Girardi et al. (2000) was
dated). It is obvious that in cases of negligible interstellar extinction the MSTO
approach may work very efficiently with GAIA, providing reliable ages of stellar
populations up to ∼ 8 kpc. However, even mild interstellar extinction changes
the situation dramatically, shifting the MSTO point below the detection limits
of GAIA for many stellar populations within the Galaxy.
2. AGB stars with GAIA: star formation histories
A possible way to complement the MSTO approach would be with tracers that
are intrinsically brighter than MSTO stars, which could thus be used for probing
stellar populations on larger distance scales. We have shown, that stars on early-
AGB (non-thermally pulsing) may be well suited for this purpose, providing
reliable ages of stellar populations from the isochrone fitting to AGB sequences
on the observed HR diagrams (Kucinskas et al. 2002).
Though similar approaches have been used in the past (e.g., employing
AGB sequences in color-magnitude diagrams), the precision in derived ages was
low. We propose to work in the Mbol vs. Teff plane, using effective temperatures
of individual AGB stars derived from the fits of the observed spectral energy
distributions (SEDs) with synthetic SEDs. This yields a significant increase in
precision of derived population ages, which is a result of the precisely derived
Teff of individual AGB stars. Since distances will be known for the majority
of Galactic AGB stars with GAIA (which, combined with data from ground-
and/or space-based infrared surveys, will give precise Mbol), AGB stars will
allow stellar populations on large distance scales to be traced with GAIA.
-6
-5
l
o
b
M
-4
-3
-2
AGB stars as tracers of SFHs: implications for GAIA
3
0.2 Gyr
0.5 Gyr
1 Gyr
2 Gyr
5 Gyr
10 Gyr
0.2 Gyr
0.5 Gyr
1 Gyr
2 Gyr
5 Gyr
10 Gyr
0.2 Gyr
0.5 Gyr
1 Gyr
2 Gyr
5 Gyr
10 Gyr
NGC 1783
NGC 2121
3.60
3.55
3.50
3.60
3.55
log Teff
3.50
3.60
3.55
Kron 3
3.50
Figure 2. Observed HR diagrams with AGB sequences in star clus-
ters of the Magellanic Clouds: NGC 1783, 2121 (both LMC) and Kron
3 (SMC). Isochrones are from Bertelli et al. (1994; B94).
We illustrate the proposed approach employing the star clusters NGC 1783,
NGC 2121 and Kron 3 in the Magellanic Clouds. The observed SEDs of AGB
stars were constructed using BVRIJHK fluxes obtained from the literature. The
BaSel 2.2 library of stellar spectra (Lejeune et al. 1998) was used to produce a
template of synthetic photometric colors at different Teff and for the metallic-
ities of individual MC clusters. HR diagrams of individual clusters with AGB
sequences were constructed using Mbol derived from observed integrated fluxes
of individual stars and Teff obtained from the SED fitting procedure (Fig. 2).
AGB ages of individual clusters derived using different sets of isochrones, as well
as those obtained from MSTO in previous studies, are given in Table 1.
There is indeed a good agreement between AGB ages obtained using iso-
chrones of Bertelli et al. (1994; B94) or Lejeune & Schaerer (2001; LS01) on
one hand, and MSTO estimates on the other. AGB ages obtained with Girardi
et al. (2000; G00) isochrones are however considerably older than MSTO esti-
mates (Table 1). This effect tends to increase with decreasing metallicity.
It
should be stressed, thus, that these intrinsic discrepancies between different sets
AGB ages of star clusters in the Magellanic Clouds derived
Table 1.
in this work using different sets of isochrones (two metallicities are used
for Kron 3).
Cluster
[Fe/H]
AGB ages, this work (Gyr)
LS01
B94
G00
MSTO ages Ref
NGC 1783 −0.4
NGC 2121 −0.7
−1.3
−1.0
Kron 3
0.8 ± 0.1
3.8 ± 0.8
10.3 ± 2.8
4.9 ± 1.1
1.1 ± 0.2
6.0 ± 1.6
> 15
> 15
--
--
7.0 ± 1.6
--
0.9 ± 0.4 M89
3.2 ± 0.5
R01
8.0 ± 0.5
R00
6.0 ± 1.3 M98
4
Kucinskas, Lindegren, Tanab´e, Vansevicius
Table 2.
Predicted accuracies, versus V magnitude, for Teff , [M/H],
EB−V (individual AGB stars, 1X photometric system -- Vansevicius &
Bridzius 2002), and ages from AGB stars derived with GAIA. d is the
maximum distance at which these accuracies may be expected.
V
σ(log Teff ) σ([M/H]) σ(EB−V ) σ(log t)
dEB−V =0
dEB−V =2
18
20
< 0.01
0.03
< 0.2
0.3
< 0.03
0.06
0.15
0.45
kpc
100
250
kpc
6
15
of isochrones clearly indicate a need for better understanding of stellar evolu-
tion on the AGB, to fully exploit the potential of AGB stars for tracing stellar
populations in the Galaxy and beyond.
Possibilities for tracing star formation histories using AGB stars with GAIA,
employing photometric metallicities, gravities and reddenings (GAIA 1X medium
band photometric system, Vansevicius & Bridzius 2002) are summarized in Ta-
ble 2 (the error in age for a single star, σ(log t), is a lower limit since it reflects
only the errors in Teff and EB−V ). We conclude that, contrary to MSTO dat-
ing, AGB stars may provide precise estimates of ages (t > 0.5 Gyr) throughout
the Galaxy and beyond (d ∼< 250 kpc) for populations within a large range of
metallicities, even if interstellar extinction is non-negligible. To achieve this,
metallicities to ±0.2 dex, and effective temperatures to ±0.01 dex are highly
desirable. A precise knowledge of interstellar reddening (σ(EB−V ) ∼ 0.03) is
also essential.
Acknowledgments. This work was supported by the Wenner-Gren Foun-
dations.
References
Bertelli, G., Bressan, A., Chiosi, C., Fagotto, F., Nasi, E. 1994, A&AS, 106, 275
(B94)
Cacciari, C. 2002, EAS Publ. Ser., 2, 163
Girardi, L., Bressan, A., Bertelli, G., Chiosi, C. 2000, A&AS, 141, 371 (G00)
Kucinskas, A., Vansevicius, V., Tanab´e, T. 2002, Ap&SS, 280,151
Lejeune, T., Schaerer, D. 2001, A&A, 366, 538 (LS01)
Lejeune, T., Cuisinier, F., Buser, R. 1998, A&AS, 130, 65
Marigo, P., Girardi, L. 2001, A&A, 377, 132
Mighell, K.J., Sarajedini, A., French, R.S. 1998, AJ, 116, 2395 (M98)
Rich, R.M., Shara, M.M., Fall, S.M., Zurek, D. 2000, AJ, 119, 197 (R00)
Rich, R.M., Shara, M.M., Zurek, D. 2001, AJ, 122, 842 (R01)
Mould, J., Kristian, J., Nemec, J., Jensen, J., Aaronson, M. 1989 ApJ, 339, 84
(M89)
Vansevicius, V., Bridzius, A. 2002, this volume
|
astro-ph/0406044 | 1 | 0406 | 2004-06-01T21:45:13 | A 2 x 2 Degree I-band Survey around PKS 1343-601 | [
"astro-ph"
] | Motivated by the possibility that the highly obscured (A_B = 12 mag) radio galaxy PKS 1343-601 at (l,b,cz) = (309.7, +1.8, 3872km/s) might constitute the center of a heavily obscured cluster in the Great Attractor region, we have imaged about 2 x 2 degree of the core of this prospective cluster in the I-band using the WFI at the ESO 2.2m telescope at La Silla. We were able to identify 49 galaxies and 6 uncertain galaxy candidates. Although their distribution does not resemble a centrally condensed, massive cluster, its appearance -- severely influenced by the strong dust gradient across our surveyed region -- is entirely consistent with a cluster. | astro-ph | astro-ph |
Nearby Large-Scale Structures and the Zone of Avoidance
ASP Conference Series, Vol. NNN, 2004
A.P. Fairall, and P.A. Woudt, eds.
A 2 x 2 Degree I-band Survey around PKS 1343−601
Ren´ee C. Kraan-Korteweg
Depto. de Astronom´ıa, Universidad de Guanajuato, Apdo. Postal 144,
Guanajuato, GTO 36000, M´exico
Manuel Ochoa
Depto. de Astronom´ıa, Universidad de Guanajuato, Apdo. Postal 144,
Guanajuato, GTO 36000, M´exico
Instituto de Astronom´ıa, UNAM, Apdo. Postal 70-264, M´exico D.F.
04510, M´exico
Patrick. A. Woudt
Dept. of Astronomy, Univ. of Cape Town, Private Bag, Rondebosch
7700, South Africa
Heinz Andernach
Depto. de Astronom´ıa, Universidad de Guanajuato, Apdo. Postal 144,
Guanajuato, GTO 36000, M´exico
Abstract. Motivated by the possibility that the highly obscured (AB =
12.m) radio galaxy PKS 1343−601 at (ℓ, b, cz) = (309.◦7, +1.◦8, 3872 km s−1)
might constitute the center of a heavily obscured cluster in the Great At-
tractor region, we have imaged about 2◦
× 2◦ of the core of this prospec-
tive cluster in the I-band using the WFI at the ESO 2.2 m telescope at
La Silla. We were able to identify 49 galaxies and 6 uncertain galaxy
candidates. Although their distribution does not resemble a centrally
condensed, massive cluster, its appearance -- severely influenced by the
strong dust gradient across our surveyed region -- is entirely consistent
with a cluster.
1.
Introduction
The Great Attractor (GA), a large extended mass overdensity in the nearby
Universe, was discovered from the infall pattern of elliptical galaxies (Dressler
et al. 1987). Kolatt et al. (1995) determined its center and found it to ly
exactly behind the southern Milky Way at (ℓ, b, cz) = (320◦, 0◦, 4000 km s−1).
Because of the increasing dust absorption at lower Galactic latitudes it has
remained difficult to assess whether this mass density is being traced by the
galaxy distribution.
The deep optical galaxy search of partially obscured galaxies in the GA re-
gion (Woudt & Kraan-Korteweg 2001) revealed the Norma cluster (ACO 3627)
1
2
Kraan-Korteweg et al.
at (ℓ, b) = (325.3◦, −7.2◦) to be a region of very high galaxy density in the GA
region. Follow-up redshift observations found the Norma cluster to be compa-
rable in size, richness and mass to the Coma cluster, albeit nearer by a factor of
1.4 (Kraan-Korteweg et al. 1996; Woudt et al. these proceedings). It is therefore
the most likely candidate to constitute the previously unrecognized bottom of
the potential well of the Great Attractor (GA) overdensity.
Is Norma the only massive cluster behind the Milky Way in the GA region
or might other clusters form part of a much broader and hence more massive
core of the GA? The identification of possible further dynamically important
mass contributors to the GA at lower latitudes is problematic, however, because
of the high extinction in the optical, star-crowding in the near-infrared surveys
DENIS and 2MASS, and even in X-rays -- despite the fact that dust extinction
and stellar confusion are unimportant -- because of the photoelectric aborption
at high Galactic HI column densities.
1.1. The PKS 1343−601 Galaxy
PKS 1343−601 is the 2nd strongest radio continuum source in the southern sky
(f20cm = 79 Jy; Mc Adam 1991 and references therein) and lies at very low
latitudes (ℓ, b) = (309.◦7, +1.◦8). Woudt (1998) and Kraan-Korteweg & Woudt
(1999) did suspect that this radio galaxy might constitute the central galaxy of
a cluster: this galaxy lies behind an obscuration layer of about 12m extinction
in the B-band, as estimated from the DIRBE/IRAS extinction maps (Schlegel,
Finkbeiner, & Davis 1998) and is not visible in the optical. With a diameter of
28 arcsec in the Gunn-z filter and, based on the Hα emission line, a recession
velocity of cz = 3872 km s−1 (West & Tarenghi 1989), it must be a giant elliptical
galaxy (a diameter of about 4 arcmin if corrected for extinction effects using
Cameron's (1990) extinction laws). Such galaxies are not generally isolated but
found at the cores of clusters.
If PKS 1343−601 marks the dynamical center of a cluster, then the Abell
radius, defined as 1.′7/z where z is the redshift, corresponds to RA = 2.◦2 =
3 h−1
50 Mpc) on the sky at the redshift-distance of PKS 1343−601. In the optical,
only a handful of highly obscured galaxies (at the highest latitudes and lowest
extinction levels) could be identified within this radius and even 2MASS reveals
only 9 galaxy candidates within this radius.
As rich clusters generally are strong X-ray emitters, we searched for evidence
of such emission. PKS 1343−601 has not been detected with ROSAT but the
soft X-ray emission would be strongly absorbed by the Galactic HI at that po-
sition. It has, however, been detected with ASCA (Tashiro et al. 1998) showing
slightly extended diffuse hard X-ray emission at the position of PKS 1343−601.
The flux of kT = 3.9 keV is quite high for a single galaxy, hence could be in-
dicative of emission from a cluster. However, recent higher resolution X-ray
observations with XMM do not support this supposition (see Schroder et al.,
these proceedings). The extended X-ray emission is thus probably due to the
Inverse Compton process in the radio lobes, as already suggested by Ebeling,
Mullis, & Tully in 2002.
Interestingly enough, the ZOA Parkes Multi Beam H I survey does uncover a
significant excess of galaxies at this position in velocity space (Kraan-Korteweg
et al. 2004; Henning, Kraan-Korteweg, & Staveley-Smith, these proceedings).
A 2 x 2 Degree I-band Survey around PKS 1343−601
3
8000
6000
4000
2000
0
2
4
6
Figure 1.
The velocity distribution as a function of distance from
PKS 1343−601. The crossed circles are galaxies detected in the Parkes
ZOA Multibeam Survey (Henning et al., in prep.), the filled circles
from optical velocity data as given in LEDA.
However, no "finger of God" is obvious, the characteristic signature of a cluster
in redshift space. Then again, HI is not a good tracer of high-density regions
such as cluster cores, since spiral galaxies generally avoid the cores of clusters
or are HI depleted (Bravo-Alfaro et al. 2000, Vollmer et al. 2001).
The HI-velocities plus the few known optical velocities within the Abell
radius and its immediate surroundings provide some dynamical support for the
existence of this cluster. Not only do we find a significant peak in the velocity
histogram at the velocity of PKS 1343−601, but the velocities as a function of
distance from the cluster center lie within a narrow range of the central radio
source and are well separated in velocity space from field galaxies (see Fig. 1).
To verify whether PKS 1343−601 indeed marks the center of an obscured
cluster, we have imaged the core of this prospective cluster in the I-band in
which extinction effects are less severe (AI = 0.45AB) using the Wide Field
Imager WFI of the ESO/MPG 2.2-m telescope at la Silla.
2. The 2 x 2 Degree I-band Survey
A total of 16 WFI fields, each 34′
× 2◦
were observed in May 1999. The surveyed area is outlined in Figs. 3 and 4
together with the Abell radius of the prospective cluster. The field centered on
PKS 1343−601 has a slightly higher exposure time (1500s compared to 600s)
and consists of 5 dithered exposures to improve the spatial resolution.
× 33′, covering a total area of about 2◦
After standard reduction with IRAF , all the images were inspected visually.
The small images of the strongly obscured galaxies and the heavy star-crowding
(partly covering the galaxies) make an automatic detection algorithm impossible.
A constant changing of the intensity and contrast levels while viewing a field
brings out the more extended low surface brightness borders of galaxies and also
emphasizes the difference in the light distribution between stars and galaxies.
This procedure makes the identification of such obscured objects more feasible.
In this way, 49 galaxies were identified, 25 of them probably of elliptical
morphology, in addition to 6 uncertain galaxy candidates. A sample of galaxy
images, from the brightest to the faintest galaxies, is displayed in Fig. 2.
4
Kraan-Korteweg et al.
Figure 2.
An image gallery of 8 of the 49 certain galaxy candidates
identified around PKS 1343−601 on the I-band images. Each image is
about 36′′
× 36′′.
Galaxy 33 is the giant radio galaxy PKS 1343−601 (AI = 5.m5). The spiral
galaxy (# 28) is by far the largest galaxy identified in this survey but it is
located in an area of relatively low extinction (AI = 2.m2). Galaxies 29, 30,
32, 33 and 34 are from the central field and although some of them look like
mere smudges (e.g. # 32) all of them are independently confirmed by a NIR
survey (J, H, K ′) of the central 36′
× 36′ region of this possible cluster using
the Japanese/South African Infrared Survey Facility (Nagayama et al. 2004;
Nagayama at al., these proceedings). In fact these recent observations reveal
the very low surface brightness galaxy # 34 at AI = 5.m5 to actually be a very
extended edge-on spiral galaxy.
3. The Detected Galaxy Distribution
The distribution in Galactic coordinates of the unveiled galaxies is shown in
Fig. 3. The outlined square region indicates the area imaged with the WFI in
the I-band (16 fields of 34′
× 33′ each) around PKS 1343−601 (large dot), and
the circle the Abell radius of RA = 2.◦2 = 3 h−1
50 Mpc. The filled circles mark the
25 elliptical galaxies and the crossed circles the 24 spiral galaxies. Note that the
morphology is very uncertain due to the heavy obscuration. The open circles
show another 6 uncertain candidates.
The distribution shows that is was possible to identify galaxies to extinction
levels of AI ∼< 5.m0, i.e., to similar extinction levels within this photometric
passband compared to our deep optical galaxy searches in the B-band where we
identified galaxies down to AB ∼< 5.m0 (Kraan-Korteweg 2000; Woudt & Kraan-
Korteweg 2001). Ellipticals are found mainly at higher extinction levels while
at lower extinction levels spirals predominate. As the cores of ellipticals and
A 2 x 2 Degree I-band Survey around PKS 1343−601
5
Figure 3.
The distribution in Galactic coordinates of the identified
galaxies. The surveyed area is marked as the tilted square. The circle
corresponds to the Abell radius. Filled, crossed and open circles mark
elliptical, spiral and uncertain galaxies. The contours indicate extinc-
tion levels in the I-band: AI = 2.m0, 3.m0, 4.m0, 5.m0 (solid), 7.m5 (long
dash) and 10.m0 (short dash) following Schlegel et al. 1998.
bulges of spirals contain mainly the red old star population this trend was to
be expected. It is still interesting to note that the density of ellipticals seems
to show a concentration around PKS 1343−601, even though the distribution
does not match our expectations of a normal, centrally condensed, rich cluster.
Its appearance is, however, strongly modulated by the extinction gradient. To
assess whether the unveiled distribution really is consistent with a rich cluster
hidden behind an increasingly thickening exinction layer, we simulated how a
rich cluster would appear at the position of PKS 1343−601.
4. Does the Galaxy Distribution Indicate a Galaxy Cluster?
For the simulation, we have used the deep Coma cluster catalog of 6724 galaxies
by Godwin, Metcalfe, & Peach (1983). First we move the rich Coma cluster at
the radial velocity of PKS 1343−601. This results in an extension of the cluster
size and galaxy diameters by a factor of f = 1.77 and an increase in brightness
of ∆m = 1.m24. We then transform the B-band magnitudes to the I-band
with a mean color term of (B − I) = 2.m0 which is representative of early-type
galaxies as well as bulges of spiral galaxies. Although it is difficult to assess the
magnitude limit obtained with the here used non-standard narrow I-band filter
(∆λ = 275 A centered on λ = 9148 A), we estimate to be able to find galaxies to
a standard I-band magnitude of Ilim ≈ 17.m5. Applying this cut-off would leave
us with 1578 galaxies in our surveyed cluster area. The left panel in Fig. 4 shows
the resulting distribution, where the symbols are proportional to brightness.
However, we still have to take the effect of dust absorption into account.
Even in the I-band, the minimum absorption is AI ≈ 2m increasing to a maxi-
6
Kraan-Korteweg et al.
Left panel:
Figure 4.
Simulation of the distribution of Coma
cluster galaxies if it would be positioned at the redshift space of
PKS 1343−601, and would have been observed under the same con-
ditions with the with the Wide Field I-band imager at La Silla. Right
panel: Remaining galaxies if the cluster galaxies are subjected to the
same foreground extinction (extinction contours as in Fig. 3).
mum of AI ≈ 10m. Absorbing each galaxy individually according to the DIRBE
extinction value at their respective positions (see contour levels in Fig. 3) us-
ing the inverse Cameron laws (1990), leaves only 51 identifiable galaxies. Their
distribution is shown in the right panel of Fig. 4.
This number is entirely consistent with the number of galaxies identified in
our real search (49 certain and 6 uncertain galaxy candidates). The resulting
distribution furthermore is a near replica of the actual detections (compare right
panel of Fig. 4 with Fig. 3): galaxies are found down to the same extinction level
of about AI ∼< 5m. Only Coma's two cD galaxies peak through slightly thicker
extinction layers -- similar to PKS 1343−601 -- and the galaxy numbers within
the various extinction intervals are quite similar.
5. Discussion
The simulation thus seems to suggest that the uncovered galaxy distribution
in our I-band survey around PKS 1343−601 is consistent with being due to
a galaxy cluster at the position and distance of PKS 1343−601. Care should,
however, be taken when interpreting the galaxy simulations. Small changes in,
e.g. the estimated I-band magnitude limit, the actual distance, the Cameron
(1990) extinction corrections (which are type-dependent and determined for the
B-band), assumption of a fixed B − I color term, etc., might alter the simulated
galaxy distribution significantly. Furthermore, the fact that no X-ray emission
typical of a massive cluster (Ebeling et al. 2002, Ebeling et al., these pro-
ceedings, Schroder et al., these proceedings) argues against a very rich cluster.
The most likely interpretation therefore is, that there exists a cluster around
A 2 x 2 Degree I-band Survey around PKS 1343−601
7
PKS 1343−601, but that it rather is an intermediate-mass cluster like Hydra or
Centauraus, which is in agreement with the independent analysis by Nagayama
et al. 2004, Nagayama et al., these proceedings.
Acknowledgments. This research used the Lyon-Meudon Extragalactic
Database (LEDA), supplied by the LEDA team at the Centre de Recherche
Astronomique de Lyon, Obs. de Lyon. RCKK, MO and HA thank CONACyT
for their support (research grants 27602E and 40094F). PAW acknowledges the
National Research Foundation for financial support.
References
Bravo-Alfaro, H., Cayatte, V., Van Gorkom, J.H., & Balkowski, C. 2000, AJ,
119, 580
Cameron, L.M. 1990, A&A, 233, 16
Dressler, A., Faber, S.M., Burstein, D., et al. 1987, ApJ, 313, 37
Ebeling, H., Mullis, C.R., & Tully, R.B. 2002, ApJ, 580, 774
Godwin, J.G., Metcalfe, N., & Peach J.V. 1983, MNRAS, 202, 113
Jones, P.A., & McAdam, W.B. 1992, ApJS, 80, 137
Kolatt, T., Dekel, A., & Lahav, O. 1995, MNRAS, 275, 797
Kraan-Korteweg, R.C. 2000, A&AS, 141, 123
Kraan-Korteweg, R.C., & Lahav, O. 2000, A&ARv, 10, 211
Kraan-Korteweg, R.C., & Woudt, P.A. 1999, PASA, 16, 53
Kraan-Korteweg, R.C., Fairall, A.P., Balkowski, C. 1995, A&A, 297, 617
Kraan-Korteweg, R.C., Staveley-Smith, L., Donley, J. et al. 2004, in ASP Conf.
Ser., Maps of the Cosmos, ed. M. Colless & L. Staveley-Smith, (San
Francisco: ASP), in press
Kraan-Korteweg, R.C., Woudt, P.A., Cayatte, V., et al. 1996, Nature, 379, 519
McAdam, W.B. 1991, PASA, 9, 255
Nagayama, T., Woudt, P.A., Nagashima, C., et al. 2004, MNRAS, in press
Schlegel, D.J., Finkbeiner, D.P., & Davis M. 1998, ApJ, 500, 525
Tashiro, M., Kaneda, H., Makishima, K., et al. 1998, ApJ, 499, 713
Vollmer, B., Cayatte, V., van Driel, W., et al. 2001, A&A, 369, 432
West, R.M., & Tarenghi, M. 1989, A&A, 223, 61
Woudt, P.A. 1998, Ph.D. thesis, Univ. of Cape Town
Woudt, P.A., & Kraan-Korteweg, R.C. 2001, A&A, 380, 441
|
0801.2698 | 1 | 0801 | 2008-01-17T15:10:53 | Microlensing probes the AGN structure of the lensed quasar J1131-1231 | [
"astro-ph"
] | We present the analysis of single epoch long slit spectra of the three brightest images of the gravitationally lensed system J1131-1231. These spectra provide one of the clearest observational evidence for differential micro-lensing of broad emission lines (BELs) in a gravitationally lensed quasar. The micro-lensing effect enables us: (1) to confirm that the width of the emission lines is anti-correlated to the size of the emitting region; (2) to show that the bulk of FeII is emitted in the outer parts of the Broad Line Region (BLR) while another fraction of FeII is produced in a compact region; (3) to derive interesting informations on the origin of the narrow intrinsic MgII absorption doublet observed in that system. | astro-ph | astro-ph |
To appear in "Proceedings of the Huatulco meeting on AGN (2007)"
RevMexAA(SC)
MICROLENSING PROBES THE AGN STRUCTURE OF THE LENSED
QUASAR J1131-1231
D. Sluse1, J.-F. Claeskens2, D. Hutsem´ekers2,3 and J.Surdej2,4
RESUMEN
ABSTRACT
We present the analysis of single epoch long slit spectra of the three brightest images of the gravitationally
lensed system J1131-1231. These spectra provide one of the clearest observational evidence for differential
micro-lensing of broad emission lines (BELs) in a gravitationally lensed quasar. The micro-lensing effect
enables us: (1) to confirm that the width of the emission lines is anti-correlated to the size of the emitting
region; (2) to show that the bulk of Fe II is emitted in the outer parts of the Broad Line Region (BLR) while
another fraction of Fe II is produced in a compact region; (3) to derive interesting informations on the origin of
the narrow intrinsic Mg II absorption doublet observed in that system.
Key Words: GRAVITATIONAL LENSING - QUASARS:INDIVIDUAL: RXS J113155.4-123155
1. INTRODUCTION
J1131-1231 is one of the nearest gravitationally
lensed AGN. The lensing galaxy (zl = 0.295) splits
the light rays from the source (zs = 0.66) into four
macro-images: three bright images (A-B-C) sepa-
rated on the sky by typically 1′′ and a fainter compo-
nent (D) located at 3.6′′ from A (Sluse et al. 2003).
The present study focuses on the lensed images A-
B-C. The lensing effects in a system like J1131-1231
occur for two different regimes. First, the lensing
galaxy produces four resolved macro-images of the
background source. Second, the compact masses in
that galaxy (typically 10−6 < M < 106 M⊙) split
each macro-lensed image into multiple of unresolved
micro-images separated by a few micro-arcseconds.
Because gravitational lensing magnifies the source,
each (unresolved) macro-image i is amplified by a
factor Mi associated with macro-lensing and by a
factor µi due to micro-lensing. The typical length
scale for micro-lensing (in the quasar plane) is the
Einstein radius RE of the micro-lens:
RE =r 4GM
c2
DlsDos
Dol
= 14.3sM h−1
M⊙
light − days,
(1)
where Dos, Dls, Dol are the angular-size distances
between observer and source (os), lens and source
1Laboratoire
Ecole Polytechnique
F´ed´erale de Lausanne (EPFL) Observatoire, 1290 Sauverny,
Switzerland ([email protected]).
d'Astrophysique,
2Institut d'Astrophysique et de G´eophysique, Universit´e de
Li`ege, All´ee du 6 Aout 17, B5C, B-4000 Sart Tilman, Belgium
3Maıtre de recherches du F.N.R.S (Belgium)
4Directeur de recherches honoraire du F.N.R.S.(Belgium)
(ls) and observer and lens (ol) (see e.g. Wambs-
ganss 2006 for a review on micro-lensing). Micro-
lensing acts as a magnifying glass which amplifies
regions of the source on scales smaller or equal to
the micro-lens Einstein radius. For J1131-1231, a
solar mass Einstein radius has a size similar to the
usually assumed size of the BLR, meaning that re-
gions as large as the BLR can be amplified by micro-
lensing. Because micro-lensing occurs independently
in each macro-lensed images, we can track for spec-
tral differences between multiple images of a lensed
quasar. Such differences reveal the selective micro-
amplification of small regions of the source. Unfor-
tunately, spectral differences between macro-lensed
images are not only due to micro-lensing. The time
delay between the macro-lensed images can also in-
duce spectral differences between images (because
each lensed image is a snapshot of the source at a
different epoch). Hopefully, the time delay between
images A-B-C of J1131-1231 is less than a few days
(e.g. Saha et al. 2006) and thus can be neglected.
Differential reddening between lensed images due to
the lensing galaxy might also exist. This effect seems
however to be negligible in the case of J1131-1231
(Sluse et al. 2006, 2007). After a brief description
of the data (Sect. 2), we interpret the main spectral
differences observed between images A-B-C of J1131-
1231 in the micro-lensing framework (Sect. 3). More
results can be found in Sluse et al. (2007; SLU07).
2. DATA
We present long slit spectra of J1131-1231 ob-
tained with the FORS2 instrument (ESO Very Large
1
2
SLUSE ET AL.
Telescope) on April 26th 2003. These data consist
of spectra obtained with the 1′′ slit oriented along
the lensed images B-A-C. We used 2 grisms which
cover the wavelength ranges 3890 < λ < 6280 A and
6760 < λ < 8810 A with a resolving power around
800 and 1500 at the respective central wavelengths.
A total exposure time of 960s has been devoted to
each grism. This observational set up enables us
to cover the (AGN) rest-frame range 2500-5600 A.
Standard reductions steps have been followed. Due
to blending between the spectra of images A-B-C, we
extracted the spectra by fitting three Moffat profiles
along the spatial direction for each wavelength bin
independently. Comparison of the fitted 2D spec-
tra with the observations confirmed that this proce-
dure provides an optimal extraction of the individual
spectra (except in the [O III] region where residuals
up to 0.5% of the [O III] flux are observed).
600
500
400
300
200
100
0
A
B(shift+100)
C
4000
5000
6000
7000
8000
9000
Observed wavelength (Angstroms)
Fig. 1. Spectra of images A-B-C in J1131-1231. Arbi-
trary flux units.
3. RESULTS
The spectra of A-B-C are shown on Fig. 1. Be-
cause spectral differences between A-B-C are only
due to micro-lensing, one can assume that the ob-
served spectra Fi are simply made of a superposition
of a spectrum FM which is only macro-lensed and
of a spectrum FMµ both macro and micro-lensed.
Using pairs of observed spectra, it is then easy to
extract both components FM and FMµ. Defining
M = M1/M2 (> 0) as the macro-amplification ratio
between image 1 and image 2 and µ as the micro-
lensing factor affecting image 1 (image 2 assumed
not to be micro-lensed), then we have:
F1 = M FM + M µFMµ
F2 = FM + FMµ .
(2)
The latter equations can be rewritten:
FM = F1/M −µF2
FMµ = F2−F1/M
1−µ
1−µ
,
(3)
where µ must be chosen to satisfy the positivity con-
straint FM > 0 and FMµ > 0. Assuming that the
narrow line region is too large to be micro-lensed,
one easily retrieves M and µ for any pair of images.
One can show that image B is not affected by micro-
lensing (SLU07) so that this decomposition applied
to image pairs A-B and C-B reveals which regions
of the quasar are micro-lensed in images A and C
(Fig. 2). We see that the broad emission lines (BELs)
are nearly completely micro-lensed in image C and
partially micro-lensed in image A. This difference of
behaviour is likely due to the larger RE of the micro-
lens affecting C. Indeed, the micro-lensed fraction of
a given emitting region increases with the RE of the
micro-lens (Eq. 1). In this context, an emission re-
gion that is micro-lensed in image A should also be
micro-lensed in image C and should be very compact,
while a larger emitting region might be micro-lensed
only in image C. Finally, even larger regions should
not be micro-lensed neither in A nor in C. Conse-
quently, this analysis provides information about the
AGN structure based on the identification and char-
acterization of the emitting regions micro-lensed in
A & C (a complementary analysis technique is pre-
sented in SLU07). Based on Fig. 2, we conclude that
the smallest regions (micro-lensed in both A and C)
are the regions emitting the very broad component of
Mg II, the Fe II emission observed in the range 3080-
3540 A and a fraction of the Fe IIopt in the range 4630-
4800 A. The remaining of the Fe II emission as well
as the bulk of the BELs are emitted in regions with
a larger size (because they are micro-lensed only in
C). Finally, the largest regions are those emitting the
narrow core of the BELs, which is not micro-lensed.
We notice that microlensing in both A and C of the
wings of the BELS is easily explained with an out-
flowing BLR. Although less likely, micro-lensing of
a BLR with a rotating accretion disk can produce a
similar observational signature (Abajas et al. 2002).
Another remarkable result is the presence in
our spectra of an intrinsic Mg II absorption doublet
blueshifted at z =0.654 (∆v ∼ -660 km/s). These
lines disappear from the spectrum ratios A/B and
C/B. This indicates that the absorbed flux is propor-
tional to the flux coming from the continuum+BLR.
This implies that the region at the origin of the ab-
sorption must cover both the continuum and the
BLR and that, within the uncertainties, their depths
are identical in the spectra of images A, B, C. One
MICRO-LENSING PROBES THE AGN STRUCTURE OF J1131-1231
3
3
2.5
2
1.5
1
0.5
0
FM
FMµ
MgII
FeII (3080-3540 A)
FeIIopt
<---------------->
<---------------------------------->
4
3.5
3
2.5
MgII
FeII UV
-
-
-
FeII UV
-
-
-
FeII UV
FeII UV
-
-
-
-
-
-
A band
FeII
-
-
-
-
-
-
-
-
Hγ
-
-
---
Hβ
2
FeII (3080-3540 A)
<---------------->
FeIIopt
<---------------------------------->
Hβ
Hγ
A band
---
-
-
-
-
-
-
-
-
1.5
1
0.5
0
FM
FMµ
2000
2500
3000
3500
4000
4500
5000
5500
6000
2000
2500
3000
3500
4000
4500
5000
5500
6000
Wavelength (Angstroms)
Wavelength (Angstroms)
Fig. 2. Fraction of the spectrum affected (FM µ, dotted) and unaffected (FM , solid) by micro-lensing (arbitrary units)
for image A (left) and C (right); rest frame wavelengths. For image A, the broadest components of Mg II, H γ and H β as
well as some Fe II multiplets (in the range 3080-3540 A and around 4600 A) are micro-lensed. For image C, nearly the
whole BELs and Fe II emissions are micro-lensed.
can see the differential micro-lensing at work in
J1131-1231 as a probe of the inhomogeneities in the
absorbing medium (Fig. 3). Indeed, by lensing more
(less) strongly some regions of the source, differential
micro-lensing increases (decreases) the contribution
of a fraction of the intervening absorber to the total
absorption. The nearly identical absorption depths
seen in the three images indicate that both the spa-
tial distribution and the optical depth of the absorb-
ing clouds must be quite homogeneous over the con-
tinuum and BLR. This is compatible with an absorp-
tion region constituted of a large number of small
absorbing clouds, their projected sizes being signifi-
cantly smaller than the local continuum region (i.e.
continuum+BLR).
We have shown (SLU07) that micro-lensing is a
very useful tool to study the AGN structure. A step
forward is to perform a spectroscopic monitoring of
micro-lensing events (the latter occur at any time
in quadruply imaged quasars). Such a program will
allow to derive absolute sizes (or tight upper limits)
of the BLR and maybe to probe their geometry.
This work is supported by the Swiss National Sci-
ence Foundation, by ESA PRODEX under contract
PEA C90194HST and by the Belgian Federal Science
Policy Office.
REFERENCES
[]Abajas, C. et al. 2002, ApJ, 576, 640
[]Saha, P. et al. 2006, A&A, 450, 461
[]Sluse, D. et al. 2003, A&A, 406, L43
[]Sluse, D. et al. 2006, A&A, 449, 539
[]Sluse, D. et al. 2007, A&A, 468, 885 (SLU07)
[]Wambsganss, J. in Gravitational lensing: strong, weak
and micro. Saas-Fee Advanced Course 33. Editors G.
Meylan, P. Jetzer and P. North.
Fig. 3. Cartoon picture of the effect of micro-lensing on
intrinsic absorbers. Left: absorbing clouds (black tiny
circles) in front of the pseudo-continuum region (contin-
uum+BLR, assumed to emit uniformly for clarity; large
circle; solid contours) for two kinds of cloud distribution:
homogeneous distribution of clouds (up) and inhomoge-
neous distribution (down). The small circle (dashed con-
tours) shows the fraction of the emission region amplified
by micro-lensing. Right: corresponding absorption dou-
blet (normalized to the local pseudo continuum) without
micro-lensing (solid) and with micro-lensing (dashed).
|
0704.1300 | 1 | 0704 | 2007-04-10T20:02:48 | The obscured X-ray source population in the HELLAS2XMM survey: the Spitzer view | [
"astro-ph"
] | Recent X-ray surveys have provided a large number of high-luminosity, obscured Active Galactic Nuclei (AGN), the so-called Type 2 quasars. Despite the large amount of multi-wavelength supporting data, the main parameters related to the black holes harbored in such AGN are still poorly known. Here we present the results obtained for a sample of eight Type 2 quasars in the redshift range 0.9-2.1 selected from the HELLAS2XMM survey, for which we used Ks-band, Spitzer IRAC and MIPS data at 24 micron to estimate bolometric corrections, black hole masses, and Eddington ratios. | astro-ph | astro-ph |
The obscured X-ray source population in the
HELLAS2XMM survey: the Spitzer view
Cristian Vignali∗,†, Francesca Pozzi∗, Andrea Comastri†, Lucia Pozzetti†,
Marco Mignoli†, Carlotta Gruppioni†, Giovanni Zamorani†, Carlo Lari∗∗,
Francesca Civano∗, Marcella Brusa‡, Fabrizio Fiore§ and Roberto
Maiolino§
∗Dipartimento di Astronomia, Università di Bologna, Via Ranzani 1, I -- 40127 Bologna, Italy
†INAF -- Osservatorio Astronomico di Bologna, Via Ranzani 1, I -- 40127 Bologna, Italy
∗∗INAF -- Istituto di Radioastronomia (IRA), Via Gobetti 101, I -- 40129 Bologna, Italy
‡Max Planck Institut für Extraterrestrische Physik (MPE), Giessenbachstrasse 1, D -- 85748
§INAF -- Osservatorio Astronomico di Roma, Via Frascati 33, I -- 00040 Monteporzio-Catone (RM),
Garching bei München, Germany
Italy
Abstract. Recent X-ray surveys have provided a large number of high-luminosity, obscured Active
Galactic Nuclei (AGN), the so-called Type 2 quasars. Despite the large amount of multi-wavelength
supporting data, the main parameters related to the black holes harbored in such AGN are still poorly
known. Here we present the preliminary results obtained for a sample of eight Type 2 quasars in
the redshift range ≈ 0.9 -- 2.1 selected from the HELLAS2XMM survey, for which we used Ks-band,
Spitzer IRAC and MIPS data at 24 m m to estimate bolometric corrections, black hole masses, and
Eddington ratios.
Keywords: Galaxies: active -- Galaxies: nuclei -- X-rays: galaxies
PACS: 98.54.-h, 98.58.Jg
INTRODUCTION
Over the last six years, the X-ray surveys carried out by Chandra and XMM-Newton
(e.g., [1, 2, 3]; see [4] for a review) have provided remarkable results in resolving a
significant fraction of the cosmic X-ray background (XRB; [5, 6]), up to ≈ 80% in the
2 -- 8 keV band (e.g., [7, 8]). Despite the idea that a large fraction of the accretion-driven
energy density in the Universe resides in obscured X-ray sources has been widely sup-
ported and accepted (e.g., [9, 10]), until recently only limited information was available
to properly characterize the broad-band emission of the counterparts of the X-ray ob-
scured sources and provide a reliable estimate of their bolometric output.
In this context, Spitzer data have provided a major step forward the understanding
of the broad-band properties of the X-ray source populations. If, on one hand, Spitzer
data have allowed to pursue the "pioneering" studies of [11] on the spectral energy
distributions (SEDs) of broad-line (Type 1), unobscured quasars at higher redshifts (e.g.,
[12]), on the other hand they have produced significant results in the definition of the
SEDs of narrow-line (Type 2), obscured AGN (e.g., [13]).
In this work we aim at providing a robust determination of the bolometric luminosity
for hard X-ray selected obscured AGN. This result can be achieved by effectively
disentangling the nuclear emission related to the active nucleus from the host galaxy
starlight, which represents the dominant component (at least for our obscured sources)
at optical and near-infrared (near-IR) wavelengths.
SAMPLE SELECTION AND Ks-BAND PROPERTIES
The sources presented in this work were selected from the HELLAS2XMM survey ([3])
which, at the 2 -- 10 keV flux limit of ≈ 10−14 erg cm−2 s−1, covers ≈ 1.4 square degrees
of the sky using XMM-Newton archival pointings ([14]). Approximately 80% of the
HELLAS2XMM sources have a spectroscopic optical classification in the final source
catalog (see [15] for details). In particular, we selected eight sources from the original
sample of [16] which are characterized by faint (23.7 -- 25.1) R-band magnitudes and
bright Ks-band counterparts (≈ 17.6 -- 19.1); all of our sources are therefore classified
as extremely red objects (EROs, R − Ks > 5 in Vega magnitudes). From the good-
quality Ks-band images, [16] were able to study the surface brightness profiles of these
sources, obtaining a morphological classification. While two sources are associated with
point-like objects, the remaining six sources are extended, showing a profile typical of
elliptical galaxies. In this latter class of sources, the active nucleus, although evident in
the X-ray band, appears hidden or suppressed at optical and near-IR wavelengths, where
the observed emission is clearly dominated by the host galaxy starlight. The relatively
good constraints on the nuclear emission in the near-IR represent a starting point for the
analysis of the Spitzer IRAC and MIPS data.
Due to the faint R-band magnitudes of our sources, optical spectroscopy was not
feasible even with the 8-m telescope facilities; however, the bright near-IR counterparts
of our sources allowed us to obtain spectroscopic redshifts in the Ks band with ISAAC at
VLT for two sources: one point-like AGN is classified as a Type 1.9 quasar at a redshift
of 2.09, while one extended source has line ratios typical of a LINER at z=1.35 (see [17]
for further details on these classifications). For the remaining sources, the redshift has
been estimated using the optical and near-IR magnitudes, along with the morphological
information, as extensively described in §5.1 of [16]; all of the redshifts are in the range
≈ 0.9 -- 2.1. The large column densities [≈ 1022 -- a few×1023 cm−2] and the 2 -- 10 keV
luminosities [≈ (1 − 8) × 1044 erg s−1, once corrected for the absorption] place our
sources in the class of the high-luminosity, obscured AGN, the so-called Type 2 quasars
(see, e.g., [18] and references therein).
SPITZER DATA
For our sample of eight sources, we obtained IRAC observations of 480 s integration
time and MIPS observations at 24 m m for a total integration time per position of
≈ 1400 s. All of the sources are detected in the four IRAC bands and in MIPS; the
faintest source in MIPS has a 24 m m flux density of ≈ 150 m Jy (≈ 5s detection; see
[19] for further details on data reduction and cleaning procedures).
ANALYSIS OF THE TYPE 2 QUASAR SPECTRAL ENERGY
DISTRIBUTIONS
A reliable determination of the bolometric output of our AGN sample requires that the
nuclear component, directly related to the accretion processes, is disentangled from the
emission of the host galaxy, which provides a dominant contribution in the optical and
Ks bands (in the case of extended sources, see [16]). To achieve this goal, we constructed
SEDs for all our sources over the optical, near- and mid-IR range. At the same time, we
used Spitzer data to improve our previous estimates on the source redshift when possible.
In the following, we consider the sample of six extended sources and two point-like
objects separately, since a different approach has been adopted for the two sub-samples.
Extended sources
As already pointed out, from the Ks-band morphological analysis carried out by [16],
we know that at least up to 2.2 m m (observed frame) the stellar contribution is mostly
responsible for the emission of these sources. At longer wavelengths, the emission of
the active nucleus is expected to arise as reprocessed radiation of the primary emission,
while the emission from the galaxy should drop significantly, assuming reasonable
elliptical templates. Although many models have been developed in the past to deal
with circum-nuclear dust emission (including the effects of the torus geometry and
opening angle, grain size distribution and density), in our study we adopted a more
phenomenological approach. To reproduce the observed data, we used a combination of
two components, one for the host galaxy and another related to the reprocessing of the
nuclear emission.
For the galaxy component, we adopted a set of early-type galaxy templates obtained
from the synthetic spectra of [20], assuming a simple stellar population spanning a large
range of ages (see [19] for details). For the nuclear component, we adopted the templates
of [21], which are based on the interpolation of the observed nuclear IR data (at least, up
to ≈ 20 m m) of a sample of local AGN through the radiative transfer models of [22]. The
strength of such an approach is that the nuclear templates depend upon two quantities,
the intrinsic 2 -- 10 keV luminosity (which provides the normalization of the SED) and
the column density (responsible for the shape of the SED), and these are known directly
from the X-ray spectra ([23]), once the redshift is known.
We also used all the available information, extended over the Spitzer wavelength
range, to place better constraints on the source redshift than those reported in [16].
Overall, we find a good agreement with the redshifts presented in [16], although Spitzer
allows us to provide estimates with lower uncertainties; only for one source the redshift
is significantly lower (z ≈ 1 instead of ≈ 2) and likely more reliable.
The data are well reproduced by the sum of the two components; the emission from
the galaxy progressively becomes less important at wavelengths above ≈ 4 m m (in the
source rest frame), where the nuclear reprocessed emission starts emerging significantly
(see Fig. 1, left panel), being dominant in MIPS at 24 m m. Furthermore, the latter is
fully consistent with the upper limits provided in the Ks band by [16].
FIGURE 1. Rest-frame SEDs for two representative Type 2 quasars of the current sample: an obscured
AGN hosted by an elliptical galaxy (on the left) and a point-like AGN (on the right). (Left) The observed
data (filled circles) are reproduced by summing up (solid line) the contribution of an early-type galaxy
template (dot-dashed line) to the reprocessed nuclear component (dashed line). The dotted line shows
the nuclear component obtained from the templates of [21], normalized using the X-ray luminosity and
column density (i.e., without fitting the data; see text and [19] for details). The downward-pointing arrow
indicates the constraint on the nuclear emission derived from the Ks-band data ([16]). The combination of
the two templates is also consistent with the R-Ks color. (Right) The observed data (filled circles) are well
reproduced by the red quasar template from [13] (solid line).
Point-like sources
For the two point-like sources, we adopted a different strategy, since their emission
in the near-IR is dominated by the unresolved AGN. To reproduce their observed SEDs,
we extincted a Type 1 quasar template from [11] with several extinction laws, but we
were not able to find a satisfactory solution. Then we used the recently published red
quasar template from [13] and found good agreement with the data (Fig. 1, right panel),
consistently with the results obtained for some obscured AGN in the ELAIS-S1 field
([24]). As in the AGN sub-sample described above, most of the uncertainty lies in the
far-IR, where a proper study of the SEDs would require MIPS data at 70 and 160 m m.
BOLOMETRIC CORRECTIONS
The determination of the SEDs is meant to be the first step toward the estimate of
the bolometric luminosities (Lbol) of obscured AGN. The bolometric luminosities can
be estimated from the luminosity in a given band by applying a suitable bolometric
correction kbol; typically, to convert the 2 -- 10 keV luminosity into Lbol, kbol≈ 30 is
assumed, although this value was derived from the average of few dozens of bright,
mostly low-redshift Type 1 quasars ([11]). For obscured sources, only few estimates are
present in literature (e.g., [13]). We derived kbol by integrating the quasar SEDs over the
X-ray (0.5 -- 500 keV) and IR (1 -- 1000 m m) intervals; in the X-ray band, we converted
the 2 -- 10 keV luminosity assuming a power law with photon index G = 1.9 (typical for
AGN emission) and the observed column density ([23]).
To derive the bolometric corrections, we accounted for both the covering factor of
the absorbing material (i.e., the opening angle of the torus) and the anisotropy of the IR
emission. According to unification models of AGN, the former effect should be directly
related to the observed fraction of Type 2/Type 1 AGN which, in the latest models of
[6], is ≈ 1.5 in the luminosity range of our sample. Furthermore, the torus is likely to
re-emit a fraction of the intercepted radiation in a direction which does not lie along
our line-of-sight; the correction for this anisotropy, according to the templates of [21], is
≈ 10 -- 20% (given the column densities of our sources). Once these corrections are taken
into account, we obtain hkboli ≈ 35 (median kbol≈ 26), similar to the average value of
[11]; the Type 1.9 quasar at z=2.09 has the highest kbol (≈ 97); see [19] for a discussion
on the uncertainties in these estimates.
BLACK HOLE MASSES AND EDDINGTON RATIOS
For the six AGN hosted by elliptical galaxies, we can derive both the galaxy and
black hole masses. Since the near-IR emission is dominated by the galaxy starlight,
we computed the rest-frame LK assuming the appropriate SED templates and then the
galaxy masses using M⋆/LK ≈ 0.5 − 0.9 ([20]); all of our AGN are hosted by massive
galaxies (≈ 1 − 6 × 1011 M⊙).
To estimate the black hole masses, we used the local MBH -- LK ([25]) which, along
with the M⋆/LK values, provides a MBH -- M⋆ relation. Despite several attempts in the
recent literature to investigate whether and how the black hole mass vs. stellar mass
relation evolves with cosmic time, there is no consensus yet. In this work, we assume
the findings of [26], who found that in the redshift range covered by our sources, the
MBH-M⋆ relation evolves by a factor of ≈ 2 with respect to the local value; see [19]
for an extensive discussion. Under this hypothesis, we obtain black hole masses for the
six obscured quasars hosted by elliptical galaxies of ≈ 2.0 × 108 − 2.5 × 109 M⊙; these
values are broadly consistent with the average black hole masses obtained by [27] for
the Sloan Digital Sky Survey (SDSS) Type 1 quasars (using optical and ultra-violet mass
scaling relationships) in our redshift range (≈ 3.5 × 108 − 8.6 × 108 M⊙).
As a final step, we derived the Eddington ratios, defined as Lbol/LEdd, where LEdd is
the Eddington luminosity. We note that the uncertainties related to these estimates are
clearly large, due to the uncertainties of the approach adopted to derive the bolometric
luminosities (through the templates of [21]) and the black hole masses (see above).
The average Eddington ratio is ≈ 0.05, suggesting that our obscured quasars may have
already passed their rapidly accreting phase and are reaching their final masses at low
Eddington rates. The Eddington ratios of our sources are significantly lower than those
derived for the SDSS Type 1 quasars in the same redshift range (≈ 0.3 -- 0.4, see [27]).
SUMMARY
We used optical, near-IR, and Spitzer IRAC and MIPS (at 24 m m) data to unveil the re-
processed nuclear emission of eight hard X-ray selected Type 2 quasars at z ≈ 0.9 − 2.1.
From proper modelling of the nuclear SEDs, we derived a median (average) bolometric
correction of ≈ 26 (≈ 35). For the six obscured sources dominated by the host galaxy
starlight up to near-IR wavelengths, we also derived black hole masses of the order of
2.0 × 108 − 2.5 × 109 M⊙ and relatively low Eddington ratios (≈ 0.05), suggestive of a
low-activity accretion phase.
ACKNOWLEDGMENTS
The authors acknowledge partial financial support by the Italian Space Agency under
the contract ASI -- INAF I/023/05/0.
REFERENCES
1. R. Giacconi et al., ApJS 139, 369 -- 410 (2002).
2. D. M. Alexander et al., AJ 126, 539 -- 574 (2003).
3. F. Fiore et al., A&A 409, 79 -- 90 (2003).
4. W. N. Brandt and G. Hasinger, ARA&A 43, 827 -- 859 (2005).
5. A. Comastri, G. Setti, G. Zamorani and G. Hasinger, A&A 296, 1 -- 12 (1995).
6. R. Gilli, A. Comastri and G. Hasinger, A&A 463, 79 -- 96 (2007).
7. F. E. Bauer, D. M. Alexander, W. N. Brandt, D. P. Schneider, E. Treister, A. E. Hornschemeier and
G. P. Garmire, AJ 128, 2048 -- 2065 (2004).
11. M. Elvis et al., ApJS 95, 1 -- 68 (1994).
12. G. T. Richards et al., ApJS 166, 470 -- 497 (2006).
13. M. Polletta et al., ApJ 642, 673 -- 693 (2006).
14. A. Baldi, S. Molendi, A. Comastri, F. Fiore, G. Matt and C. Vignali, ApJ 564, 190 -- 195 (2002).
15. F. Cocchia et al., A&A, in press, astro-ph/0612023 (2007).
16. M. Mignoli et al., A&A 418, 827 -- 840 (2004).
17. R. Maiolino et al., A&A 445, 457 -- 463 (2006).
18. C. Vignali, D. M. Alexander and A. Comastri, MNRAS 373, 321 -- 329 (2006).
19. F. Pozzi et al., A&A, in press, arXiv:0704.0735 (2007).
20. G. Bruzual and S. Charlot, MNRAS 344, 1000 -- 1028 (2003).
21. L. Silva, R. Maiolino and G. L. Granato, MNRAS 355, 973 -- 985 (2004).
22. G. L. Granato and L. Danese, MNRAS 268, 235 -- 252 (1994).
23. G. C. Perola et al., A&A 421, 491 -- 501 (2004).
24. C. Gruppioni et al., A&A, in preparation.
25. A. Marconi and L. K. Hunt, ApJ 589, L75 -- L77 (2003).
26. C. Y. Peng, C. D. Impey, L.C. Ho, E. J. Barton and H.-W. Rix, ApJ 640, 114 -- 125 (2006).
27. R. J. McLure and J. S. Dunlop, MNRAS 352, 1390 -- 1404 (2004).
8. R. C. Hickox and M. Markevitch, ApJ 645, 95 -- 114 (2006).
9. A. J. Barger, L. L. Cowie, R. F. Mushotzky, Y. Yang, W.-H. Wang, A. T. Steffen and P. Capak, AJ 129,
10. P. F. Hopkins, L. Hernquist, T. J. Cox, T. Di Matteo, B. Robertson and V. Springel, ApJS 163, 1 -- 49
578 -- 609 (2005).
(2006).
|
astro-ph/0104239 | 2 | 0104 | 2002-04-27T13:38:26 | Density Contrast-Peculiar Velocity Relation in the Newtonian Gauge | [
"astro-ph"
] | In general relativistic framework of the large scale structure formation theory in the universe, we investigate the relation between density contrast and peculiar velocity in the Newtonian gauge. According to the gauge-invariant property of the energy-momentum tensor in the Newtonian gauge, we consider the perturbation of velocity in the energy-momentum tensor behaves as the Newtonian peculiar velocity. It is shown that in the relativistic framework, the relation between peculiar velocity and density contrast has an extra correction term with respect to the Newtonian Peebles formula which in small scales, can be ignorable . The relativistic correction of peculiar velocity for the structures with the extension of few hundred mega parsec is about few percent which is smaller than the accuracy of the recent observations for measuring peculiar velocity. The peculiar velocity in the general relativistic framework also changes the contribution of Doppler effect on the anisotropy of CMB. | astro-ph | astro-ph |
Density Contrast -- Peculiar Velocity Relation in the
Newtonian Gauge
Sohrab Rahvar∗
Department of Physics, Sharif University of Technology,
P.O.Box 11365 -- 9161, Tehran, Iran
and
Institute for Studies in Theoretical Physics and Mathematics,
P.O.Box 19395 -- 5531, Tehran, Iran
October 26, 2018
Abstract
In general relativistic framework of large scale structure formation theory in
the universe, we investigate relation between density contrast and peculiar veloc-
ity in the Newtonian gauge. According to the gauge -- invariant property of the
energy -- momentum tensor in the this gauge, the velocity perturbation behaves as
the Newtonian peculiar velocity. In this framework, relation between peculiar veloc-
ity and density contrast with respect to the Newtonian Peebles formula has an extra
correction term which is ignorable for the small scales structures. The relativistic
correction of peculiar velocity for the structures with the extension of hundred mega
parsec is about few percent which is smaller than the accuracy of the recent peculiar
velocity measurements. We also study CMB anisotropies due to the Doppler effect
in the Newtonian gauge comparing with using the Newtonian gravity.
1
Introduction
In the large structure formation theory, structures originate from small density fluctua-
tions that are amplified by gravitational instability. These initial fluctuations are assumed
to be generated during inflationary epoch and are inflated to the beyond of the horizon
while the size of horizon remains constant during the inflation. These quantum fluctua-
tions beyond the horizon freeze and evolve like classical perturbations. They reenter the
horizon at a later time when the horizon grows to the size of perturbation. The observa-
tional evidence for these small density fluctuations can be seen at the decoupling epoch
as the anisotropy of CMB and also existence of the large scale structures in the universe.
One of the consequences of density fluctuation is metric perturbations in the Friedman --
Robertson -- Walker (FRW) universe and deviation of velocity field from the Hubble law
which is called peculiar velocity. Hence studying peculiar velocity could be a useful indirect
∗E-mail:[email protected]
1
method to find out the structures in the FRW universe. In the Newtonian linear struc-
ture formation theory, relation between density contrast and peculiar velocity has been
obtained by Peebles [1]. From the experimental point of view, Bertschinger and Dekel
introduced POTENT method to reconstruct velocity potential from the radial component
of velocity field [2]. Recently Branchini et al combined measurement of radial velocity of
galaxies and independent measurement of density contrast to estimate Ω [3].
On the scales well below than the Hubble radius, the Newtonian theory of gravitation is
a good approximation and Peebles' approach is widely used. However for observations
and simulations are probing scales which are a significant fraction of the Hubble radius,
the light-cone effect should taken into account [4]. Our aim in this work is to generalize
relation between density contrast and peculiar velocity to general relativistic framework.
The relativistic linear perturbation of FRW universe can be studied by two type of formu-
lations exist in the literature. The 'gauge fixing' formulation, originated by Lifshitz [5],
considers perturbed components of the metric which are related to the energy momentum
tensor. According to the gauge freedom problem in the theory, it is necessary to put
a constraint on perturbed part of metric for gauge fixing (e.g. synchronous gauge). In
the Second approach, initiated by Bardeen in 1980 [6] gauge invariant variables can be
made by the combining perturbed metric elements. These gauge-invariant variables are
well known from other physical theories. For example, in classical electrodynamics it is
usually more physical to work in terms of gauge-invariant electromagnetic fields rather
than in terms of gauge-dependent scalar and vector potentials. The subsequent works in
FRW perturbation theory can be found in some textbooks [7]-[10].
General relativistic analysis of peculiar velocity has been studied in the 'covariant' fluid
flow [11], 'Harmonic gauge' [4] and ' quasi-Newtonian gauge' fixing methods. In this work
we are going to use conformal Newtonian gauge also known as the (Longitudinal gauge) to
obtain relation between density contrast and peculiar velocity. The conformal Newtonian
gauge was introduced by Mukhanov et al [12] as a simple gauge used for scalar mode of
metric perturbation. In this gauge perturbation of metric elements are gauge-invariant
variables. For the case of absence of non-diagonal space -- space component in energy --
momentum tensor, perturbation of metric can be interpreted as relativistic generalization
of Newtonian gravitational potential. We generalize this idea for peculiar velocity field
according to its gauge -- invariant property in the Energy -- Momentum tensor. Here, we
restrict ourselves to the flat universe, in the matter and radiation dominant epochs, and
obtain relativistic relation between density contrast and peculiar velocity field. It is shown
that in small scales compare to the size of horizon, where general relativistic effect due
to the light cone effect is negligible, our formulation reduces to Peebles equation. For the
structures with 300Mpc extension, the relativistic correction is about one percent. We
also calculate the relativistic Doppler effect contribution on the CMB anisotropies and
compare it with the Newtonian one.
The organization of the paper is as follows. In Sec. II we write down perturbation of met-
ric in Newtonian gauge. In Sec. III we introduce energy -- momentum tensor and Einstein
equations in this gauge and In Sec. IV, we obtain relativistic relation between density
contrast and peculiar velocity for flat universe in radiation and matter dominant epochs.
In Sec. V , we obtain signature of Doppler effect on the anisotropies of CMB, using
relativistic peculiar velocity. We conclude in Sec. VI with a brief summary and some
discussions.
2
2 Perturbation of FRW in the Newtonian Gauge
Consider a small perturbation of metric hµν with respect to FRW background gµν
ds2 = (gµν + hµν)dxµdxν
(1)
where gµν = a(τ )2(dτ 2 − γijdxidxj) and τ is conformal time. Greek and Roman letters go
from 0 to 3 and 1 to 3, respectively. The metric perturbation hµν could be categorized into
three distinct types like Scalar, vector and tensor perturbation. This classification refers
to the way in which hµν transforms under three -- space coordinate transformation on the
constant time hyper-surface. A covariant description of tensor decomposition has been
shown in [13]. Vector and tensor modes exhibit no instability. Vector perturbation decays
kinematically in an expanding universe and tensor perturbation leads to gravitational
waves which do not couple to energy density and pressure inhomogeneities. However,
scalar perturbation may lead to growing inhomogeneities of matter.
The most general form of the scalar metric perturbation is constructed by using four
scalar quantities φ, ψ, B and E:
h(s)
µν = 2φ
B;i 2(ψγij − E;ij) ! .
−B;i
(2)
One of the main difficulties of the relativistic structure formation theory is that there is
gauge freedom in the theory and one can make an artificial perturbation by coordinate
transformation in such a way that infinitesimal space -- time distance between the two event
remains constant. One way to overcome this problem is using gauge fixing and the other
way is gauge invariant method [6]. In the latter method, some quantities are found to
be gauge invariant under coordinate transformation. It seems that the gauge -- invariant
quantities are similar to the electric and magnetic fields in the theory of electrodynamics
where they are measurable quantities in contrast to the vector and scalar potentials that
are gauge dependent parameters. The simplest gauge -- invariant quantities from the linear
combination of φ, ψ, B and E which span the two -- dimension space of gauge -- invariant
variables are:
Φ(gi) = φ +
Ψ(gi) = ψ −
1
a
a′
a
[(B − E
′
′
)a]
(B − E
′
)
(3)
(4)
where prime represents derivation with respect to conformal time and index (gi) stands for
gauge invariance. The above variables were first introduced by Bardeen [6]. Newtonian
gauge is defined by choosing B = E = 0 [12]. It is seen that in this gauge φ and ψ become
gauge -- invariant variables and the metric can be written as follows:
ds2 = a(τ )2h(1 + 2φ)dτ 2 − (1 − 2ψ)γijdxidxji .
(5)
3 Einstein Equation in The Newtonian Gauge
In the perturbed background of the FRW metric, the linear perturbation of Einstein
equation can be written as follows:
δGµ
ν = 8πGδT µ
ν.
(6)
3
The left hand side of eq.(6) can be obtained from perturbation of metric in the Newtonian
gauge and the right hand side of it is obtained by perturbed energy -- momentum tensor.
Restricting our attention to scalar perturbation of energy -- momentum tensor, we can
express the most general first order energy -- momentum tensor in terms of four scalars as
follows [6]:
δT µ
ν =
δρ
−(ρ + p)a−1ui
(ρ + p)aui
−δp + σ;ij
! ,
(7)
where δρ and δp are the perturbations of energy density and pressure, W is potential
for the three -- velocity field vi = a(t)ui(x, τ ) = ∇iW and σ is anisotropic stress. Bardeen
has shown that in the Newtonian gauge: δT µ(gi)
ν [6]. According to this result,
vi = a(t)ui may be regarded as Newtonian peculiar velocity. For a perfect fluid, the
anisotropic stress which leads to non-diagonal space -- space component of the energy --
momentum tensor vanishes. In this case it can be shown that φ = ψ. Using the metric of
eq.(5) in the left hand side and eq.(7) in the right hand side of eq.(6), Einstein equation
can be rewritten:
ν = δT µ
∇2φ − 3Hφ′ − 3(H2 − κ)φ = 4φGδρ,
(aφ)′,i = −4φG(ρ + p)ui,
φ′′ + 3Hφ′ + (2H
+ H2 − κ)φ = 4φGδp,
′
(8)
(9)
(10)
where H is the Hubble constant in conformal time. For simplicity in calculation we
consider the following change of variable.
φ = 4πG(ρ + p)1/2ω = (4πG)1/2[(H2 − H
′
+ κ)a2]1/2ω.
(11)
By substituting eq.(11) into eq.(10), this equation is rewritten in terms of ω as follows:
where
′′
ω
− cs
2∇2ω −
θ′′
θ
ω = 0,
θ =
1
a
(
ρ
ρ + p
)1/2(1 −
3κ
8πGρa2 )1/2.
(12)
(13)
and for the adiabatic perturbations c2
can obtain exact solution for differential equation (12), considering (cs = 0), yields:
δρ . For the case of matter dominant epoch one
s = δp
ω(x, τ ) = E1(x)θ(τ ) + E2(x)θ(τ )Z dτ
θ2 .
(14)
Our aim is to obtain density contrast and peculiar velocity in terms of ω. Density contrast
δ = δρ
0 component of the FRW equation
(H9 + κ = 8πGρa2
δ can be obtained by dividing eq.(8) to G0
):
3
δ =
2
3(H2 + κ)
(∇2φ − 3Hφ
′
− 3(H2 − κ)φ).
(15)
In what follows, Eqs.(9) and (15) will be our main equations in the following sections. It
is seen that these two equations are as function of φ which, can be obtained by solving
eq.(12).
In the next section we find an explicit relation between density contrast and
peculiar velocity in the framework of general relativistic perturbation theory in the New-
tonian gauge. we restrict ourselves to the case of radiation and matter dominant epochs
in the flat universe.
4
4 Density Contrast -- Peculiar Velocity Relation
From the recent CMB and Supernova type I experiments, it seems that the curvature of
the universe is flat [14] (κ = 0). In this section according to sequence of the radiation
and matter dominant epochs from early universe, we apply Eqs.(9) and (15) for these two
regimes.
4.1 Radiation dominant epoch
In the radiation dominant epoch, for the flat universe, according to FRW equations, scale
factor evolves like a = a0τ . In this case eq.(13) reduce to:
θ = s 3
4
1
a
.
(16)
For simplicity in the calculation we consider structures larger than Hubble radius, so the
second term in the right hand side of eq.(12) can be ignored with respect to other terms
and ω(x, τ ) can be obtained form eq.(14). Substituting eq.(16) into eq.(14), the dynamics
of ω is obtained:
1
D(x, τ )
ω =
(H2 − H′)4/2 (
a
′
)
,
(17)
where D(x, τ ) = [E1 sin(ντ ) + E2 cos(ντ )]e(ik·x) and ν = k
eqs.(11), (15) and (9), we obtain following equations for φ(x, τ ), δ(x, τ ) and ui(x, τ ):
√3. Substituting eq.(17) into
φ(x, τ ) =
β(x, τ ) =
1
τ 3 [(ντ cos(ντ ) − sin(ντ ))C1 − (ντ sin(ντ ) + cos(ντ ))C6] e(ik·x),
4
τ 3 [([(ντ )2 − 1] sin(ντ ) + ντ [1 −
1
2
(ντ )2] sin(ντ ))C2)]e(ik·x),
(ντ )2] cos(ντ )C1
+ ([1 − (ντ )2] cos(ντ ) + ντ [1 −
1
2
∇ · v =
([(ντ cos(ντ ) − cos(ντ ))C1 − (ντ sin(ντ ) + cos(ντ ))C2] e(ik·x)
a1k2
a2τ
k2a0
2a2 [(ν cos(ντ ) − ν2τ sin(ντ ) − ν cos(ντ ))C1
− (ν sin(ντ ) + ν2τ cos(ντ ) − ν sin(ντ ))C2])e(ik·x).
−
(18)
(19)
(20)
According to our consideration for the structures larger than Hubble radius ντ << 1,
eqs.(19) and (20) are reduced as follows:
∇ · v = −
e(ik·x)C2,
k2a0
a5τ
δ =
4
τ 3 e(ik·x)C2.
(21)
(22)
Now one can divide eq.(21) by eq.(22) to obtain the explicit relation between density
contrast and the peculiar velocity:
∇ · v = −
3
4
Hδ(ντ )2.
(23)
It is seen from eq.(23) that in the radiation dominant epoch, the relativistic peculiar
velocity for a given density contrast is much smaller than what one can expect from the
Newtonian formula.
5
4.2 Matter dominant epoch
In the matter dominant epoch, for κ = 0, eq.(13) is simplified into this form:
θ =
1
a
.
(24)
For the flat universe, the scale factor grows as a = a0τ 2/2. By substituting eq.(24) in
eq.(14), the dynamics of ω as a function of conformal time is obtained:
ω = E1(x)τ 3 + E2(x)τ −2,
(25)
where E1(x) and E2(x) are arbitrary functions of spatial coordinates. Using ω in the
eq.(11), the value of φ obtain as follows:
φ(x, τ ) = C1(x) + τ −7C2(x).
(26)
Substituting eq.(26) into eqs.(9) and (15), δ and u are obtained as functions of conformal
time, C1(x) and C2(x) as:
δ =
ui(x, τ ) =
1
1
a 5
(
C2(x),i
C1(x),i
−
2
3
τ 6
).
τ
6 h(∇2C6(x) − 81C2(x)) − τ −5(τ 2∇2C2(x) + 18C2(x))i ,
Neglecting the decaying modes, we rewrite eqs.(27) and (28) in the Fourier space:
δk = −
ik · vi(k)
a
=
1
6
2
3a1
(k2τ 2 + 12)C1(k),
k2
τ
C1(k).
(27)
(28)
(29)
(30)
Dividing eq.(30) by eq.(29), one can obtain an explicit relation between density contrast
and peculiar velocity in the following form:
ik.vk
a
= −
Hδk
1 + 3a2H 2
k2
(31)
where, H is the Hubble constant in physical time and vk is the Fourier transformation of
the peculiar velocity. By using the definition of scalar potential for the peculiar velocity,
vk = ik
a
Wk, eq.(31) can be written:
Hδk =
k2
a2
Wk + 3H 2Wk.
In the real space, eq.(32) changes to:
∇ · v = −H
δ
b
+ 3H 2W
(32)
(33)
where b is biasing factor and W = R v · dl is the scalar potential of velocity field and can
be obtained by radial integrating of velocity field. The second term in the right hand side
6
of eq.(33) is a result of general relativistic corrections. This correction can be estimated
by dividing the second term by the first term in the right hand side of eq.(32).
3H 2Wk
k2
a2 Wk
≃ (
λ
dH
)2
(34)
where dH is the size of horizon and λ = a
k is the size of structure. Is is seen that for the
small scale structures ( compare to the size of horizon), the general relativistic correction is
negligible. This correction for the structures with 300Mpc extension is about one percent.
5 The Effect of Relativistic Peculiar Velocity on CMB
The inhomogeneities in the universe induce anisotropies in the distribution of relic back-
ground of the photons on the CMB. The anisotropy of CMB can be caused by several
reasons. One of the main reasons is that the matter which scattered the radiation in
our direction had a peculiar velocity with respect to comoving frame when the scattering
occurred. Since the peculiar velocity of matter is different at different directions in the
sky, this leads to an anisotropy on the CMB. The universe at the last scattering epoch has
a redshift of z ≈ 1000. Since the equality of matter and radiation occurred at z ≈ 5 × 104,
we can consider that at last scattering epoch universe resides in the matter dominant.
Using eq.(31) for relativistic peculiar velocity in matter dominant epoch, the relativistic
anisotropy due to Doppler effect on CMB is obtained as follows:
δT
T
≃ v =
λ
dH
1 + 3( λ
dH
)2
δ
(35)
where λ is the size of structure and dH is the Hubble radius at the decoupling epoch. It is
seen from eq.(35) that the temperature perturbation of the CMB due to peculiar velocity
in the relativistic framework can be expressed in the term of Newtonian one:
(
δT
T
)Rel =
( δT
T )N ew
1 + 3( θ
)2
θH
(36)
dH
3M pc)
T )N ew = λ
δ is the contribution of Newtonian Doppler anisotropy, θ = 34′′(Ωh)( λ
where ( δT
is the angular size of structure [8] and θH ≃ 1◦ is the angular size of Hubble radius. It is
seen from eq.(36) that for the large angles, the relativistic contribution of Doppler effect
due to the peculiar velocity on the fluctuations of CMB is smaller than the contribution
of Newtonian anisotropy.
We compare the relativistic Doppler effect also with Sachs -- Wolf effect (which is one of
the reasons for the the anisotropy of CMB arising from the variation in the gravitational
potential at the last scattering [15]). It is shown that the contribution of this effect on
anisotropy of CMB can be given by:
δT
T
=
1
3
(
θ
θH
)2δ.
(37)
Eqs.(35) and (37) shows that Sachs -- wolf effect at large angles (θ > θH ) is dominant term
on the anisotropies of CMB while for the small angles Doppler term is dominant. Fig.
7
1 shows the fluctuation of temperature as a function of angular size of the anisotropy
for the case of Newtonian and relativistic peculiar velocity and Sachs -- wolf effect which
normalized to the density contrast. It can be shown that the contribution of relativistic
peculiar velocity affect the power spectrum of CMB.
6 Conclusion
In this letter, we have tried to identify the peculiar velocity in general relativistic theory of
structure formation. We obtain a relation between the density contrast and the peculiar
velocity for the matter and radiation dominant epochs.
It has been shown that the
relativistic correction is about few percent for structures with extension of hundred mega
parsec. The effect of the relativistic peculiar velocity as Doppler effect on the CMB also
has been calculated. It was shown that the contribution of relativistic peculiar velocity
on the anisotropy of CMB is less than what one expected from the Newtonian theory.
References
[1] Peebles, P. J. E. The large -- Scale Structure of the Universe (Princeton University
Press, Princeton, NJ, 1980).
[2] Bertschinger E., Dekel, A, APJ, 336, L5 (1989).
[3] Branchini, E., Zehavi, I., Plionis, M and Dekel, A, MNRAS, 313 419 (2000).
[4] Mansouri, R., Rahvar, S, Int. J. Mod. Phys. D, 11, 321 (2002); astro-ph/9907041.
[5] Lifshitz, E. M, Phys. USSR., 10, 116 (1946).
[6] Bardeen, J, Phys. Rev. D, 22, 1882 (1980).
[7] Weinberg, S, Graviation and Cosmology (Wiley, New York, 1972).
[8] Padmanabhan, T, Structure Formation in Universe (Cambridge University Press,
Cambridge, England, 1993).
[9] Bertschinger, E ,in Porceeding of the Cosmology and Large Scale Structure, edited
by R. Schaefer, J. Silk, M. Spiro and J. Zinn-Justin ( Elsevier Science, Amesterdam,
1996), p. 273.
[10] Bertschinger, E (astro-ph/0101009).
[11] Bruni, M., Lyth, D. H, Phys. Lett. B, 323, 118 (1994).
[12] Mukhanov, V. F., Feldman, H. A. and Brandenberger, R. H, Phys. Rep., 215, 1
(1992).
[13] Stewart, J, Class. Quantum. Grav., 7, 1169 (1990).
[14] Lange, A. E et. al, Phys. Rev. D, 63 (2001); astro-ph/0004389.
[15] Sachs, R. K., Wolfe, A. M, APJ, 147, 73 (1967).
8
7 Figure Caption
This figure shows the normalized perturbation of temperature to density contrast ( δT
T )/δ
as a function of normalized angular size of structures to the size of horizon θ
. Solid line,
θH
dashed line and dashed -- dot line represent the contribution of Newtonian, Relativistic
Doppler effect and Sachs -- Wolf effect on the anisotropy of CMB, respectively.
9
Newtonian Doppler
Relativistic Doppler
Sachs-Wolf
2
1.5
1
0.5
δ
/
)
T
/
T
δ
(
0
0
0.5
1
θ/ θ
H
1.5
2
|
astro-ph/0303167 | 1 | 0303 | 2003-03-07T16:17:20 | Chandra Observations of the Mice | [
"astro-ph"
] | Presented here are high spatial and spectral resolution Chandra X-ray observations of the famous interacting galaxy pair, the Mice, a system similar to, though less evolved than, the well-known Antennae galaxies. Previously unpublished ROSAT HRI data of the system are also presented. Starburst-driven galactic winds outflowing along the minor axis of both galaxies (but particularly the northern) are observed, and spectral and spatial properties, and energetics are presented. That such a phenomenon can occur in such a rapidly-evolving and turbulent system is surprising, and this is the first time that the very beginning - the onset, of starburst-driven hot gaseous outflow in a full-blown disk-disk merger has been seen. Point source emission is seen at the galaxy nuclei, and within the interaction-induced tidal tails. Further point source emission is associated with the galactic bar in the southern system. A comparison of the source X-ray luminosity function and of the diffuse emission properties is made with the Antennae and other galaxies, and evidence of a more rapid evolution of the source population than the diffuse component is found. No evidence for variability is found between the Chandra and previous observations. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 4 June 2018
(MN LATEX style file v2.2)
Chandra Observations of the Mice
Andrew M. Read1
1 School of Physics and Astronomy, University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK
(E-mail: [email protected])
Accepted ..............................; Received ..............................; in original form ..............................
ABSTRACT
Presented here are high spatial and spectral resolution Chandra X-ray observa-
tions of the famous interacting galaxy pair, the Mice, a system similar to, though less
evolved than, the well-known Antennae galaxies. Previously unpublished ROSAT HRI
data of the system are also presented.
Starburst-driven galactic winds outflowing along the minor axis of both galaxies
(but particularly the northern) are observed, and spectral and spatial properties, and
energetics are presented. That such a phenomenon can occur in such a rapidly-evolving
and turbulent system is surprising, and this is the first time that the very beginning
− the onset, of starburst-driven hot gaseous outflow in a full-blown disk-disk merger
has been seen.
Point source emission is seen at the galaxy nuclei, and within the interaction-
induced tidal tails. Further point source emission is associated with the galactic bar in
the southern system. A comparison of the source X-ray luminosity function and of the
diffuse emission properties is made with the Antennae and other galaxies, and evidence
of a more rapid evolution of the source population than the diffuse component is found.
No evidence for variability is found between the Chandra and previous observations.
Key words: galaxies: individual: NGC4676 − galaxies: starburst − galaxies: ISM −
galaxies: haloes − X-rays: galaxies − ISM: jets and outflows
1
INTRODUCTION
Though galaxies were once thought of as 'Island Universes',
evolving slowly in complete isolation, this is now known not
to be the case. Galaxies interact in a wide variety of ways
with their environment, and collisions and mergers of galax-
ies are now believed to be one of the dominant mechanisms
in galactic evolution (Schweizer 1989). There are probably
very few galaxies today that were not shaped by interactions
or even outright mergers. Indeed, Toomre's (1977) hypoth-
esis, whereby elliptical galaxies might be formed from the
merger of two disc galaxies, is now generally accepted, such
behaviour having been modelled in many N-body simula-
tions of mergers (e.g. Toomre & Toomre 1972; Barnes 1988).
During such an encounter, the conversion of orbital to inter-
nal energy causes the two progenitor disks to sink together
and coalesce violently into a centrally condensed system.
The 'Toomre sequence' (Toomre 1977) represents probably
the best examples of nearby ongoing mergers, from disk-disk
systems to near-elliptical remnants.
One of these is the famous binary interacting system
NGC4676A/B (also Arp242). First described by Vorontsov-
Vel'yaminov (1957) as a pair of "playing mice", ever since,
the name 'the Mice' has become universally accepted. It con-
sists of two distinct spiral galaxies (the edge-on NGC4676A
to the north, and the more face-on NGC4676B to the south)
with long tidal tails, and the presence of these, along with an
obvious bridge between the two galaxies and an increase in
IR activity compared to typical field galaxies, indicates that
the galaxies have begun to interact, and have passed one an-
other. It is one of the original systems presented by Toomre
& Toomre (1972) as a classic example of a pair of galaxies
undergoing tidal interaction. It lies second in the proposed
evolutionary sequences of both Toomre (1977) − the Toomre
sequence, and of Hibbard & van Gorkom (1996). It is also
placed second in the first study of the X-ray properties of an
evolutionary sample of merging galaxies; the ROSAT work
of Read & Ponman (1998) (hereafter RP98).
The systematic velocity of the system appears well-
defined in several different regions of the electromagnetic
spectrum (e.g. Stockton 1974; Smith & Higdon 1994; Hib-
bard & van Gorkom 1996; Sotnikova & Reshetnikov 1998),
and this converts, assuming H0 = 75 km s−1 Mpc−1, to the
distance to the Mice used in this paper; 88 Mpc. It is gen-
erally agreed, using both kinematical work (e.g. Toomre &
Toomre 1972; Mihos, Bothun & Richstone 1993) and multi-
wavelength observations (e.g. Hibbard & van Gorkom 1996),
that both galaxies are in the throes of a prograde encounter,
2
A.M. Read
and have their northern edges moving away from us. Hence,
NGC4676A's tail is on the very furthest side, swinging away
from us (the tail has a systematic velocity almost 300 km
s−1 faster than the NGC4676A nucleus; Sotnikova & Reshet-
nikov 1998), and NGC4676B is rotating clockwise, with its
northeastern edge lying closest to us.
Both galaxies appear (e.g. from the optical data of
Schombert, Wallin & Struck-Marcell 1990) to have shapes
and colours consistent with those of early-type spirals,
though the disc regions are strongly distorted or absent.
Both tails, although bluer than the galaxies' central colours,
are in agreement with the colours of outer-disc regions. The
northern tail has a very high surface brightness (µB = 23.0−
23.5 per square arcsecond) when compared with similar fea-
tures in other galaxies, whereas the southern tail is roughly
1 magnitude per square arcsecond fainter (Schombert et al.
1990). The tails account for 16% of the total Hα emission,
are quite luminous, containing one-third of the total R-band
luminosity of the system, and have a high atomic gas con-
tent (Hibbard & van Gorkom 1996). Local Hα maxima are
also observed in the northern tail (Sotnikova & Reshetnikov
1998). The northern galaxy appears to exhibit a 6.6h−1 kpc
plume of Hα along its minor axis, and the southern galaxy
possesses an ionized gas bar, as produced in Barnes & Hern-
quist's (1991, 1996) merger simulations, offset with respect
to the stellar bar (Hibbard & van Gorkom 1996). Angular
momentum transfer between the two bars is able to force
large amounts of gas towards the galactic centre. Recent
mapping of the CO emission in the Mice (Yun & Hibbard
2001) detect a compact ( <
∼2 kpc), disklike or ringlike com-
plex, perhaps accounting for 20% of the total nuclear mass,
centered on the inner stellar disk of NGC4676A, with kine-
matics consistent with simple rotation. Within NGC4676B,
the CO emission is far less bright and occurs along the 7 kpc
stellar bar.
A very relevant work is the set of papers dealing with
the Chandra X-ray emission from the Antennae (Fabbiano,
Zezas & Murray 2001, Zezas et al. 2002, Zezas & Fabbiano
2002), a very similar system to the Mice, in that it involves
the merger of two equal-sized gas rich spiral galaxies. The
Antennae however lie at a later evolutionary epoch, in that
the two galactic disks have begun to interact violently with
each other. Also, the Antennae lie at a much closer dis-
tance than the Mice, aiding the detection of low luminosity
sources, and decreasing the effects of source confusion. Vari-
ous aspects of the two systems' X-ray emission are compared
throughout this paper.
In this paper results of a ∼30.5 ks Chandra ACIS-S ob-
servation of the Mice are presented. The fundamental advan-
tage of Chandra over any other previous or current X-ray
mission, is its excellent spatial resolution. With a spatial
resolution some 10 times better than the ROSAT HRI and
XMM-Newton, it is possible to resolve emission on scales
down to ∼210 pc (≈ 70% encircled energy PSF [on-axis at
<6 keV]) at the distance of the Mice. This allows the removal
of point source emission, and the analysis of any unresolved,
perhaps diffuse emission. In addition, the spectral resolution
of Chandra is comparable to that of ASCA, and far better
than that of the ROSAT PSPC (the ROSAT HRI note, hav-
ing essentially no spectral resolution whatsoever), allowing
us to study the X-ray spectral properties of the sources and
Figure 1. Contours of ROSAT HRI emission over the full channel
range superimposed on an optical Digital Sky Survey (DSS) image
of the Mice. The 2′′ resolution X-ray image has been smoothed
with a Gaussian of FWHM 9.4′′. Contours are at values of 2, 3,
5 and 9σ (σ being 52.4 counts/arcmin2) above the background
(157.3 counts/arcmin2).
emission regions seen, providing additional information as
to their nature.
The X-ray observations of the Mice prior to Chandra
are described in the following subsection. Section 2 describes
the Chandra observations and the data reduction techniques
used. Discussion of the spatial, spectral and temporal prop-
erties of the source and diffuse emission components follow
in Section 3, and in Section 4, the conclusions are presented.
1.1 Previous X-ray Observations
The Mice have only previously been observed in X-rays with
ROSAT. The PSPC observations were described in RP98,
and show rather an amorphous X-ray structure, with only
one source detected, lying between the two galactic nuclei.
Little could be said as regards the spectral properties of
the source. The general form of the PSPC emission how-
ever, seemed to follow the optical 'heads' of the two galax-
ies, running essentially north-south, with some tentative ev-
idence for extension in the east-west direction, especially
around NGC4676A. Though only a small number of counts
were obtained, the fitted temperature to the diffuse emis-
sion spectrum was seen to be in good agreement with fitted
temperatures of nearby, known, starburst winds (Heckman
1993; Read, Ponman & Strickland 1997), suggesting a star-
burst origin for the diffuse emission.
The Mice were also observed with the ROSAT HRI.
The previously unpublished results of these observations are
presented here, as it serves both as a useful introduction to
the system and its X-ray structure, and as a stepping-stone
from the (relatively) low-resolution of the ROSAT PSPC to
the high resolution of Chandra.
The 33.4 ks of ROSAT HRI data (ID 601091), taken at
the end of 1997, were initially screened for good time inter-
vals longer than 10 s. No further selection on low background
times was made, as we were mainly interested in point-like
sources (which the HRI is far better suited to), and for this
purpose, the data were still photon-limited. Source detection
and position determination was then performed over the full
field of view with the EXSAS local detect, map detect and
maximum likelihood algorithms (Zimmermann et al. 1994),
using images of pixel size 5′′.
Fig. 1 shows smoothed contours of ROSAT HRI emis-
sion (over the full channel range) from the Mice super-
imposed on an optical Digital Sky Survey (DSS) image.
Though there are very few counts, it is clearly seen that
the HRI is able to resolve the PSPC emission into two dis-
tinct sources at the positions of the two galaxies. These
were the only two sources detected in the vicinity of the
optical galaxies. The northern source is detected (with
22 source counts at α[2000.0] = 12h46m9.95s δ[2000.0] =
+30◦43′54.3′′) at a significance of 5.9σ with a (0.1−2.4 keV)
count rate of (6.8±1.7)×10−4 counts s−1, the southern
source (with 10 source counts at α[2000.0] = 12h46m11.04s
δ[2000.0] = +30◦43′21.4′′) at 3.5σ with a count rate of
(3.1±1.2)×10−4 counts s−1. Errors on the HRI positions are
1.4−1.6′′. Further, perhaps extended emission is suggested
to the north and east of NGC4676A, and this is borne out by
a low-significance extent within the source-searching results
to the northern source's emission.
2 CHANDRA OBSERVATIONS, DATA
REDUCTION AND RESULTS
The Mice were observed with Chandra on May 29th, 2001
for a total of just over 30 ks, with the back-illuminated ACIS-
S3 CCD chip at the focus (Observation ID: 2043). Data
products, correcting for the motion of the spacecraft and
applying instrument calibrations, were produced using the
Standard Data Processing (SDP) system at the Chandra X-
ray Center (CXC). These products were then analysed using
the CXC CIAO software suite (version 2.2.1). A lightcurve
extracted from a large area over the entire observation was
seen to be essentially constant and consistent with a low-
level rate. Consequently, no screening to remove periods of
high background flaring was performed.
2.1 Overall X-ray structure
shows
Fig. 2 (left)
contours of adaptively smoothed
(0.2−10 keV) Chandra ACIS-S X-ray emission from the field
surrounding the Mice system, superimposed on an image
from the Hubble Space Telescope's Advanced Camera for
Surveys (ACS) instrument. The adaptive smoothing of the
X-ray emission attempts to adjust the smoothing kernel to
obtain a constant signal-to-noise ratio across the image.
Several things are immediately evident from the im-
age. As suggested by the HRI data, the emission is resolved
into two main components associated with the two galaxy
'heads'. The factor 10 higher resolution over the HRI how-
ever, is able to shed far more light on the situation. The
emission associated with NGC4676A is centered close to the
galaxy nucleus, and appears predominately diffuse and ex-
tended, especially in the east-west direction (i.e. along the
Chandra Observations of the Mice
3
minor axis of the galaxy, NGC4676A being seen almost ex-
actly edge on). The X-rays within NGC4676B appear far
more centrally concentrated at the nucleus, though some
evidence for extension is also seen again along NGC4676B's
minor axis (i.e. from south-east to north-west). Orthogo-
nal to this, hotspots of X-ray emission are seen along the
galaxy disk, equidistant from the nucleus, especially to the
south-west. Other interesting features are seen associated
with the tails, one source lying directly in the northern tail,
and a further source lying to the east of the southern tail.
No significant features are seen associated with the more
distant, off-image tail regions. One last interesting feature is
the source seen ∼40′′ west of NGC4676A.
Fig. 2 (right) shows a true colour X-ray image, with red
corresponding to 0.2−0.9 keV, green to 0.9−2.5 keV and blue
to 2.5−10 keV. The individual adaptively-smoothed images
were obtained as described above. The image is to the same
scale as Fig. 2 (left), and all the X-ray features are visi-
ble. The northern head shows a medium energy nucleus,
with cooler, complex, structured emission surrounding. The
southern head is brighter and hotter than the surrounding
emission which again is cool (white indicates strong emission
in all three bands). The nearby feature SW of the southern
head appears soft, as does the feature in the northern tail.
The feature just east of the southern tail and the source west
of the system appear somewhat harder.
2.2 Point sources: spatial and spectral properties
The CIAO tool wavdetect was used to search for point-like
sources, on scales from 1−16 pixels (0.5−8′′). A total of 6
sources were detected in the 0.3−10 keV band within or close
to the optical confines of the galaxies (whether the galaxy
'heads' or tails), and their X-ray properties are summarized
in Table 1. Sources are detected (see Fig. 3) at the positions
of the two galaxies (A and B), and a further source (B2)
is detected just south-west of NGC4676B. A source (N) is
detected within the northern tail, and a source (S) is de-
tected close to the eastern edge of the southern tail. Finally,
a source (W) is detected west of the the main body of the
system. The X-ray properties given in Table 1 are as fol-
lows; Right Ascension and Declination (2000.0) are given in
cols. 2 and 3, together with the positional error (in arcsec-
onds) in col. 4. Net source counts (plus errors) are given in
col. 5, and the source significance is given in col. 6. Hardness
ratios (plus errors) are given in col. 7. These were calculated
by performing additional detection runs (using again scales
of 1−16 pixels) in a soft (0.3-2 keV) and a hard (2-10 keV)
band. The tabulated values are (H −S)/(H +S), H being the
net source counts in the hard band, S being the net source
counts in the soft band. Finally, (0.3−10 keV) X-ray emitted
and intrinsic (i.e. corrected for absorption) luminosities (ob-
tained as described in the spectral fitting section) are given
in cols. 8 and 9.
Source spectra were extracted in the range 0.3−10 keV
at the exact positions given by the 0.3−10 keV detection
analysis. The regions output by the detection routines, de-
fined to include as many of the source photons as possible,
but minimizing the background contamination, were invari-
ably near-circles of radius ∼ 5 pixels (partly due to the
Mice only occupying the very centre of the ACIS-S3 chip),
and consequently, extraction circles of radius 5 pixels (2.5′′)
4
A.M. Read
Figure 2. (Left) Contours of adaptively-smoothed (0.2−10 keV) Chandra ACIS-S X-ray emission from the field surrounding the Mice
system, superimposed on a Hubble Space Telescope ACS image. The X-ray contours increase by factors of two. For scale, the X-ray
sources at the two galaxy 'heads' lie ≈36′′ apart. (Right) 'True colour' X-ray image of the Mice to the same scale. Red corresponds to
0.2−0.9 keV, green to 0.9−2.5 keV and blue to 2.5−10 keV.
Table 1. Sources detected by wavdetect in the 0.3−10 keV band within or close to the optical confines of the Mice. Columns are described
in the text. Luminosities assume (except for source B; see text) a photon index 1.5 power law plus Galactic absorption, and a distance
of 88 Mpc.
Src.
RA
Dec.
(2000.0)
Pos.err.
(arcsec)
Counts(err)
Sig.
HR
LX (0.3−10 keV)
(1039 erg s−1)
(emitted)
(intrinsic)
12 46 10.09 +30 43 55.8
A
12 46 11.23 +30 43 22.0
B
12 46 11.08 +30 43 15.8
B2
12 46 10.36 +30 44 22.2
N
12 46 13.62 +30 43 03.3
S
W 12 46 07.05 +30 43 49.4
0.22
0.12
0.23
0.14
0.13
0.14
26.52± 5.29
92.98± 9.80
10.29± 3.31
27.25± 5.29
17.36± 4.24
23.26± 4.90
10.64
31.59
4.66
12.25
7.96
10.47
-0.45±0.27
-0.48±0.12
-0.66±0.36
-0.59±0.21
-0.48±0.36
-0.44±0.23
11.92
23.99
3.22
8.72
4.98
7.82
12.17
26.14
3.30
8.91
5.04
7.98
were used for all the sources in Table 1. Background regions
were source-free circular annuli surrounding each source (in
order to minimize effects related to the spatial variations
of the CCD response), though regions of apparent diffuse
emission were also avoided in the extraction of background
regions; in the cases of A, B and B2, background extrac-
tion annuli of inner-to-outer radii 15−22.5′′ were used. In
the cases of N, S and W, background extraction annuli of
2.5−12.5′′ were used. All source and background extraction
regions are shown in Fig. 3.
ACIS spectra were extracted using Pulse Invariant (PI)
data values, and were binned together to give a minimum
of 10 counts per bin after background subtraction. Hence
χ2 statistics could be used. Response matrices and ancillary
response matrices were created for each spectrum, using the
latest calibration files available at the time of writing.
Standard spectral models were fit to the spectral data
using the CIAO spectral fitting software. Events above 7 keV
(of which there were very few) and below 0.3 keV were ex-
cluded from the fitting on the grounds of uncertainties in
Chandra Observations of the Mice
5
Figure 4. 99%, 90% and 68% confidence contours in the pho-
ton index-absorption column plane for the power-law fit to the
spectrum of source B.
source B assume this best-fitted model, while for the other
sources, the model assumed is of fixed (Galactic) absorption
plus a power-law of photon index 1.5.
2.3 Residual emission: spatial and spectral
properties
The existence of residual, likely diffuse emission, surround-
ing the two head sources A and B is very evident in the
figures. Spectra from these regions of apparent diffuse emis-
sion were extracted in the range 0.3−10 keV. An annulus of
2.5−15′′ centred on source A was used to extract a spectrum
of the 'diffuse' emission associated with NGC4676A (here-
after 'diffA'). An identical annulus around source B (except-
ing a 2.5′′ circular region surrounding source B2) was used
to extract a spectrum of the emission around NGC4676B
('diffB'). Background regions were as for sources A and B.
These diffuse and background extraction regions are shown
in Fig. 3.
The spectral fitting was performed as for the point
sources, using the same models, and using both methods to
correct for the degradation in the ACIS QE. With ≈80−100
non-background counts in each spectrum (see Table 2), it
was possible to place some constraints on the spectral prop-
erties of the residual emission. In both cases, diffA and diffB,
an absorption plus power-law model was unable to fit the
data satisfactorily. Only reduced χ2's of 1.5 (diffA) and 1.4
(diffB) were attainable using power-law models. Thermal
models proved much better, and the best thermal fits (again
using an absorption plus mekal model) are summarized in
Table 2; the absorbing column, the fitted temperature and
metallicity (an 'F' indicating a fixed value), the reduced χ2,
and the emitted and intrinsic (i.e. absorption-corrected) X-
ray luminosity is given. Essentially identical temperatures
(≈ 0.5 ± 0.2 keV) are obtained for both diffuse emission fea-
tures, and the data plus best fit models are shown in Fig. 5.
Figure 3. Adaptively-smoothed (0.2−10 keV) ACIS-S image
(scale as for Fig. 2) labelled with sources detected in the
0.3−10 keV band. Circles show the source, diffuse and background
regions used in the extraction of the spectra (see text).
the energy calibration. It is now known that there has been
a continuous degradation in the ACIS QE since launch. A
number of methods now exist within the community to cor-
rect for this. These include the release of an XSPEC model
(ACISABS) to account for this degradation, and the exis-
tence of software (corrarf) to correct the ancillary response
files. Both methods have been used here in the spectral fit-
ting, and very similar results were obtained. In both cases,
the time since launch of the observations (here, 678 days) is
used in the correction. Although the calibration at energies
below 1.0 keV is believed to be uncertain, data in this range
were kept, as the statistical error on these data points is
still greater than the errors due to the uncertainties in the
calibration.
Two models, one incorporating absorption fixed at the
value out of our Galaxy (1.28×1020 cm−2) and a 5 keV mekal
thermal plasma, the other incorporating absorption (again,
fixed) and a power-law of photon index 1.5 were fit to the
data. Only for source B were there sufficient counts to let
the model parameters fit freely. A good fit (reduced χ2 <
∼1)
was obtained using a power-law model of photon index 1.52
and an absorbing column of 4.19×1020 cm−2. Fig. 4 shows
the 99%, 90% and 68% confidence contours in the photon
index-absorption column plane for the power-law fit to the
spectrum of source B. The luminosities quoted in Table 1 for
6
A.M. Read
Table 2. Best results of fitting thermal models to the spectra of residual emission around sources A and B. Luminosities assume a
distance of 88 Mpc (see text).
Diff.
Src.
diffA
diffB
Counts(err)
NH
(1020 cm−2)
kT
(keV)
Z
χ2
(solar)
(red.)
LX (0.3−10 keV)
(1039 erg s−1)
(emitted)
(intrinsic)
103.4±12.7
80.55±11.9
1.28(F)
1.28(F)
0.50+0.18
−0.13
0.46+0.20
−0.12
0.3(F)
0.3(F)
1.20
0.45
10.13
7.34
10.86
7.89
calculated for the ROSAT HRI count rates, using the ACIS-
S best-fit models to the source A and B spectra. Compar-
ison of the HRI and ACIS-S PSFs indicates that the HRI
would essentially 'see' source A and most of the residual
emission around source A (diffA) as one point source. Simi-
larly, the HRI would detect the emission from sources B, B2
and from diffB as one source. One can therefore only study
the temporal variation of the emission from the 'heads' of
the two galaxies; NGC4676A (where the total ACIS-S (in-
trinsic) LX is the sum of the X-ray luminosities of source A
and diffA) and NGC4676B (the total LX being the sum
for sources B, B2 and diffB). In Fig. 6, these two crude X-
ray lightcurves are shown. Errors on the luminosities are
taken as the statistical errors on the count rates (the er-
rors on the HRI luminosities are very large, due to there
being very few counts, and a higher background level). For
the southern galaxy NGC4676B, the level of the emission
appears not to have varied at all between the two observa-
tions. If one uses a power-law model to calculate the HRI
luminosity of NGC4676A (the upper 1997 point in the left-
hand panel of Fig. 6), then it appears that some variation
may have occurred (at least at the ∼ 2σ level). However, re-
calculating an intrinsic (0.3−10 keV) X-ray luminosity for
the NGC4676A ROSAT HRI count rate, but using instead
the best-fit model to the diffA spectrum, results in a value
(the lower 1997 point in the left-hand panel of Fig. 6) very
in accordance with the ACIS-S value. The true HRI lumi-
nosity lies somewhere between these two points, and one can
conclude therefore that there are no significant signs of any
X-ray variations between the two observations.
3 DISCUSSION
3.1 Point sources
With the Chandra ACIS-S instrument, it is possible to re-
solve more clearly the source populations within the Mice,
that were suggested by the ROSAT HRI observations.
Complex emission regions are seen at the centres of the
two galaxies, surrounded by what appears to be residual,
diffuse X-ray emission, and further high-luminosity sources
are detected close to the southern nucleus and within the
tidal tails.
Recent mapping of the 2.6 mm CO J = 1 → 0 emis-
sion in the Mice (Yun & Hibbard 2001) has revealed large
amounts of molecular gas. A compact molecular complex is
seen well centred on the inner stellar disk of NGC4676A,
forming a disk- or ring-like structure, seen within 5◦−10◦of
being edge-on, with a deconvolved size of 1.8 kpc in radius
and a thickness (FWHM) of 250 pc. The gas kinematics ob-
Figure 5. Data (points) plus best-fit thermal mekal models
(lines) for the residual emission around source A (top) and
source B (bottom).
2.4 Temporal properties
As we have two observations of the Mice where one can
distinguish discrete emission regions (the first being the
ROSAT HRI observations, described in Sect. 1.1), some con-
straints can be placed on the temporal properties of those re-
gions. X-ray luminosities in the 0.3−10 keV band have been
NGC4676A
NGC4676B
Lx (0.3−10 keV) (10e39 erg/s)
Lx (0.3−10 keV) (10e39 erg/s)
60
50
40
30
20
60
50
40
30
20
1998
1999
2000
2001
1998
1999
2000
2001
Year
Year
Figure 6. Crude X-ray (intrinsic) luminosity lightcurves of the
emission regions associated with the 'heads' of the two galax-
ies; NGC4676A (left) and NGC4676B (right). The two 1997
NGC4676A points refer to the usage of two different models (see
text).
served are consistent purely with rotation. A total molecular
gas mass of 5.5×109 M⊙ is detected (about twice as much
as in our Galaxy), making up a significant fraction (20%)
of the total mass in the nuclear region. This is a very high
fraction for ordinary disk galaxies (Young & Scoville 1991).
In NGC4676B, the situation is different in that only
weak CO emission is detected, and then only along the
7 kpc stellar bar. Large peaks in the CO emission are seen in
NGC4676B at the ends of the bar, as is often seen in other
barred galaxies. The CO kinematics are consistent with solid
body rotation of the bar, and the molecular gas fraction is
seen to be about 7%, which is more typical of undisturbed
disk galaxies.
The fact that several smaller CO clumps are seen in
the bridging region between the two Mice 'heads' indicates
that the disruption of the inner disks has begun within the
Mice. This, along with the obvious morphological indicators
of interaction, and the fact that an increase in far-infrared
luminosity is observed within the system, is very suggestive
of strong starburst activity having begun within the one or
both of the galactic nuclei. The Mice have LF IR/LB values
more like those of field starburst galaxies, than of normal
galaxies, and a large value of the far-infrared dust colour
temperature S60/S100, a diagnostic seen to increase with
interaction strength (Telesco et al. 1988), is also observed
within this system (RP98).
Significant X-ray emission is seen at the centres of
the two galaxies. Dealing first with source at centre of
NGC4676B, source B, it is possible to place some constraints
as to the spectral properties of the source, in that an ab-
sorbed power-law is seen to fit the data very well. This is
not too constraining however, as the fit is not significantly
better than with an absorbed (9.9×1020 cm−2) fixed-5 keV
mekal model. Though it is difficult to say whether the emis-
sion at the centre of NGC4676B is thermal or not, it appears
Chandra Observations of the Mice
7
Figure 7. 0.2−10 keV ACIS-S X-ray emission (white contours,
as per Fig. 2a) superimposed on the Yun & Hibbard (2001) figure
of velocity-integrated CO (1−0) maps (black contours) and Hα
(greyscale).
to be rather absorbed, a fact not too surprising, given the
inclined orientation nature of the southern galaxy.
Fig. 7 shows white contours of 0.2−10 keV ACIS-S X-
ray emission (as per Fig. 2a) superimposed on Yun & Hi-
bbard's (2001) figure of velocity-integrated CO (black con-
tours) and Hα (greyscale). Fig. 8 details the Hα emission
and shows contours of Hα (increasing by factors of 1.7783),
from Hibbard & van Gorkom (1996) (and reproduced in Yun
& Hibbard (2001)), superimposed on the 0.2−10 keV ACIS-
S X-ray emission (greyscale).
Source B is very bright, appears very compact, and lies
coincident with the peak in the southern galaxy's Hα emis-
sion, a site of intense current star-formation. Little CO emis-
sion is seen here however. In contrast, source A, the X-ray
source at the centre of NGC4676A, is not as bright, nor
as centrally compact, and appears coincident with the peak
in the CO emission disk, a site containing large amounts
of molecular gas. There is a peak in the Hα emission at
the position of source A, indicating that a good deal of
star-formation is ongoing, though the main Hα peak in the
northern galaxy lies some 13′′ to the south, at the edge
of the CO disk. Yun & Hibbard (2001) suggest that this
near anticorrelation between the CO and the Hα emission
in the northern galaxy is due to extinction in the nearly
edge-on disk. In fact the peak CO flux corresponds to an
NH of ≈ 6 × 1022 cm−2, and this may explain why source A
appears rather different to source B in X-rays, i.e. we are
not really seeing many of the X-rays (or indeed the Hα
photons), because of the intervening absorbing material in
8
A.M. Read
Figure 8. Contours of Hα (increasing by factors of 1.7783),
from Hibbard & van Gorkom (1996), superimposed on 0.2−10 keV
ACIS-S X-ray emission (greyscale).
the disk. One cannot say anything using source A's X-ray
spectrum as regards absorption, but assuming again a 1.5
power-law model, but with an absorption suggested by the
CO data, then an intrinsic (0.3−10 keV) X-ray luminosity
of 2.1×1040 erg s−1 is obtained, more similar to that of
source B. Sources A & B may indeed be very similar.
The fact that the X-ray emission at the centres of the
two galaxies appears spatially complex, with associated dif-
fuse hot gas structures, and significant amounts of molecular
gas and Hα emission, suggests strongly (especially for source
A) that the nuclear X-ray sources are starburst-like in ori-
gin (as opposed to being significantly dominated by AGN
activity).
Such X-ray luminosities for the nuclear sources in the
Mice are certainly 'Super-Eddington' (the Eddington limit
for a 1M⊙ accreting object is ∼ 1.3 × 1038 erg s−1), but
these nuclear sources may not be single sources, and we
would need even higher spatial resolution than that offered
by Chandra to untangle any complex source confusion. It is
interesting to note that the luminosities of sources A & B
are somewhat greater than those of the two nuclear sources
in the Antennae (Zezas et al. 2002). Though there may be
single massive accreting 'active' sources at the centres of
these two galaxies, more likely is that these sources are col-
lections of supernovae and stellar winds plus some smaller
accreting sources, all embedded in diffuse hot gas − a star-
burst. The fact that much apparently diffuse X-ray emission
is seen surrounding the nuclei certainly supports this. This
diffuse emission is discussed in the next section.
soft source according to its hardness ratio (Table 1) and
its appearance in the' true colour' image (Fig. 2b). It was
not detected by ROSAT, on account of its dimness and close
proximity to the southern nuclear source. No noteworthy Hα
features are associated with this feature (Fig. 8), excepting
that B2 lies along the general disk of Hα that follows the
galactic disk. More interesting is the comparison with the
CO emission (Fig.7) where it is seen that B2 lies coincident
with the CO feature corresponding to the south-western end
of the NGC4676B bar (also both B2 and the correspond-
ing CO feature show some extension to the northwest). A
less bright X-ray feature (not formally detected within the
source-searching analysis) is seen on the opposite side of the
southern nucleus, at the other end of the bar, and coinci-
dent with a smaller knot of CO emission. Gas dynamical
models of barred galaxies (e.g. Engelmaier & Gerhard 1997)
show strong gas accumulation at the ends of bars, due to
corotation of the bar structure with the disk, leading to en-
hanced star formation. The CO emission and the general
Hα structure here seem to support this, and source B2 and
the X-ray feature on the other side of the bar are very likely
due to this bar-enhanced star-formation. Similar behaviour
has been seen in a number of other barred galaxies (e.g.
NGC4303; Tschoke, Hensler & Junkes 2000).
Two sources are also seen in the tidal tails; N in the
northern tail, and S at the eastern edge of the southern.
Interestingly, both sources sit at similar distances along the
tails from their respective 'heads'. No other significant X-ray
features are seen associated with the tidal tails. An enhance-
ment in the ROSAT PSPC contours is seen at the position of
N (RP98), and, though nothing significant is seen at the po-
sition of S, one would not expect to see anything significant,
given the brightness of S. The ROSAT HRI image (Fig. 1)
shows an interesting, though very low significance, plume to
the north of NGC4676A. None of the features within the
plume however, coincide with the position of N, and, given
the ACIS-S count-rate of N, we would be very fortunate to
detect any emission in the HRI. Nothing is seen in the HRI
at the position of S.
N lies well within the northern tail, about 1.5′′ east of
the centre of the tail. Though many Hα features are seen
within both tails (Hibbard & van Gorkom 1996), no bright
knots lie particularly close to N (the nearest is relatively
dim, and lies some 1.5′′ away). Interestingly, a knot in the
HST ACS image is seen poking eastwards out of the main
central shaft of the tail at the position of N. Furthermore
there appears to be a neutral hydrogen H i enhancement at
the position of N (Yun & Hibbard 2001). S lies on the very
eastern edge of the southern tidal tail. No significant opti-
cal feature, either within the tail or more isolated, appears
coincident in the HST ACS image, or in coarser Digital Sky
Survey images. No Hα features appear coincident either.
Finally, a bright star lies close to Source W (visible at
the very right-hand edge of Fig. 2), but is too far offset to
be associated. Instead, a very faint optical counterpart can
be seen at the exact position of source W in the HST ACS
image. The ROSAT PSPC contours are seen to stretch over
to encompass this source (RP98), and a feature is seen at
this position in the ROSAT HRI data (Fig. 1). This source
is assumed not to be associated with the Mice, and is likely
a background AGN.
Close by to the southern nucleus is source B2, a rather
Indeed, one can use the the logN−logS relation of Gi-
acconi et al. (2001) to estimate the expected number of
sources not physically associated with the Mice. At most,
only one background source is expected over the area cov-
ered by Figs. 2 & 3 (∼ 3.2 sq. ′) at the detection limit seen
here (5 × 1039 erg s−1 at 88 Mpc). It may also be the case
that Source S has nothing to do with the Mice, given that it
lies some distance from the southern tail, but for source N,
lying as it does, directly on the thin northern tail, the case
is more concrete.
As a last point, it is worth mentioning that Chandra
is able to throw some light upon some of the curious X-ray
features previously seen by ROSAT, but not associated with
the Mice. The strange extension to the south-west (away
from the tidal tails) seen in the PSPC image (RP98) lies
towards the direction of two dim Chandra X-ray sources; a
≈ 40 ct source at α = 12h46m8.30s, δ = +30◦41′55.5′′, and
a ≈ 20 ct source at α = 12h46m6.06s, δ = +30◦42′24.2′′. A
final interesting source is the very dim (≈ 8 ct) source at
α = 12h46m10.42s δ = +30◦46′39.3′′, lying directly along
the line of the northern tail, but approximately 1 arcminute
beyond the optical end of the tail.
>
Though only a small number of sources are evident
within the Mice, an approximate comparison of the source
X-ray luminosity function (XLF) with the XLFs of other
galaxies can be made. This has recently been performed for
the Mice's more evolved and nearby counterpart, the Anten-
nae (Zezas & Fabbiano 2002). At the distance of the Mice,
∼5×1039 erg s−1.
one is only able to detect sources with LX
This in itself is important; all five sources detected within
the Mice are ultraluminous. This high tail to the XLF is
in sharp contrast to the Galaxy, and to a vast majority of
nearby normal and starburst galaxies (e.g. Read & Pietsch
2001, Zezas & Fabbiano 2002). It does however, appear
broadly similar to the Antennae. The eight sources in the
Antennae with LX > 5 × 1039 erg s−1 may be significantly
greater than the five in the Mice, though there are a num-
ber of notes and caveats. Zezas & Fabbiano (2002) use a
distance to the Antennae calculated using H0 = 50 km s−1
Mpc−1. Using H0 = 75 km s−1 Mpc−1, the eight high-LX
Antennae sources would reduce to around five. Using an An-
tennae distance of 25 Mpc, as in RP98, the number would be
6−7. Furthermore, the Antennae is slightly more massive (in
terms of LB; RP98), and has a correspondingly higher SFR
(in terms of LF IR; RP98) by similar factors of ≈ 1.3. Activ-
ity indicators, such as far-infrared dust colour temperature
S60/S100, and (naturally, from above) LF IR/LB are almost
identical for the two systems. Scaling the XLFs by mass (or
SFR) would bring the two systems into alignment. Note that
the usage of the optical luminosity LB as a measure of mass
is generally not ideal, as the blue luminosity to mass ratio is
quite sensitive to the age of a stellar population. Here how-
ever, as we are dealing with two very similar systems, the
comparison is quite justified (the usage of various mass and
activity indicators for galaxies is discussed in Read & Pon-
man 2001). Though it would seem therefore, that in terms
of the high end of the XLF, the Mice and the Antennae may
appear rather similar, a final caveat to consider is that it is
likely, even with the excellent spatial resolution of Chandra,
that source confusion is prevalent in the more distant Mice.
Such confusion can only strengthen the tail of the XLF, and
the likely true situation is that there are fewer than five
high-LX sources in the Mice. It is difficult to be conclusive,
Chandra Observations of the Mice
9
but it appears that there are significant number of high-LX
sources in the Mice, a number probably less than is seen in
the more evolved Antennae. If the Mice is to evolve into a
system like the Antennae, then it is expected that the num-
ber of high-LX sources in the Mice should increase slightly
with time.
3.2 Diffuse emission
There exists much evidence that starburst-driven diffuse
emission features are seen around both galaxies within the
Mice. As discussed with regard to the point sources, there
is very little doubt that a good deal of star-formation is
taking place at the two nuclei. Around the two nuclei, one
sees extended, clumpy, structured emission that is seen to be
soft and well-fitted with a thermal (0.5 keV) spectral model.
Furthermore, emission appears to be significantly more ex-
tended along the minor axis of each galaxy (this is more
evident in the northern galaxy, due likely to NGC4676A ly-
ing almost exactly edge-on). Furthermore, in the northern
galaxy, the Hα emission is also seen to be significantly ex-
tended along the minor axis (especially to the west; Fig. 8).
All this is very reminiscent of these features being starburst-
driven galactic winds, as seen in famous nearby starburst
galaxies such as M82 and NGC253 (e.g. Strickland et al.
2002; Bravo-Guerrero, Read & Stevens 2002). Such nearby
systems however, are rather isolated, and classic bipolar
winds in even one member of a strong, rapidly-evolving
interacting pair such as the Mice, is something that has
not been seen before. It is believed from ROSAT (RP98)
that starburst- and star formation-driven diffuse emission
is prevalent in interacting and merging galaxies, but this
was only seen in mid-stage (e.g. the Antennae) and ultralu-
minous (nuclear contact) mergers, not at early (e.g. Mice)
stages. In these post-Mice cases, the systems are evolved to
such a degree, that any classic starburst winds (were they to
have existed), would have been distorted out of recognition,
in agreement with observed peculiar morphologies (RP98).
More recently, higher-resolution Chandra observations of the
(post-Mice) Antennae show a great deal of hot diffuse gas,
but it has become all pervasive, extending further than the
stellar bodies of the galaxies (Fabbiano et al. 2002). It is
believed we have now seen, in the Mice, the onset of intense
star formation, starburst activity and starburst-driven dif-
fuse gaseous outflows in a classic full-blown disk-disk merger.
Though one cannot probe too far, due to lack of counts,
there even appears to be some temperature structure in the
NGC4676A wind. The eastern wind appears to be brighter
and spectrally harder (greener, rather than red; see Fig. 2b)
to the north than to the south. To the west of NGC4676A,
the situation is reversed, and all this may well be an indi-
cation that the rapid evolution and slewing of the galaxies
is already beginning to effect even these very nascent star-
burst winds. The existence of a classic starburst wind in a
disk-disk merger may be a very short-lived affair.
One can infer mean physical properties of the hot gas
around the northern and southern galaxies once some as-
sumptions have been made regarding the geometry of the
diffuse emission. The gas around each galaxy is assumed to
be contained in a spherical bubble of radius r, taken to be
the average radius of the lowest contour level seen in Fig.2a,
i.e. ≈7.5′′ (for A) and ≈6.0′′ (for B). Using these volumes,
10
A.M. Read
the fitted emission measure ηn2
eV (where η is the 'filling fac-
tor' - the fraction of the total volume V which is occupied
by the emitting gas) can be used to infer the mean elec-
tron density ne, and hence, assuming a plasma composition
of hydrogen ions and electrons, the total mass Mgas and
thermal energy of the gas Eth. Approximate values of the
cooling time tcool of the hot gas, and also the mass cooling
rate Mcool and adiabatic expansion timescale texp, can also
be calculated. The resulting gas parameters for A and B are
listed in Table 3.
Comparing the diffuse gas parameters in Table 3 with
those for isolated normal and starburst galaxies (Read, Pon-
man & Strickland 1997), and for merging galaxies (RP98),
one can say that the diffuse outflows in NGC4676A and
NGC4676B appear to be small, and probably young. Ex-
tents of 9−14 kpc are seen for the classic winds of M82 and
NGC253, and the outflowing, turbulent ISM of the Anten-
nae is seen out to ≈8 kpc (Fabbiano et al. 2002). Perhaps
5−10 times as much gas mass is contained within the M82
or NGC253 diffuse features than in either of the Mice, and
the diffuse gas mass of the Antennae may be up to 20 times
that of the Mice. The outflows in the Mice appear, in terms
of their extent and mass, more like those of small wind/large
corona systems, such as NGC891 and NGC4631.
In terms of temperature, it is believed that the hot ISM
of starburst galaxies has a a multi-temperature structure,
and as such, given the number of diffuse counts from the
Mice, little conclusive can be said. However, the single tem-
peratures obtained from the spectral fitting of the diffuse
spectra (≈0.50 keV) is wholly consistent with the range ob-
tained for other starburst and merging galaxies.
3.3 X-ray emission from the Mice
It has been seen that the X-ray emission from the Mice is
made up of point source and diffuse emission. In calculating
a total LX for the Mice, source W has been omitted, and
the remaining 5 sources and the two diffuse emission features
have been summed together. Total intrinsic (and emitted)
0.3−10 keV X-ray luminosities (in units of 1040 erg s−1) are
as follows: Sources − 5.56 (5.28), Diffuse emission − 1.88
(1.75), Total − 7.43 (7.03), leading to a fraction of the total
emission that is diffuse of 25%. As such it is rather brighter,
both fractionally and in absolute terms, than most normal
and starburst galaxies.
It is very instructive to compare these values with
those of the Antennae (Fabbiano et al. 2001), where ap-
proximately half of the total (0.1−10 keV) X-ray luminosity
(2.3×1041 erg s−1) is attributable to point sources, and half
is due to the extended diffuse thermal component. Given
the uncertainties in the assumed distances to both galaxies,
and noting the slight difference in galaxy masses, it appears
that, whereas the X-ray point source populations in the two
systems may be rather similar (at least at the high-LX end),
the extended diffuse emission component is far fainter and
less extended in the Mice than it is in the more evolved
Antennae. This indicates that, if both systems are at dif-
ferent evolutionary stages of an encounter involving similar
gaseous disks, then at the Mice stage, the ULX populations
have evolved more rapidly than the diffuse gas structures,
whereas at the Antennae stage, the diffuse gas structures
have caught up. This is not surprising given that the ULXs
are likely due to the evolved components of single massive
stars (e.g. Roberts et al. 2002), and thus can be created over
short (106 yr) timescales, whereas the diffuse components are
mainly due to galactic winds, galactic fountains and chim-
neys, and the hot phases of the ISM, requiring the agglomer-
ation of several, perhaps thousands, of supernova remnants,
thus requiring a far longer (few ×107 yr) timescale.
The Mice and the Antennae appear to be at differ-
ent stages of very similar evolutionary tracks (involving two
similarly-sized gas-rich spiral disks), the only difference be-
tween the systems being (apart from their merger epoch)
that the Antennae appears slightly more massive. It is there-
fore attractive to predict that that the Mice will evolve into
a system very like the Antennae within the next 108 years or
so. Though the high-LX end of the point source populations
of the two systems are already very similar, the diffuse gas
component of the Mice's emission requires a few ×107 years
worth of starburst-injected mass and energy, before it can
compare with the diffuse emission seen in the Antennae. At
present, the outflows in the Mice, in terms of their size and
mass, are not yet especially significant, resembling struc-
tures observed in nearby systems with small winds and/or
large coronae.
4 CONCLUSIONS
Presented here are high spatial and spectral resolution
Chandra ACIS-S X-ray observations of the famous inter-
acting galaxy pair, the Mice, a system very similar to the
Antennae galaxies (albeit at a much larger distance), in con-
sisting of the merger of two equally-sized spiral galaxies. The
Mice are at an earlier interaction stage however, the galaxy
disks not yet having begun to merge. Previously unpublished
ROSAT HRI data of the system are also presented. The pri-
mary results can be summarized as follows:
• Of great interest is the discovery of what appears to be
starburst-driven galactic winds outflowing along the minor
axes of both galaxies (particularly the northern). Spectrally
soft X-ray emission extends beyond the northern edge-on
galactic disk, into regions occupied by plumes of Hα emis-
sion. Fitting of the diffuse emission spectra indicates tem-
peratures wholly consistent with well-known nearby galactic
winds. That such classic winds could exist in the Mice is per-
haps surprising, as one would expect the rapid evolution and
violence of the environment to distort the wind structures
very quickly. The phenomenon may not last long however
− in terms of extent and mass, the winds are small, low-
luminosity, perhaps newly-formed and already show complex
spatial and spectral structure, indicating that they could be
being disrupted by the encounter. These low-temperature,
diffuse X-ray features, extending out of the galactic disks
appear to be the very beginnings of starburst-driven hot
gaseous outflows in a full-blown disk-disk merger.
• In addition, five bright (LX (0.3−10 keV) >
∼5×1039 erg
s−1) point sources are formally detected, associated with the
Mice system. The sources detected at the nuclei of the two
galaxies are very bright, though perhaps not single, indi-
vidual sources, and appear quite similar. Emission from the
northern nucleus is likely absorbed by the intervening edge-
on disk of the northern galaxy. Both nuclear X-ray sources,
given the spatial structure of the nuclear and surrounding
Chandra Observations of the Mice
11
Diff.
Src.
kT
r
(keV)
(kpc)
A
B
0.50
0.46
3.2
2.6
ne
(cm−3)
(×1/√η)
0.019
0.025
Mgas
(M⊙)
(×√η)
6.5×107
4.3×107
Eth
(erg)
(×√η)
1.8×1056
1.2×1056
tcool
(Myr)
(×√η)
480
380
Mcool
(M⊙ yr−1)
texp
(Myr)
0.13
0.11
8.8
7.3
Table 3. Values of physical parameters for the diffuse gas associated with NGC4676A and NGC4676B. The values quoted are for a
bubble model (see text), and η is the filling factor of the gas.
Heckman T.M., 1993, in Schlegel E.M., Petre R., eds, AIP Conf.
Proc. 313, The Soft X-ray Cosmos. AIP Press, Woodbury,
NY, p.139
Hibbard J.E., van Gorkom J.H., 1996, AJ, 111, 655
Mihos J.C., Bothun G.D., Richstone D.O., 1993, ApJ, 418, 82
Read A.M., Pietsch W., 2001, A&A, 373, 473
Read A.M., Ponman T.J., D.K., Strickland D.K., 1997, MNRAS,
286, 626
Read A.M., Ponman T.J., 1998, MNRAS, 297, 143 (RP98)
Read A.M., Ponman T.J., 2001, MNRAS, 328, 127
Roberts T.P., Warwick R.S., Ward M.J., Murray S.S., 2002, MN-
RAS, 337, 677
Schombert J.M., Wallin J.F., Struck-Marcell C., 1990, AJ, 99,
497
Schweizer F., 1989, Nat, 338, 119
Smith B.J., Higdon J.L., 1994, AJ, 108, 837
Sotnikova N.Y., Reshetnikov V.P., 1998, Astronomy Letters, 24,
73
Stockton A., 1974, AJ, 187, 219
Strickland D.K., Heckman T.M., Weaver K.A., Hoopes C.G.,
Dahlem M., 2002, ApJ, 568, 689
Toomre A., 1977, in Evolution of Galaxies and Stellar Popula-
tions, Ed. B.M.Tinsley, R.B.Larson, p.401, (New Haven: Yale
University Observatory)
Toomre A., Toomre J., 1972, ApJ, 178, 623
Tschoke D., Hensler G., Junkes N., 2000, A&A, 360, 447
Vorontsov-Vel'yaminov B.A., 1957, Astron. Tsirk., 178, 19
Young J.S., Scoville N.Z., 1991, ARA&A, 29, 581
Yun M.S., Hibbard J.E., 2001, ApJ, 550, 104
Zezas A., Fabbiano G., 2002, 2002, ApJ, 577, 726
Zezas A., Fabbiano G., Rots A.H., Murray S.S., 2002, ApJ, 577,
710
diffuse X-ray emission, and the large amounts of associated
molecular gas and Hα emission, are likely starburst regions.
A source is detected coincident with a molecular gas com-
plex at the tip of one edge of the southern galaxy's bar. A
smaller X-ray enhancement (together with molecular emis-
sion) is observed at the opposite end. Two further spectrally
soft sources are detected within the two tidal tails.
• The source X-ray luminosity function appears in sharp
contrast to that of other normal and starburst galaxies. It
is similar however to that of the more evolved Antennae,
though the number of high-LX sources in the Mice is prob-
ably less.
• Whilst the point source XLFs of the Mice and Anten-
nae appear similar, far less diffuse emission is detected in
the Mice. This indicates that the high-luminosity source
populations in these systems evolve more rapidly than the
diffuse gas structures. For the Mice to evolve, within the
next 108 years, into a system like the Antennae, much of the
starburst-injected energy and mass needs to feed the diffuse
gas component.
• There is no evidence for any variability between the
Chandra data and the ROSAT HRI data.
ACKNOWLEDGEMENTS
AMR acknowledges the support of PPARC funding, and
thanks the referee for useful comments which have improved
the paper. AMR also thanks John Hibbard for making his
Hα data available, and Trevor Ponman for carefully reading
the manuscript. Optical images are based on photographic
data obtained with the UK Schmidt Telescope, operated by
the Royal Observatory Edinburgh, and funded by the UK
Science and Engineering Research Council, until June 1988,
and thereafter by the Anglo-Australian Observatory. Origi-
nal plate material is copyright (c) the Royal Observatory Ed-
inburgh and the Anglo-Australian Observatory. The plates
were processed into the present compressed digital form with
their permission. The Digitized Sky Survey was produced at
the Space Telescope Science Institute under US Government
grant NAG W-2166.
REFERENCES
Barnes J.E., Hernquist L., 1991, ApJ, 370, L65
Barnes J.E., Hernquist L., 1996, ApJ, 471, 115
Bravo-Guerrero J., Read A.M., Stevens I.R., submitted to MN-
RAS
Engelmaier P., Gerhard O., 1997, MNRAS, 287, 57
Fabbiano, G., Zezas, A. Murray, S.S., 2001, ApJ, 554, 1035
Giacconi R., et al., 2001, ApJ, 551, 624
|
astro-ph/0608645 | 2 | 0608 | 2006-12-15T19:00:19 | 9.7 um Silicate Features in AGNs: New Insights into Unification Models | [
"astro-ph"
] | We describe observations of 9.7 um silicate features in 97 AGNs, exhibiting a wide range of AGN types and of X-ray extinction toward the central nuclei. We find that the strength of the silicate feature correlates with the HI column density estimated from fitting the X-ray data, such that low HI columns correspond to silicate emission while high columns correspond to silicate absorption. The behavior is generally consistent with unification models where the large diversity in AGN properties is caused by viewing-angle-dependent obscuration of the nucleus. Radio-loud AGNs and radio-quiet quasars follow roughly the correlation between HI columns and the strength of the silicate feature defined by Seyfert galaxies. The agreement among AGN types suggests a high-level unification with similar characteristics for the structure of the obscuring material. We demonstrate the implications for unification models qualitatively with a conceptual disk model. The model includes an inner accretion disk (< 0.1 pc in radius), a middle disk (0.1-10 pc in radius) with a dense diffuse component and with embedded denser clouds, and an outer clumpy disk (10-300 pc in radius). | astro-ph | astro-ph |
9.7 µm Silicate Features in AGNs: New Insights into Unification
Models
Y. Shi1, G. H. Rieke1, D. C. Hines2, V. Gorjian3, M. W. Werner3, K. Cleary3, F. J. Low1,
P. S. Smith1, J. Bouwman4
ABSTRACT
We describe observations of 9.7 µm silicate features in 97 AGNs, exhibiting a
wide range of AGN types and of X-ray extinction toward the central nuclei. We
find that the strength of the silicate feature correlates with the HI column density
estimated from fitting the X-ray data, such that low HI columns correspond to
silicate emission while high columns correspond to silicate absorption. The be-
havior is generally consistent with unification models where the large diversity in
AGN properties is caused by viewing-angle-dependent obscuration of the nucleus.
Radio-loud AGNs and radio-quiet quasars follow roughly the correlation between
HI columns and the strength of the silicate feature defined by Seyfert galaxies.
The agreement among AGN types suggests a high-level unification with similar
characteristics for the structure of the obscuring material. We demonstrate the
implications for unification models qualitatively with a conceptual disk model.
The model includes an inner accretion disk (< 0.1 pc in radius), a middle disk
(0.1-10 pc in radius) with a dense diffuse component and with embedded denser
clouds, and an outer clumpy disk (10-300 pc in radius).
Subject headings: galaxies: active -- galaxies: nuclei
1.
Introduction
The arrangement of material around the central supermassive blackhole (SMBH) in
active galactic nuclei (AGNs) is crucial in unification schemes, where the active nuclei clas-
1Steward Observatory, University of Arizona, 933 N Cherry Ave, Tucson, AZ 85721, USA
2Space Science Institue 4750 Walnut Street, Suite 205, Boulder, Colorado 80301
3Jet Propulsion Laboratory, MC 169-327, California Institute of Technology, 4800 Oak Grove Drive,
Pasadena, CA 91109
4Max-Planck-Institut fu r Astronomie, D-69117 Heidelberg, Germany
-- 2 --
sified as type 2 arise from the obscuration of broad optical emission lines in type 1 ob-
jects (Antonucci 1993). The structure of the circumnuclear material may be critical to
the growth of the SMBH and may influence the feedback from the AGN and its effects
on the formation and evolution of the host galaxies. The infrared (IR) spectral energy
distribution (SED) arises from the heating of surrounding material, and its shape should
indicate how the material is organized. There are at least three different possibilities: com-
pact (Pier & Krolik 1992, 1993), extended (Granato & Danese 1994; Granato et al. 1997)
and cloudy (Krolik & Begelman 1986; Rowan-Robinson 1995; Nenkova et al. 2002) struc-
ture models all fit the data well (assuming an extra component of star formation in the
compact model). The silicate emission and absorption features now being observed with
Spitzer (e.g., Weedman et al. 2005) can be used to probe these possibilities further.
In this paper, we present Spitzer Infrared Spectrograph (IRS; Houck et al. 2004) ob-
servations of 9.7 µm silicate features in 93 AGNs, including examples from the literature
(Siebenmorgen et al. 2005; Hao et al. 2005; Sturm et al. 2005; Ogle et al. 2006). We com-
plement them with groundbased measurements for 4 more (Roche et al. 1991). By combining
these observations with X-ray data that explore the line of sight toward the central engine,
we can gain new insights into the unification scheme.
2. Sample and Data Reduction
Our sample (see on-line Table 1) includes a variety of types of AGN: radio-loud quasars,
FR II radio galaxies, Seyfert 1 and Seyfert 2 galaxies, broad absorption-line quasars (BALQs),
low-ionization nuclear emission-line region (LINER) galaxies, and non-BAL radio-quiet quasars
including optically selected Palomar-Green (PG) quasars and IR quasars as selected by the
Two-Micron All Sky Survey (2MASS) from Smith et al. (2002). Except for the objects from
the literature, quasars of all types and FR II radio galaxies are the most luminous objects
in their parent samples (which include upper limits in redshift in their definitions) while
Seyfert 1 and Seyfert 2 galaxies are derived to have high brightness and a broad range in HI
column density from Turner & Pounds (1989) and Risaliti et al. (1999), respectively. The
whole sample spans a range of HI column density from 1020 cm−2 up to 1025 cm−2. We focus
on the sample of 85 objects with available HI column densities in this study, although all
97 objects have available silicate data. X-ray spectra were retrieved from the Chandra data
archive for 8 sources. The data reduction and the measurement of column densities through
power-law fits are as described in Shi et al. (2005). The column densities of most of the
2MASS QSOs are based on the X-ray hardness ratio from Wilkes et al. (2002) assuming an
intrinsic power law X-ray spectrum with a photon index of 1.7. The associated uncertainty
-- 3 --
is estimated as a factor of 3 corresponding to a change of ∼1 in the photon index. The
column density for the remaining sources is obtained from the literature as shown in the
on-line Table 1.
The IRS spectra presented for the first time in this paper, except for those of the Seyfert
galaxies, were obtained using the standard staring mode. The intermediate products of the
Spitzer Science Center (SSC) pipeline S11.0.2 were processed within the SMART software
package (Higdon et al. 2004). The background was subtracted using associated spectra from
the two nodded, off-source positions. This also subtracts stray light contamination from the
peak-up apertures, and adjusts pixels with anomalous dark current. Pixels flagged by the
SSC pipeline as "bad" were replaced with a value interpolated from an 8-pixel perimeter sur-
rounding the suspect pixel. The spectra were extracted using a 6.0 pixel fixed-width aperture
for the short low module (SL), and 5.0 pixels for the long low module (LL). The spectra were
calibrated using a spectral response function derived from IRS spectra and Kurucz stellar
models for a set of 16 Sun-like stars that exhibit: 1) high signal-to-noise observations, 2) no
residual instrumental artifacts, and 3) no signs of IR excess. The absolute flux density scale
is tied to calibrator stars observed by the IRS instrument team and referenced to calibrated
stellar models provided by the SSC (see also Bouwman et al. 2006). The uncertainties in
the final calibration are dominated by random noise. The relative flux calibration across the
spectrum is accurate to ∼ 1 − 2% (Hines et al. 2006; Bouwman et al. 2006).
The data reduction was slightly different for the Seyfert galaxies first presented in this
paper. The whole Seyfert sample consists of a mixture of point-like and extended sources.
They are at relatively low redshift (z ∼0.01) and are lower luminosity AGNs, so care was
taken to minimize the contribution of the host galaxies. For the extended sources, a fixed-
width column extraction that is narrow enough to exclude most of the extended source
would be most appropriate. However, narrow fixed-width extraction windows introduced
artifacts into the extracted spectra. Instead, a narrow expanding-aperture extraction was
performed on all sources, point-like or extended. This was deemed to be an acceptable
compromise between rejecting the extended source component and introducing artifacts into
the spectrum. The spectra were extracted using an expanding extraction aperture defined
to be 4 pixels wide at the wavelengths of 6 µm and 12 µm for the second and the first
order of SL, respectively, and of 16 µm and 27 µm for the second and the first order of LL,
respectively. At the wavelength of 10 µm, the extraction aperture is around 3.3 pixels. A
detailed description of the data reduction is given by Gorjian et al.
(2006).
The silicate feature strengths were estimated as follows. Narrow emission lines were
removed from the spectra and they were then smoothed to a resolution of 0.1 µm. The
continuum was defined by using a spline interpolation between the blue and red ends of the
-- 4 --
IRS spectral range. We defined (all in rest-frame wavelengths) the blue end as 5-7.5 µm,
while the red end was defined as 13-14 µm for the 13 objects observed only with the SL
module. For six BALQs at higher redshifts, the red end was chosen to be 10.5-13 µm. For
the remainder of the sample, the red spectral end was defined as 25-30 µm. For nine objects
with the red end contaminated by silicate emission (Figure 1(b)), the continuum at this end
was estimated to be below that observed by <20% at the reddest point, based on the spectra
of objects without contamination. For twelve objects with the blue end contaminated by
the strong 6-8 µm aromatic feature (Figure 1(c)), the continuum in this bandpass was given
by two segments between 5 and 5.5 µm where the contamination is negligible. The fit was
judged to be good when it matched the flux at both ends within the noise and the curvature
of the continuum varied gradually from one end to the other over the silicate region. The
uncertainty in continuum fitting was obtained by adjusting the flux of the continuum up
and down over the fitting spectral range. For the objects with continuum not contaminated
by either aromatic or silicate features as shown in Figure 1(a), we adjusted the fit until the
continuum was just above (for upper error) or below (for lower error) the observed flux over
the whole spectral fitting range. To estimate the uncertainty for the objects with red end
contaminated by the silicate feature (Figure 1(b)), we first fixed the blue end and adjusted
the red end continuum up to the red end point that matched the upper-envelope of the
underlying continuum plus noise and down by the same amount as an estimate of the errors
of the continuum fit. To account for the effects of noise in the blue end, we added another 4%
uncertainty (the mean value for objects without aromatic or silicate feature contamination).
For the objects with blue end contaminated by the aromatic feature (Figure 1(c)), we fixed
the red end and adjusted the blue end to get the uncertainty in a similar way where the
upper-limit of the underlying continuum is the flux at 7 µm (the median wavelength between
the peaks of the 6.2 µm and 7.7 µm aromatic features). Another 4% uncertainty was added
to this uncertainty to account for the effects of noise in the red end. There are no objects
with red end contaminated by the silicate feature and blue end contaminated by the aromatic
features.
The strength of the 9.7 µm silicate feature is defined as (Ff −Fc)/Fc, where Ff and Fc are
the observed flux density and underlying continuum flux density, respectively, at the peak
(for emission) or the minimum (for absorption) of the silicate feature. The feature strength
in this definition is a direct measure of the optical depth for the silicate absorption. Figure 2
shows the IRS spectra presented for the first time in this study in order of HI column density.
No ice or hydrocarbon absorption is found in Figure 2. Especially at 5-8 µm, where there is
no contamination of the silicate feature, the spectra are well described by power laws except
for those with aromatic emission. Therefore, once the continuum had been fitted, we could
calculate the feature strengths unambiguously.
-- 5 --
3. Results
Figure 3 shows the strength of the 9.7 µm silicate feature as a function of HI column den-
sity. The right y-axis of Figure 3 shows the IR-absorbing column density estimated from the
silicate absorption feature by assuming τ9.7µm = ln(Fc/Ff), Av/A9.7µm=19 (Roche & Aitken
1985) and Av/NH=0.62×10−21cm−2 (Savage & Mathis 1979). The silicate feature varies
from emission (+) to absorption (-) as the X-ray spectra become more heavily obscured.
The trend is also demonstrated by the composite spectra in different bins of HI column as
shown in Figure 4. Since the silicate feature is broad, the composite spectra are the geomet-
ric mean spectra to conserve the global spectral shape (e.g. Vanden Berk et al. 2001). The
relationship between HI column and silicate feature behavior is generally consistent with the
AGN unification scheme where material surrounding the central SMBH obscures both the
X-ray and silicate emissions.
The correlation (black lines in Figure 3) is defined by all types of AGN. To investigate
the behavior for individual AGN types, we unified Seyfert 1 and Seyfert 2 galaxies as one type
of AGN (Seyfert galaxies). Similarly, FR II radio galaxies and radio-loud QSOs are classified
as radio-loud AGNs while PG and 2MASS QSOs are classified as radio-quiet QSOs. Table 2
shows the linear fits to these AGN types where the limits to the HI column density are treated
as detections during the fitting. The fits for different AGN types are characterised by large
uncertainty and are generally consistent with each other within the uncertainty. Radio-quiet
QSOs (2MASS and PG QSOs) exhibit a small (less than two standard deviations) deviation
from other two types of AGN. As shown in Figure 3, this deviation arises because the PG
QSOs with low HI columns have larger silicate strengths.
To quantify the discrepancy between different types of AGN, we examined the proba-
bility that the other AGN types follow the Seyfert correlation by producing a Monte Carlo
theoretical distribution with 10000 data points using the Seyfert correlation (red line in Fig-
ure 3) with associated scatter. Table 3 shows the result of three tests: 1.) the probability
from a K-S test that the given AGN type has the same distribution as the theoretical distri-
bution; 2.) the significance that two distributions have the same mean (indicated by a value
greater than 0.05); and 3.) the significance that two distributions have the same variance
(indicated by a value greater than 0.05). The high K-S probability for the Seyfert galaxies
indicates that the theoretical distribution is a good representative of them (as it should be).
As shown in Table 3 and Figure 3, radio-loud AGNs (radio-loud QSOs and FR II galaxies)
are consistent with the Seyfert correlation and radio-quiet QSOs (2MASS and PG QSOs)
show slightly higher silicate strengths at low HI column compared to Seyfert galaxies.
Sturm et al. (2006) recently found that six type 2 QSOs in the HI column range of 1021.5-
1024 cm−2 do not show any silicate feature, slightly higher but still consistent with those of
-- 6 --
the Seyfert galaxies at the same range of HI columns. As shown in Figure 3, the three BALQs
and the one LINER in this study deviate from the Seyfert correlation significantly. More
objects are needed to address whether these two types follow the Seyfert correlation. Given
that most PG QSOs have upperlimit measurements of HI columns, the intrinsic deviation of
radio-quiet QSOs from the Seyfert trend should be smaller than we calculate. Although the
intrinsic slope of the Seyfert correlation should be smaller due to the lowerlimit measurement
of the HI column for half the Seyfert 2 galaxies, the effect on the comparision is small because
it is dominated by the column range where Seyfert galaxies have detected HI columns. In
Table 2, we also list the linear fits to the Seyferts and the whole sample excluding objects
with limit measurements of the HI columns. The comparision between the new slopes shows
that the consistency between different AGN types may become even better.
Therefore, except for LINERs and BALQs, the remaining non-Seyfert AGN types follow
more or less the Seyfert correlation. The rough agreement among AGN types indicates
that the geometry of the circumnuclear material and our viewing angle are the primary
factors influencing the relation between the silicate feature and the X-ray attenuation. This
correlation is nonetheless curious because, as shown in Figure 3, the column required for the
observed X-ray obscuration levels is up to ∼1025 cm−2, two orders of magnitude larger than
that required to produce the silicate absorption, ∼ 1023 cm−2. This implies that along many
lines of sight there must be X-ray absorbing material that is not contributing to the silicate
absorption.
The dispersion in the correlation is much larger than the uncertainties in silicate-feature
strength and HI column density. The large dispersion appears to be characteristic of all the
AGN types with large numbers of objects (Seyfert galaxies, radio-quiet QSOs and radio-
loud AGNs). This is consistent with a common mechanism (the circumnuclear geometry)
regulating the correlation and dispersion in Figure 3. Due to different physical aperture
diameters from 500 pc to galaxy-scale in size, the differing amounts of extended emission
from star-forming regions may contribute to the scatter. However, such contributions must
be small. As shown in Figure 2, twelve of 85 objects have strong aromatic features and their
silicate features may be contaminated by star-forming regions. However, the dispersion of the
correlation is almost the same for the 73 objects without aromatic features. Another possible
contribution to the scatter is the gas-to-dust ratio. However, as indicated by Figure 3, the
gas-to-dust ratio needs to vary nearly three orders of magnitude to account for the scatter
in the correlation. Such large variation is unreasonable given the similar IR SEDs and IR
luminosities for objects with similar HI column but different silicate features as shown in
Figure 2, for example, NGC 4941 versus NGC 3281.
Another characteristic of the correlation is that several Compton-thick AGNs have sil-
-- 7 --
icate absorptions that are much weaker than predicted. Again, this implies that additional
absorbing material must be present in the Compton-thick sources, but placed so it does not
obscure the IR emission.
4. Discussion
We now describe a conceptual model to explain the large difference in the absorbing
columns for the X-rays and infrared, but that can also account for the correlation in Figure 3.
The goal is to find a geometry for the circumnuclear material that can explain Figure 3 solely
in terms of variations in viewing angle. As mentioned in the Introduction, there are three
classes of model that are generally successful in fitting the IR SEDs of specific AGN types:
1.) a compact disk (e.g. Pier & Krolik 1992); 2.) a more extended disk (radius of hundreds
of pc) (e.g. Granato et al. 1997); and 3.) a disk with clumps or clouds of material (e.g.
Nenkova et al. 2002). We now explore whether the ideas behind these three types of model
can be combined to provide a possible explanation for the behavior of the silicate feature. It
is beyond the scope of this paper to compute a quantitative model to fit the data. However,
by confining our argument to combinations of features in models already shown to fit other
aspects of AGN behavior, it is likely that our explanation of the silicate-X-ray correlation
will also be compatible with the other observations of AGNs. Overall, our model requires
both a component of material similar to the cloudy model (e.g. Nenkova et al. 2002) and a
diffuse component with outer radius similar to the compact disk-model (e.g. Pier & Krolik
1992) and with column density similar to the model of the extended disk (e.g. Granato et al.
1997).
In Figure 5, we show the hypothetical disk geometry. We include an inner accretion disk
(AD), and in the same plane a middle disk (MD) with a diffuse component (grey) joining
to the AD. Denser clouds or clumps are embedded in this MD. It merges with a cloudy or
clumpy outer disk (OD). The X-ray and UV radiation from the AD heat the dust in the MD
and OD to produce IR emission and ionize the broad-emission-line clouds and the narrow-
emission-line (NEL in Figure 5) clouds, which are not shielded by the diffuse component.
Compton-thick X-ray obscuration can arise either in the AD, or in the clouds in the MD.
Irregularities in the AD or the passage of clouds through the line of sight may be responsible
for the variations in X-ray obscuration observed toward some AGNs (e.g. Elvis et al. 2004;
Risaliti et al. 2005). However, for a model with only the central disk and the clouds, there
is a large probability of viewing the central X-ray emission directly without any extinction
even for an inclined disk (along the equatorial plane). This is inconsistent with Figure 3,
where the lower-envelope of the data distribution shows a strong trend that the HI column
-- 8 --
density increases with the depth of the silicate absorption. The diffuse component of the
MD is required to obscure the X-ray emission even if the line of sight does not cross any
cloud. This component is also the source of the silicate emission. The silicate absorption
results from the material in the OD.
We argue that the outer edge of the MD should not extend beyond around 10 pc and
the dust within it can have sufficiently high temperature (> 300 K; Laor & Draine 1993)
to produce the silicate emission. Given the maximum silicate absorption of 1023cm−2 and
the maximum HI column of >1025cm−2, the clouds in the MD should have NH > 1023 cm−2
and produce significant obscuration of the X-ray emission for an intercepting line of sight
but not obscure the silicate emission significantly due to a small covering factor. For a given
strength of the silicate feature along a line of sight, the large variation in the HI column
is caused by the variation in the number or column density of clouds that the line of sight
intercepts and the minimum X-ray obscuration is due only to the diffuse component. Based
on this concept, as the line of sight varies such that the silicate strength ranges from 0.0
to -0.8 (the minimum value in the sample) as shown in Figure 3, the column density of
the diffuse component varies from 1021 cm−2 to 1023 cm−2. The evidence for the cloudy
OD is that the mid-IR images of NGC 1068 at 10 µm (Jaffe et al. 2004) and NGC 4151
(Radomski et al. 2003) at ∼10 and ∼18 µm show that the radius of any diffuse component
should be smaller than 2 pc and 35 pc, respectively. However, the dust in the OD is relatively
cool and emits any reprocessed energy mainly at far-IR wavelengths and thus may be missed
for observations at mid-IR wavelengths. Since both the MD and OD are unresolved in the
IRS beam, the observed silicate absorption can be provided by several clouds in the OD and
the depth of the absorption feature is determined by the average number of clouds along a
line of sight. This is because a cloud in the OD is not large enough to cover the whole MD.
5. Conclusions
We report observations of 9.7µm silicate features in 97 AGNs. The features vary from
emission to absorption with increasing HI column density, consistent with unification mod-
els. Radio-loud AGN (radio-loud QSOs and FR II galaxies) and radio-quiet QSOs (PG and
2MASS QSOs) lie roughly on the Seyfert-correlation between HI column and silicate feature
strength. The behaviors of LINERs and BALQs are not clear due to the small number of
objects for these two AGN types. The scatter in the relation is large and several Compton-
thick AGNs do not show deep silicate absorption. Qualitatively, the correlation requires a
circumnuclear disk geometry with an accretion disk outside of which is a middle disk with
high density and with even denser clouds embedded (0.1-10 pc in radius), co-aligned with
-- 9 --
an outer clumpy disk (10-300 pc in radius). The similarity of the behavior of various types
of AGN suggests that this disk geometry may be typical for AGN in general.
We thank the anonymous referee for detailed comments. We also thank Roberto
Maiolino for helpful suggestions. Support for this work was provided by NASA through
contract 1255094 issued by JPL/ California Institute of Technology.
Antonucci, R. 1993, ARA&A, 31, 473
REFERENCES
Belsole, E., Worrall, D. M., & Hardcastle, M. J. 2006, MNRAS, 366, 339
Bouwman, J., Henning, Th., Hillenbrand L., Silverstone, M., Meyer, M., Carpenter, J.,
Pascuci, I., Wolf, S., Hines, D. 2006, submitted.
Brunner, H., Mueller, C., Friedrich, P., Doerrer, T., Staubert, R., & Riffert, H. 1997, A&A,
326, 885
Dewangan, G. C., Griffiths, R. E., & Schurch, N. J. 2003, ApJ, 592, 52
Donato, D., Gliozzi, M., Sambruna, R. M., & Pesce, J. E. 2003, A&A, 407, 503
Elvis, M., Risaliti, G., Nicastro, F., Miller, J. M., Fiore, F., & Puccetti, S. 2004, ApJ, 615,
L25
Gallagher, S. C., Brandt, W. N., Sambruna, R. M., Mathur, S., & Yamasaki, N. 1999, ApJ,
519, 549
Gallagher, S. C., Brandt, W. N., Chartas, G., & Garmire, G. P. 2002, ApJ, 567, 37
Gorjian, V., Cleary, K., Marshall, J., DeMuth, N., Werner, M.W., Lawrence, C. 2006, ApJ,
in Prep.
Granato, G. L., & Danese, L. 1994, MNRAS, 268, 235
Granato, G. L., Danese, L., & Franceschini, A. 1997, ApJ, 486, 147
Hao, L., et al. 2005, ApJ, 625, L75
Higdon, S. J. U., et al. 2004, PASP, 116, 975
-- 10 --
Hines, D. C., et al. 2006, ApJ, in press.
Houck, J. R., et al. 2004, ApJS, 154, 18
Isobe, N., Tashiro, M., Makishima, K., Iyomoto, N., Suzuki, M., Murakami, M. M., Mori,
M., & Abe, K. 2002, ApJ, 580, L111
Jaffe, W., et al. 2004, Nature, 429, 47
Krolik, J. H., & Begelman, M. C. 1986, ApJ, 308, L55
Laor, A., & Draine, B. T. 1993, ApJ, 402, 441
Leighly, K. M., O'Brien, P. T., Edelson, R., George, I. M., Malkan, M. A., Matsuoka, M.,
Mushotzky, R. F., & Peterson, B. M. 1997, ApJ, 483, 767
Mathur, S., et al. 2000, ApJ, 533, L79
Nenkova, M., Ivezi´c, Z., & Elitzur, M. 2002, ApJ, 570, L9
Ogle, P. M., Whysong, D., Antonucci, R. 2006, astro-ph/0601485
Pier, E. A., & Krolik, J. H. 1992, ApJ, 401, 99
Pier, E. A., & Krolik, J. H. 1993, ApJ, 418, 673
Porquet, D., Reeves, J. N., O'Brien, P., & Brinkmann, W. 2004, A&A, 422, 85
Ptak, A., Terashima, Y., Ho, L. C., & Quataert, E. 2004, ApJ, 606, 173
Radomski, J. T., Pina, R. K., Packham, C., Telesco, C. M., De Buizer, J. M., Fisher, R. S.,
& Robinson, A. 2003, ApJ, 587, 117
Reeves, J. N., & Turner, M. J. L. 2000, MNRAS, 316, 234
Risaliti, G., Maiolino, R., & Salvati, M. 1999, ApJ, 522, 157
Risaliti, G., Elvis, M., Fabbiano, G., Baldi, A., & Zezas, A. 2005, ApJ, 623, L93
Roche, P. F., & Aitken, D. K. 1985, MNRAS, 215, 425
Roche, P. F., Aitken, D. K., Smith, C. H., & Ward, M. J. 1991, MNRAS, 248, 606
Rowan-Robinson, M. 1995, MNRAS, 272, 737
Savage, B. D., & Mathis, J. S. 1979, ARA&A, 17, 73
-- 11 --
Shi, Y., et al. 2005, ApJ, 629, 88
Siebenmorgen, R., Haas, M., Krugel, E., & Schulz, B. 2005, A&A, 436, L5
Smith, P. S., Schmidt, G. D., Hines, D. C., Cutri, R. M., & Nelson, B. O. 2002, ApJ, 569,
23
Sturm, E., et al. 2005, ApJ, 629, L21
Sturm, E., et al. 2006, astro-ph/0601204
Turner, T. J., & Pounds, K. A. 1989, MNRAS, 240, 833
Vanden Berk, D. E., et al. 2001, AJ, 122, 549
Wang, T., Brinkmann, W., & Bergeron, J. 1996, A&A, 309, 81
Worrall, D. M., Birkinshaw, M., Hardcastle, M. J., & Lawrence, C. R. 2001, MNRAS, 326,
1127
Wilkes, B. J., Schmidt, G. D., Cutri, R. M., Ghosh, H., Hines, D. C., Nelson, B., & Smith,
P. S. 2002, ApJ, 564, L65
Wilkes, B. J., Pounds, K. A., Schmidt, G. D., Smith, P. S., Cutri, R. M., Ghosh, H., Nelson,
B., & Hines, D. C. 2005, ApJ, 634, 183
This preprint was prepared with the AAS LATEX macros v5.2.
-- 12 --
Table 1. Source Characteristics
Sources
(1)
PG0050+124
PG0052+251
PG0804+761
PG0953+414
z
(2)
0.058
0.155
0.100
0.234
Type
Strength
Reference
(3)
(4)
(5)
N X
H
(6)
Reference
(7)
PG
PG
PG
PG
0.38
0.33+0.06
−0.06
0.60
0.40+0.08
−0.08
2
0
2
0
0.03+0.01
−0.016
0.04 +0.01
−0.01
0.03+0.00
−0.005
< 0.006
10
8
18
9
Note. -- Column (1): The sources. Column (2): Redshift. Column (3): The types
′PG′: PG quasar; ′RLQ′: radio-loud quasar; ′Sy1′: Seyfert 1 galaxies;
of AGNs.
′Sy2′: Seyfert 2 galaxies; ′BALQ′: broad absorption-line quasar; ′FRII′: FR II ra-
dio galaxies; ′LINER′:
low-ionization nuclear emission-line region; ′2MQ′: 2MASS
quasar. Column(4): The strength of the silicate feature as defined in § 2. Column (5):
The references for the silicate data: (0) This work; (1) Siebenmorgen et al. (2005);
(2) Hao et al. (2005); (3) Sturm et al. (2005); (4) Roche et al. (1991); (5) Ogle et al.
(2006). Column (6): The intrinsic HI column density in the unit of 1022cm−2. Col-
umn (7): References for the HI column densities: (6) This work; (7) Donato et al.
(2003); (8) Brunner et al. (1997); (9) Porquet et al. (2004); (10) Reeves & Turner
(2000); (11) Leighly et al. (1997); (12) Gallagher et al. (2002); (13) Mathur et al.
(2000); (14) Dewangan et al. (2003); (15) Turner & Pounds (1989); (16) Risaliti et al.
(1999); (17) Wilkes et al. (2002); (18) Wang et al. (1996); (19) Ptak et al. (2004); (20)
Gallagher et al. (1999); (21) Wilkes et al. (2005) ; (22) Worrall et al. (2001) ;(23)
Belsole et al. (2006); (24) Isobe et al. (2002)
The complete version of this table is in the electronic edition of the Journal.
-- 13 --
Table 2. The linear fits to objects with
different AGN types
Type
(1)
A
(2)
B
(3)
Number
(4)
2.6±0.7
Seyfert
2MQ & PG
5.5±1.2
FRII & RLQ 3.9±1.6
3.3±0.5
all objects
Seyfert1
4.6±1.3
all objects1
5.1±0.9
-0.12 ± 0.03
-0.25 ± 0.06
-0.17 ± 0.07
-0.15 ± 0.02
-0.21 ± 0.06
-0.23 ± 0.04
38
28
15
85
23
61
Note. -- Column(1): The AGN types. See caption to Figure
3. Column(2) and Column(3): Parameters of the linear fit S =
A + B ∗ log(N X
H ) where S is the strength of the silicate feature
as defined in § 2 and N X
H is the HI column density in the unit of
cm−2. Both parameters A and B are unitless due to the definition
of the silicate strength. Column(4): the total number of objects
for each fit.
1Fits excluding objects without measured HI columns (i.e., upper
and lower limits to the HI are excluded).
-- 14 --
Table 3. Tests of the correlation
Type
(1)
Sy
2MQ & PG
RLQ & FR II
Num
(2)
38
28
15
K-S Prob.(%) Mean
Variance
(3)
90
2
10
(4)
1.0
0.02
0.06
5
0.85
0.35
0.40
Note. -- Column(1): The AGN type. See caption to Figure 3.
Column(2): the total number of a given AGN type. Column(3):
The probability from the K-S test that the given AGN type has
the same distribution as a theoretical Monte Carlo distribution
produced by the correlation defined by the Seyfert galaxies. Col-
umn (4): The significance that the distribution of the given AGN
type and the theoretical distribution have the same mean. A value
greater than 0.05 indicates the same mean for the two distribu-
tions. Column (5): The significance that the distribution of the
given AGN type and the theoretical distribution have the same
variance. A value greater than 0.05 indicates the same variance
for the two distributions.
-- 15 --
Fig. 1. -- IRS spectra of 3 objects with fits. The dashed line is the fitted continuum and
the dotted line is the estimated uncertainty in the continuum fitting. The scaling factors for
(a) MCG-2-58-22, (b) Fairall9 and (c) NGC4388, are 30, 5 and 1.5, respectively.
-- 16 --
Fig. 2. -- The IRS spectra for objects with available X-ray data in order of the HI column
density. Narrow emission lines have been removed from the spectra (see text). The dashed
lines are the fitted continua and the dotted lines show the estimated uncertainties in the
continuum fitting. The scaling factor to normalize the spectra at 5 µm and the logarithm of
the HI column density in units of cm−2 are given for each object.
-- 17 --
Fig. 2. -- Continued
-- 18 --
Fig. 2. -- Continued
-- 19 --
Fig. 2. -- Continued
-- 20 --
Fig. 2. -- Continued
-- 21 --
Fig. 3. -- The strength of the silicate feature as a function of HI column density. The
strength of the silicate feature is defined as (Ff − Fc)/Fc, where Ff and Fc are the observed
flux density and underlying continuum flux density, respectively, at the peak (for emission)
or the minimum (for absorption) of the silicate feature. The black line is the linear fit to all
objects while the red line is the fit to Seyfert galaxies. ′PG′: PG quasar; ′RLQ′: radio-loud
quasar; ′Sy1′: Seyfert 1 galaxies; ′Sy2′: Seyfert 2 galaxies; ′BALQ′: broad absorption-line
quasar; ′FRII′: FR II radio galaxies; ′LINER′:
low-ionization nuclear emission-line region;
′2MQ′: 2MASS Quasar.
-- 22 --
Fig. 4. -- The composite spectra of AGNs in different HI column bins. The number in
parenthesis is the number of objects used for the composite spectrum in each bin.
-- 23 --
Fig. 5. -- The structure of the material surrounding the central blackhole in the first quarter
section. The whole structure is symmetric about the disk axis and the equatorial plane. From
inside to outside: 1.) the inner accretion disk (AD), which produces X-ray and UV radiation
ionizing the narrow-emission-line (NEL) and broad-emission-line clouds, and heating the
dust; 2.) the middle disk (MD) with a diffuse component (grey) and with denser embedded
clouds -- the diffuse component produces the silicate emission while the embedded clouds
heavily obscure the central X-ray emission when the line of sight intercepts them; and 3.) the
outer disk (OD) with clouds that obscure the silicate and X-ray emission, and are responsible
for the far-IR emission. Four lines of sight indicate: 1.) silicate emission with low HI column;
2.) silicate emission with high HI column; 3.) silicate absorption with low HI column; 4.)
silicate absorption with high HI column.
|
astro-ph/9401011 | 1 | 9401 | 1994-01-10T17:05:00 | False Vacuum Inflation with Einstein Gravity | [
"astro-ph",
"gr-qc",
"hep-th"
] | We investigate chaotic inflation models with two scalar fields, such that one field (the inflaton) rolls while the other is trapped in a false vacuum state. The false vacuum becomes unstable when the inflaton field falls below some critical value, and a first or second order transition to the true vacuum ensues. Particular attention is paid to Linde's second-order `Hybrid Inflation'; with the false vacuum dominating, inflation differs from the usual true vacuum case both in its cosmology and in its relation to particle physics. The spectral index of the adiabatic density perturbation can be very close to 1, or it can be around ten percent higher. The energy scale at the end of inflation can be anywhere between $10^{16}$\,GeV and $10^{11}$\,GeV, though reheating is prompt so the reheat temperature can't be far below $10^{11}\,$GeV. Topological defects are almost inevitably produced at the end of inflation, and if the inflationary energy scale is near its upper limit they can have significant effects.
Because false vacuum inflation occurs with the inflaton field far below the Planck scale, it is easier to implement in the context of supergravity than standard chaotic inflation. That the inflaton mass is small compared with the inflationary Hubble parameter is still a problem for generic supergravity theories, but remarkably this can be avoided in a natural way for a class of supergravity models which follow from orbifold compactification of superstrings. This opens up the prospect of a truly realistic, superstring | astro-ph | astro-ph |
KUNS 1207
LANCASTER-TH 9401
SUSSEX-AST 94/1-1
astro-ph/9401011
(January 1994)
False Vacuum Inflation with Einstein Gravity
Edmund J. Copeland1 , Andrew R. Liddle1 , David H. Lyth2 ,
Ewan D. Stewart3 and David Wands1
1 School of Mathematical and Physical Sciences,
University of Sussex,
Falmer, Brighton BN1 9QH, U. K.
2 School of Physics and Materials,
Lancaster University,
Lancaster LA1 4YB, U. K.
3Department of Physics,
Kyoto University,
Kyoto 606,
Japan
Abstract
We present a detailed investigation of chaotic inflation models which feature two scalar fields,
such that one field (the inflaton) rolls while the other is trapped in a false vacuum state. The
false vacuum becomes unstable when the magnitude of the inflaton field falls below some critical
value, and a first or second order transition to the true vacuum ensues. Particular attention
is paid to the case, termed ‘Hybrid Inflation’ by Linde, where the false vacuum energy density
dominates, so that the phase transition signals the end of inflation. We focus mostly on the case
of a second order transition, but treat also the first order case and discuss bubble production in
that context for the first time.
False vacuum dominated inflation is dramatically different from the usual true vacuum case,
both in its cosmology and in its relation to particle physics. The spectral index of the adiabatic
density perturbation originating during inflation can be indistinguishable from 1, or it can be
up to ten percent or so higher. The energy scale at the end of inflation can be anywhere
between 1016 GeV, which is familiar from the true vacuum case, and 1011 GeV. On the other
hand reheating is prompt, so the reheat temperature cannot be far below 1011 GeV. Cosmic
strings or other topological defects are almost inevitably produced at the end of inflation, and
if the inflationary energy scale is near its upper limit they contribute significantly to large scale
structure formation and the cosmic microwave background anisotropy.
Turning to the particle physics, false vacuum inflation occurs with the inflaton field far below
the Planck scale and is therefore somewhat easier to implement in the context of supergravity
than true vacuum chaotic inflation. The smallness of the inflaton mass compared with the
inflationary Hubble parameter still presents a difficulty for generic supergravity theories. Re-
markably however, the difficulty can be avoided in a natural way for a class of supergravity
models that follow from orbifold compactification of superstrings. This opens up the prospect
of a truly realistic, superstring derived theory of inflation. One possibility, which we show to
be viable at least in the context of global supersymmetry, is that the Peccei-Quinn symmetry
is responsible for the false vacuum.
PACS numbers:
98.80.Cq, 04.50.+h
1
1
Introduction
An attractive proposal concerning the first moments of the observable universe is that of chaotic
inflation [1]. At some initial epoch, presumably the Planck scale, the various scalar fields existing in
nature are roughly homogeneous and dominate the energy density. Their initial values are random,
sub ject to the constraint that the energy density is at the Planck scale. Amongst them is the inflaton
field φ, which is distinguished from the non-inflaton fields by the fact that the potential is relatively
flat in its direction. Before the inflaton field φ has had time to change much, the non-inflaton fields
quickly settle down to their minimum at fixed φ, after which inflation occurs as φ rolls slowly down
the potential.
Two possibilities exist concerning the minimum into which the non-inflaton fields fall. The
simplest possibility is that it corresponds to the true vacuum; that is, the non-inflaton fields have
the same values as in the present universe.
Inflation then ends when the inflaton field starts to
execute decaying oscillations around its own vacuum value, and the hot Big Bang (‘reheating’)
ensues when the vacuum value has been achieved and the decay products have thermalised. This
is the usually considered case, which has been widely explored. The other possibility is that the
minimum corresponds to a false vacuum, with non-zero energy density. This case may be called
false vacuum inflation, and is the sub ject of the present paper.
There are two fundamentally different kinds of false vacuum inflation, according to whether the
energy density is dominated by the false vacuum energy density or by the potential energy of the
inflaton field. (For simplicity we discount for the moment the intermediate possibility that the two
contributions are comparable, though it will be dealt with in the body of the paper.) In all cases
the false vacuum exists only when the value of the inflaton field is above some critical value.
If
the false vacuum energy dominates, a phase transition occurs promptly when the inflaton field falls
below the critical value, causing the end of inflation and prompt reheating. The result is a new
model of inflation which is dramatically different from the usual one, and at least as attractive. It
was first studied by Linde who termed it ‘Hybrid Inflation’, and it is the main focus of the present
paper. The phase transition may be of either first or second order. A first-order model of false
vacuum dominated inflation has been considered by Linde [2] and (with minor differences but more
thoroughly) by Adams and Freese [3]. A second-order model has been discussed by Linde [4, 5] and
explored in a preliminary way by Liddle and Lyth [6] and by Mollerach, Matarrese and Lucchin
[7]. As far as we know these are the only references in the literature to false vacuum dominated
inflation with Einstein gravity. Related models have been considered at some length in the context
of extended gravity theories [8, 9, 10]; although such theories can be recast as Einstein gravity
theories by a conformal transformation, the resulting potentials are of a different type and this case
is excluded from the present paper.
The opposite case where the false vacuum energy is negligible (inflaton domination) is indistigu-
ishable from the true vacuum case for couplings of order unity, though a variety of exotic effects can
occur for small couplings. This case has been studied by several authors [11, 12, 13, 14, 15, 16, 17,
18, 19], and in the present paper it is treated fairly briefly.
From the viewpoint of cosmology, false vacuum dominated inflation differs from the usual true
vacuum case in three important respects.
1. The spectral index n of the adiabatic density perturbation is typically very close to the scale
invariant value 1, and is in any case greater than 1. This is in contrast with other working
models of inflation, where one typically finds n < 1, viable models covering a range from
perhaps n ≃ 0.7 up to n ≃ 1 [6]. We shall however note that the extent to which n can exceed
unity is quite limited, contrary to claims in Refs. [5, 7].
2. Topological defects generally form at the end of inflation, in accordance with the homotopy
groups of the breaking of the false vacuum to degenerate states, provided that these groups
2
exist. The defects may be of any type (domain walls, gauge or global strings, gauge or global
monopoles, textures or nontopological textures).
3. Reheating occurs promptly at the end of inflation. In the simple models that we have explored,
this means that the reheat temperature is at least 1011 GeV. One consequence is that a long
lived gravitino must be either rather heavy (m ∼> 1 TeV) or extremely light, so as not to be
overproduced [20].
False vacuum dominated inflation is also very different from the true vacuum case from the
viewpoint of particle physics. Sticking to the chaotic inflation scenario already described, let us
consider as a specific example the inflationary potential
V (φ) = V0 +
1
2
m2φ2 +
1
4
λφ4 .
(1.1)
where V0 is the false vacuum energy density. Consider first the true vacuum case, where V0 vanishes.
Inflation occurs while φ rolls slowly towards zero, and it ends when φ begins to oscillate, which occurs
when φ is of order the Planck mass. In order to have sufficiently small cosmic microwave background
(cmb) anisotropy, one needs m ∼< 1013 GeV and λ ∼< 10−12 , with one or conceivably both of these
limits saturated if inflation is to actually generate the observed anisotropy (and a primeval density
perturbation leading to structure formation). To achieve the small λ in a natural way one should
invoke supersymmetry. As long as one sticks to global supersymmetry this presents no problem, but
there are sound particle physics reasons for invoking instead local supersymmetry, which is termed
supergravity because it automatically includes gravity. In the context of supergravity, the fact that
φ is of order the Planck mass during inflation is problematical, because in this regime it is difficult
to arrange for a sufficiently flat potential.
As will become clear, things are very different in the false vacuum case. One still needs to have
λ very small, and will still therefore wish to implement inflation in the context of supergravity. But
now φ is far below the Planck scale during inflation (after the observable universe leaves the horizon
which is the cosmologically interesting era). As a result it becomes easier to construct a viable model
of inflation, though the smallness of m in relation to the inflationary Hubble scale H still presents a
severe problem for generic supergravity theories. Remarkably though, it turns out that among the
class of supergravity models emerging from orbifold compactifications of superstring theory, one can
find a large subset for which this problem disappears. As a toy model, we will see how things work
out with a specific choice for the perturbative part of the superpotential.
Another crucial difference concerns the mass m.
In contrast with the true vacuum case, the
cmb anisotropy does not determine m in the vacuum dominated case, but rather determines V0 as
a function of m. The value m ∼ 1013 GeV that obtains in the true vacuum case is allowed as an
upper limit, but m can be almost arbitrarily small and it is natural to contemplate values down to
at least the scale m ∼ 100 GeV. The value of m chosen by nature might be accessible to observation
because it determines the spectral index n; if m is within an order of magnitude or so of its upper
limit n is appreciably higher than 1, whereas if it is much lower n is indistinguishable from 1. In
the superstring motivated models mentioned earlier, the first case probably obtains if the slope
of the inflationary potential is dominated by one-loop corrections coming from the Green-Schwarz
mechanism, in which case the value of n is determined by the orbifold. This would open up the
interesting possibility that observations of the cmb anisotropy and large scale structure provide a
window on superstring physics.
The opposite case m ∼ 100 GeV is also interesting. Supersymmetric theories of particle physics
typically contain several scalar fields with this mass. The corresponding false vacuum energy scale
V 1/4
0 ∼ 1011 GeV also appears in particle physics, as that associated with Peccei-Quinn symmetry,
a global U (1) symmetry which is perhaps the most promising explanation for the observed CP
invariance of the strong interaction. This same symmetry provides the axion, which is one of the
leading dark matter candidates, and the possibility that it might in addition provide the false
3
vacuum for inflation is to say the least interesting. We explore this possibility in the context of
global supersymmetry and find that it can easily be realised there. We have not gone on to explore
it in the context of supergravity, but there seems to be no reason why it should not be realised
within the context of the superstring derived models considered earlier.
As will be clear from this introduction, the present work is expected to be of interest to a very
wide audience, ranging from observational astronomers to superstring theorists. With this in mind
we have tried to keep separate the part of the paper that discusses the phenomenology of the false
vacuum inflation models, and the part that relates these models to particle physics.
The outline of the paper is as follows. Section 2 introduces the specific second-order model
upon which most of our discussion shall be focussed. We analyse the inflationary dynamics and
density perturbation constraints by a combination of analytic and numerical methods to delineate the
observationally viable models. Section 3 then takes our attention onto the formation of topological
defects, which (almost) inevitably form at the end of inflation. Their possible existence constrains
the models, and there is the further opportunity of a reconciliation of structure-forming defects
with inflation. In Section 4 we try to realise the model in the contexts of global supersymmetry,
supergravity and superstring derived supergravity. In Section 5 we consider the related first-order
model which also indicates the link with extended inflation models. Section 6 summarises the paper.
2
Inflationary Phenomenology
2.1 The Model
V (φ, ψ) =
Throughout this paper we assume Einstein gravity. During inflation the energy density is supposed
to be dominated by the potential of two scalar fields, which is taken to be of the form
1
1
1
λ (cid:0)ψ2 − M 2 (cid:1)2
4
2
2
This potential possesses the symmetries φ ↔ −φ and ψ ↔ −ψ , and is the most general renormalis-
able potential with this property except for a quartic term λ′′φ4 .1
We make the restrictions 0 < λ, λ′ ∼< 1, and we also require that the masses m and M fall in
the range between 100 GeV and the Planck scale mPl/√8π = 2.4 × 1018 GeV indicated by particle
physics considerations.2
Provided that φ2 > φ2
inst , where
λ′φ2ψ2 .
m2φ2 +
(2.1)
+
φ2
inst = λM 2 /λ′ ,
there is a local minimum at ψ = 0 on the constant φ slices, corresponding to a false vacuum. Our
assumption is that inflation occurs with the ψ field sitting in this false vacuum, so that the potential
is
1
1
2
4
If the false vacuum dominates, inflation ends when φ falls below φinst , the fields rapidly adjusting
to their true vacuum values ψ = M and φ = 0.
This model was first considered by Kofman and Linde [11], who pointed out that it might produce
cosmic strings with enough energy per unit length to form structure. They considered only what
we shall term the inflaton dominated regime (small false vacuum energy), as did several subsequent
authors studying this and related models [12, 13, 14, 15, 16, 17, 18, 19]. In order to obtain interesting
λM 4 +
m2φ2 .
V (φ) =
(2.3)
(2.2)
1The pure quadratic term is the simplest possibility, and is also the one favoured by particle physics considerations
(Section 4). Non-renormalisable potentials, involving higher powers of the fields, arise naturally in the context of
supergravity, but for simplicity we ignore them here. As we discuss later the ψ field can have several components, but
they do not affect the issues we discuss in the present section.
2The factor √8π is mathematically convenient and we shall follow the ma jority of authors by inserting it, though
of course our understanding of the Planck scale is quite insufficient to justify such factors from a physical viewpoint.
4
effects, these authors had to assume (at least) that the coupling λ′ was many orders of magnitude
less than unity. The case of false vacuum domination, which is our main focus, was proposed by
Linde who termed it ‘Hybrid Inflation’ [4] and has received further attention from Liddle and Lyth
[6], Linde [5], and Mollerach, Matarrese and Lucchin [7]. In this case the couplings can be of order
unity, but for completeness we explore also the regime of parameter space where they are very
different from one.
2.2 Inflationary dynamics
,
,
;
,
1
2
(2.5)
(2.6)
(2.4)
φ2 +
H 2 =
As usual, the inflationary dynamics are governed by the equations
Pl (cid:18) 1
ψ2 + V (φ, ψ)(cid:19) ,
8π
3m2
2
∂ V (φ, ψ)
φ + 3H φ = −
∂φ
∂ V (φ, ψ)
ψ + 3H ψ = −
∂ψ
for two isotropic scalar fields in an expanding universe, with H = a/a the Hubble parameter, a the
scale factor, mPl the Planck mass and dots derivatives with respect to time. Our assumption is that
there is a transitory regime during which the ψ field rolls to ψ = 0 from whatever its initial value
may have been, and is followed by sufficient inflation on the ψ = 0 tra jectory to erase any evidence
of such a transient. Inflation then proceeds according to the usual single field equation for φ in the
potential of Eq. (2.3). Without loss of generality, we shall assume that φ is initially positive.
We shall utilise the slow-roll approximation throughout. It is characterised by the conditions
η ≪ 1 ,
ǫ ≪ 1
where the two dimensionless functions ǫ(φ) and η(φ) are defined by
V (φ) (cid:19)2
m2
16π (cid:18) V ′ (φ)
Pl
m2
V ′′ (φ)
Pl
.
8π
V (φ)
η(φ) ≡
Here and throughout primes indicate derivatives with respect to the field φ. With justification from
numerical results, it is standard to assume that if the potential satisfies these conditions, then the
solutions for a broad range of initial conditions rapidly approach the attractor
3H φ ≃ −V ′ .
When this is satisfied, there exists a simple expression for the number N of e-foldings of expansion
which occur between two scalar field values φ1 and φ2
ǫ(φ) ≡
(2.10)
(2.7)
(2.8)
(2.9)
N (φ1 , φ2 ) ≡ ln
For our specific potential we have
a2
a1 ≃ −
Pl Z φ2
8π
m2
φ1
V
V ′
dφ .
η =
ǫ =
N (φ1 , φ2 ) =
,
m2m2
Pl
2π (λM 4 + 2m2φ2 )
m2
Plm4φ2
8π
1
π (λM 4 + 2m2φ2 )2 =
m2
2
Pl
2πλM 4
2π
φ1
1 − φ2
Pl (cid:0)φ2
2 (cid:1) .
m2
m2m2
φ2
Pl
ln
+
η2φ2 ,
5
(2.11)
(2.12)
(2.13)
(2.14)
Within the slow-roll approximation, the condition for inflation to occur is simply that ǫ be
less than one. However, slow-roll is automatically a poor approximation should ǫ reach this value,
though the amount of inflation that occurs as ǫ becomes large is always small. Numerical simulation
indicates that for this potential if ǫ and η grow to unity, shortly thereafter the inflationary condition
a > 0 is violated and inflation ends. The number of e-foldings that occur between these events is a
tiny fraction of unity, and can be ignored. It is therefore sensible operationally to identify the end
of inflation in this case with the precise condition that ǫ = 1, should this occur, and we shall assume
this subsequently.
There are therefore two separate ways in which inflation may end in this model, the one which
is applicable depending on the parameter values. These are
1. If φ reaches φinst while inflation is occurring, then inflation may end through the instability of
the ψ field to roll to its global minimum. As noted by Linde [5] one expects this to happen, at
least for λ and λ′ not too small, if the false vacuum term λM 4/4 dominates the potential. We
look in some detail at this possibility in Section 3, confirming the picture of rapid instability.
2. If the logarithmic slope of the potential becomes too large on the ψ = 0 trajectory, then
inflation can end while the φ field is still rolling down that tra jectory. This is symptomised
by ǫ growing to exceed unity. Some time later, φ will pass φinst and the ψ field may roll away
from ψ = 0.
φǫ =
(2.15)
The value of φ at which ǫ becomes equal to unity is3
mPl√16π 1 + s1 −
m2 ! .
λM 4
Plm2 > 1, then φǫ does not exist at all. In that case inflation must end by instability.
If 8πλM 4 /m2
In the opposite limit, the position φǫ → mPl/√4π is familiar from chaotic inflation with a single
field, and of course the standard results will be recovered in that limit with the ψ field playing no
significant role.
We need to know the number N (k) of Hubble times of inflation which occur after a given scale
leaves the horizon4 . With the assumptions (valid in our model) that H does not vary significantly
and that reheating is prompt it is given by [6]
8π
m2
Pl
N (k) = 62 − ln
1016 GeV
V 1/4
1
− ln
k
a0H0
,
(2.16)
where subscript ‘0’ indicates present value. The largest cosmologically interesting scale is of order
the present Hubble distance (roughly the size of the observable universe), k = a0H0 , and other
scales of cosmological interest leave the horizon at most a few Hubble times after this one. As for
the inflationary energy scale, true vacuum inflation typically gives V 1/4
1 ∼ 1016 GeV, which makes
the observable universe leave the horizon about 60 e-folds before the end of inflation (the fact that
reheating may be very inefficient in this model may reduce this number somewhat). As we shall
see, false vacuum dominated inflation can give values as low as V 1/4
1 ∼ 1011 GeV, which reduces the
figure 60 to about 50. However, one only needs a rough estimate of N for most purposes because the
potential is slowly varying, and for simplicity we suppose from now on that cosmologically interesting
scales leave the horizon 60 e-folds before the end of inflation.
Provided the parameters are chosen in such a way as to produce the correct level of density
perturbations to explain the COsmic Background Explorer (COBE) satellite observations [21] of
3There is also a second root at smaller φ, where ǫ drops back below unity. However, for second-order models it
is easy to show that the attractor solution Eq. (2.10) cannot be attained for φ below this root, allowing inflation to
restart, before the instability sets in.
4As usual we say that a comoving scale a/k leaves the horizon when aH/k = 1
6
the cmb anisotropies, there is no problem in obtaining sufficient inflation to resolve the horizon and
flatness problems or with ensuring that a classical description of the evolution is adequate. We thus
need only investigate the density perturbation constraint in order to completely fix the model.
2.3 Density perturbations
The adiabatic density perturbation, which is generally thought to be responsible for large scale
structure, originates as a vacuum fluctuation during inflation.
Its spectrum is determined by a
quantity δH , which loosely speaking gives the density contrast at horizon crossing and is defined
formally in [6]. The inflationary prediction for the spectrum is
,
32
75
δ2
H (k) =
1
V∗
m4
ǫ∗
Pl
where ǫ is the slow-roll parameter defined earlier and the subscript ∗ indicates that the right hand
side is to be evaluated as the comoving scale k equals the Hubble radius (k = aH ) during inflation.
By virtue of the slow-roll conditions Eqs. (2.7)-(2.10), this formula gives a value of δH (k) which is
nearly independent of k on scales of cosmological interest, in agreement with observation. For a suf-
ficiently flat spectrum, and provided that no significant generation of long wavelength gravitational
wave modes occurs, the central value of the COBE 10◦ anisotropy, 30µK , is reproduced provided
one has δH = 1.7 × 10−5 [6]5 . Thus the inflationary energy scale when cosmologically interesting
scales leave the horizon is given by
(2.17)
60 = 6ǫ1/4
V 1/4
60 × 1016 GeV ,
where a subscript 60 denotes 60 e-folds before the end of inflation.
The most efficient way to proceed is as follows. First, fix the couplings λ and λ′ . Then, having
chosen a value for the mass scale M , find the value(s) of m such that the density perturbation
constraint is satisfied. Assuming that inflation ends promptly if φ falls below φinst , we can determine
the means by which inflation ends and the corresponding value of φ
(2.18)
.
(2.20)
(2.19)
δ2
H =
φend = max{φǫ , φinst} .
We then use Eq. (2.14) to determine the value of φ 60 e-foldings from the end of inflation, φ60 , and
evaluate δH as
60 (cid:1)3
75 (cid:0)λM 4 + 2m2φ2
8π
m6
Plm4φ2
60
To find the value(s) of m which satisfy the COBE normalisation, remember that φend , and hence φ60 ,
is a function of m. In general this procedure cannot be carried out analytically, and we compute
using an iterative numerical method. However, the problem can be solved analytically and self-
consistently in two regimes. As we shall see, provided M is not too large then for each M there
are two possible choices of m which give the right perturbation amplitude. One corresponds to the
traditional polynomial chaotic inflation scenario, where the first term in Eq. (2.3) plays a negligible
role (and by implication the first term in the numerator of Eq. (2.20) likewise). [There is a variant
on this regime, also discussed below, where the instability sets in while the false vacuum energy is
still negligible.] The second, and for our purposes more interesting, possibility involves a value of
m ≪ M , and corresponds to domination by the first term in Eq. (2.3).
5This figure assumes a Gaussian beam profile, and is raised by 16% if the precise profile of the experiment is used
and a correction applied for the incomplete sky coverage inducing errors in the monopole and dipole subtractions [22].
For an accurate analysis one has to include (here and elsewhere) the effect of spectral tilt and gravitational waves on
the COBE normalisation [23]. Such changes are not significant in the present context except for extreme parameter
values, and for simplicity we shall not include them in the normalisation though we shall discuss tilt and gravitational
waves later.
7
2.4 Delineating parameter space
We shall now examine different analytic and numerical regimes. The results are concisely summarised
in Figure 1.
The inflaton dominated regime
The simplest scenario of all is one in which the ψ field plays no role whatsoever, leaving just the
φ field to govern inflation. The potential V = m2φ2 /2 was proposed by Linde [1] as a simple
realisation of chaotic inflation. With this potential, inflation ends when φ starts to oscillate around
its minimum, which occurs when ǫ ≃ 1 corresponding to
φend ≃ mPl/√4π .
The condition that our potential Eq. (2.3) be a good approximation to this one is therefore
(2.21)
8π
m2
Pl
λM 4
4m2 ≪ 1 .
In Eq. (2.14) the second term dominates, giving
φ60 ≃ r 60
2π
Note that in this regime the characteristic scale of φ60 is the Planck scale. The density perturbation
amplitude is independent of M , λ and λ′ in this limit, and the correct value is obtained with
√8π
mPl
(2.22)
(2.23)
(2.24)
mPl .
π
δH = 5.5 × 10−6 .
4√6
The condition for the validity of this approximation is therefore
√8π
mPl
m =
λ1/4M ≪ 3 × 10−3 .
The above analysis assumes that φ > φinst throughout inflation, which from Eq. (2.21) fails if
> mPl/√4π , or equivalently
φinst ∼
Pl (cid:19)2
λ′2
λM 4 (cid:18) 8π
∼< 10−11 .
λ ∼<
m2
(The final inequality is Eq. (2.25).) If φ falls below φinst , then as discussed in Section 3.1 ψ may roll
towards its minimum at fixed φ,
(2.25)
(2.26)
1
4
(2.27)
φ2
inst (cid:19) .
vac (φ) = M 2 (cid:18)1 −
ψ2
φ2
It oscillates around the minimum, losing energy through the expansion of the universe so that after
a few Hubble times ψ ≃ ψvac (if its spatial gradient is not negligible it may settle down more quickly
through thermalisation). Inserting ψ = ψvac into Eq. (2.1) gives [24]
inst (cid:19)2# ,
M 4 "1 − (cid:18)1 −
φ2
1
m2φ2 +
V (φ) =
φ2
2
λM 4
φ2
V ′ (φ) = m2φ (cid:20)1 +
inst (cid:19)(cid:21) ,
inst (cid:18)1 −
φ2
m2φ2
λM 4
φ2
V ′′ (φ) = m2 (cid:20)1 +
inst (cid:18)1 − 3
inst (cid:19)(cid:21) .
φ2
m2φ2
(2.29)
(2.28)
(2.30)
λ
4
8
Since we are in the regime φinst ∼> mPl /√8π , the condition Eq. (2.22) written down earlier guarantees
again that the modification to V will be negligible.
It therefore appears that when Eq. (2.22) is well satisfied, the evolution of φ will not be signif-
icantly affected even if φ falls below φinst . If Eq. (2.22) is only marginally satisfied, the evolution
of φ might be substantially altered, leading to a significant change in the predicted adiabatic den-
sity perturbation. The simplest assumption is that the potential is given by Eq. (2.28). In that
case, if one ignores the inhomogeneity of φ caused by the phase transition, the perturbation is still
given by the usual formula, Eq. (2.17), with the new potential [24]. However, this formula depends
crucially on the assumption that each Fourier mode of φ is in the vacuum state before leaving the
horizon, whereas the phase transition will inevitably populate some of the modes with non-zero
particle number. Taking this into account, the adiabatic perturbation on scales leaving the horizon
after φ = φinst might be quite different (non-Gaussian, with a non-flat spectrum and a different
normalisation). An additional adiabatic perturbation might also be generated by the perturbation
in ψ [13, 15, 16], as discussed in Section 3.1.
The vacuum dominated regime
We now explore the opposite regime, where the vacuum energy density notionally associated with
the ψ field dominates the potential. Special cases of this regime of parameter space have already
been considered in [4]–[7].
As noted earlier, inflation is expected to occur only if η and ǫ are small compared with unity.
The first of these parameters is independent of φ in the limit of vacuum domination, with the value
4m2
m2
Pl
λM 4 .
(2.31)
η =
8π
Thus, the requirement that η ∼< 1 in this regime is precisely the opposite of the condition Eq. (2.22)
which characterises the regime in which the vacuum does not dominate. The parameter ǫ decreases
as inflation proceeds, and during the era φ < φ60 that we are interested in we have
8π
1
m2
2
Pl
ǫ < ǫ60 =
η2φ2
60 .
(2.32)
The condition for vacuum domination is
1
2
8π
m2
Pl
φ2
60 η ≪ 1 ,
or equivalently
ǫ60 ≪ η .
The first term of Eq. (2.14) dominates the formula for φ60 , giving
(2.33)
(2.34)
(2.35)
φ2
60 =
M 2 e120η .
λ
λ′
The COBE normalisation Eq. (2.20) is therefore
√8π
√λ′M = 10√3πδH ηe60η ,
mPl
= 9.3 × 10−4ηe60η .
It involves the two masses and the two coupling constants only in the dimensionless combinations
√8π
√λ′M ,
mPl
√8π
mPl r λ′2
λ
9
M ≡
m ≡
m ,
(2.36)
(2.37)
(2.38)
(2.39)
(2.40)
(2.41)
since η = 4 m2/ M 4 . The restrictions M ∼< mPl/√8π and λ′ ∼< 1 that we have agreed to impose
because of particle physics considerations means that we are in the regime M ∼< 1, which with the
COBE constraint corresponds to η ∼< 0.15.
The situation becomes especially simple in the regime 60η ≪ 1, which we call the extreme vacuum
dominated regime. It corresponds to M ≪ 10−4 , and the quantity η varies linearly with M leading
to [4, 5]
M 5
m2 = 40√3πδH = 3.7 × 10−3 .
Inserting the Planck mass and working with the masses themselves this formula becomes
= λ′−1/10 λ−1/5 (cid:16) m
M
1 TeV (cid:17)
5.5 × 1011 GeV
In the other regime 60η ∼> 1 [6, 7], η varies only logarithmically with M , and the power m2/5
gradually changes to m1/2 .
Although the cmb constraint can be expressed in terms of just the two quantities m and M ,
the vacuum domination condition involves three quantities which are conveniently chosen to be m,
λ1/4M and λ′2 /λ. It is therefore useful to express the cmb constraint as a constraint on m and
λ1/4M at fixed λ′2 /λ. (Another good reason for doing this is that λM 4 is the false vacuum energy
density.) The extreme vacuum dominated regime 60η ≪ 1 corresponds to
λ (cid:19)−1/4
λ1/4M ≪ 4 × 10−5 (cid:18) λ′2
mPl√8π
In this regime η increases linearly with λ1/4M ,
λ (cid:19)1/4 √8π
m2
λM 4 ≃ 270 (cid:18) λ′2
η
4 ≡
mPl
while for larger values it increases only logarithmically giving the normalisation
m2
m2
Pl
λM 4 ≃ 0.004 to 0.04
8π
with the upper limit corresponding to η = 0.15.
We have yet to invoke the false vacuum domination condition Eq. (2.33). Using Eq. (2.37), it
becomes
η3 e240η ≪ 2 × 106 λ′2
λ
With λ′ 2/λ = 1, this bound is saturated for η = 0.09, and a similar limit is obtained for any value
of the ratio within a few orders of magnitude of unity. Setting η = 0.09, one learns that the false
vacuum dominated regime is restricted to
√8π
λ (cid:19)−1/4
λ1/4M ∼< 2 × 10−2 (cid:18) λ′2
mPl
√8π
λ (cid:19)−1/2
m ∼< 8 × 10−5 (cid:18) λ′2
mPl
The upper limit on λ1/4M is not far below the one following just from the fact that ǫ∗ < 1 in
the cmb constraint Eq. (2.17), which using 1
4 λM 4 < V is6
√8π
mPl
λ1/4M < 5 × 10−2 .
6To understand the relation between the two limits, note that when the vacuum domination condition Eq. (2.33)
is saturated, V = 1
2 λM 4 and ǫ60 = η.
m2
Pl
8π
λ1/4M ,
(2.44)
(2.46)
(2.47)
(2.42)
(2.43)
.
(2.45)
(2.48)
2
5
.
.
,
.
10
The intermediate regime
We have now investigated the extreme cases, first the one in which the false vacuum energy is
negligible, and second the one in which it dominates. There remains the intermediate case where
both are comparable, during at least part of the cosmologically significant era φ < φ60 .
Plotted with m the vertical axis and λ1/4M the horizontal axis, we have learned that the first
regime corresponds to a straight horizontal line, whereas the second one corresponds to a line with
positive slope. Unless λ′2 /λ is several orders of magnitude away from unity, comparison of Eqs. (2.24)
and (2.25) with Eqs. (2.47) and (2.46) shows that the right hand ends of these two lines are separated
by at most an order of magnitude or so in the m and λ1/4M variables. Therefore the intermediate
regime is not very extensive, but it is still important to investigate it in order to see if new physics
occurs.
Even in the intermediate regime the upper bound Eq. (2.48) holds. Apart from this fact, numer-
ical techniques are required to solve the density perturbation constraint. The solution, as might be
expected, is that as λ1/4M is increased the two solution branches approach each other and merge
continuously. This merger specifies the maximum allowed value of λ1/4M ; for higher values it be-
comes impossible to obtain a sufficiently low perturbation amplitude regardless of the choice of m.
(The maximum value does depends on λ and λ′ , of course.) Figure 1 illustrates the complete set
of viable models for the couplings both set to unity, showing both the asymptotic regimes and the
merger region.
2.5 Tilt and gravitational waves
Although the inflationary prediction in Eq. (2.17) for the spectrum δH (k) is almost flat, there is
always some k-dependence, usually referred to in the literature as tilt. On cosmologically interesting
scales the tilt can usually be well characterised by a constant spectral index n, such that δ2
H ∝ kn−1 ,
and in that case one learns from the slow-roll conditions Eqs. (2.7)-(2.10) that [25, 6]
n − 1 = 2η60 − 6ǫ60 .
As always, we take ‘cosmologically interesting’ to mean scales that leave the horizon 60 e-folds before
the end of inflation.
In addition to the adiabatic density (scalar) perturbation, inflation also generates gravitational
waves, whose contribution R to the cmb anisotropy (∆T /T )2 relative to that of the scalar modes is
[26, 25, 6]
(2.49)
R ≃ 12ǫ60 .
(2.50)
For true vacuum inflation with a potential V ∝ eAφ , η = 2ǫ so that n is less than 1 and
R ≃ 6(1 − n). Replacing the exponential by a power φα gives tilt n − 1 = −(2 + α)/120, still
negative, and provided that α ≥ 2 as required by particle physics it still gives R ≃ 6(1 − n). Thus,
true vacuum inflation with a φα potential typically makes n a few percent below unity and it makes
R tens of percent. Both of these predictions are big enough to be cosmologically significant.
The case of false vacuum dominated inflation is dramatically different. The condition Eq. (2.33)
for false vacuum domination is ǫ ≪ η , and unless it is almost saturated η is very small. As a result,
the tilt and gravitational wave contribution are both indistinguishable from zero, for generic choices
of the parameters. For fixed couplings, they become significant only at the upper end of the mass
range allowed by Eq. (2.33), and in contrast with the true vacuum case n is greater than one until
Eq. (2.33) is almost saturated.
The value of n for a given parameter choice is obtained by substituting Eq. (2.35) into Eq. (2.49).
With λ′2 /λ = 1, n is equal to 1.0001 for the minimum value m ∼ 100 GeV at the lower end, and
rises to a maximum value n = 1.14 near the upper end of the allowed range [6]. The biggest possible
value of n, corresponding to η = 0.15 and ǫ ≪ 1, is n = 1.30, but this is only achieved with extreme
values of the parameters M ∼ mPl/√8π and λ′ 2/λ ∼> 109 .
11
We have extended these results numerically to the intermediate regime for the case where the
couplings are unity. The result for n is shown in Figure 2 as a function of λ1/4M . The two solution
branches are as in Figure 1. The analytic vacuum dominated result agrees well with the exact one
until well beyond the maximum, and in particular the maximum value n ≃ 1.14 is essentially the
same. This number is considerably less than that which was suggested could be obtained in this
model by other authors [5, 7]; they extrapolated the vacuum dominated case beyond its regime of
validity and neglected ǫ to obtain their larger values. As we commented above, this conclusion is
not altered unless the couplings are changed (and separated) by orders of magnitude.
Another interesting feature is the dip in n to around 0.92 as one exits from the inflaton dominated
regime into the intermediate region, indicating that these models are also capable of providing a
significant (though not startling) tilt in the opposite direction. In the limit of small M , the vacuum
dominated case asymptotes to unity and the inflaton dominated case to the standard value 0.97.
We have also calculated the gravitational wave component, though we have not attempted to
include it (or the tilt) into the COBE normalisation. Gravitational waves make only a small contri-
bution to COBE except in the intermediate regime where they can reach a peak of tens of percent
(though as expected when n exceeds unity the gravitational wave component is suppressed by the
small ǫ required), indicating that a proper treatment of them is required to develop the precise
phenomenology of the intermediate regime.
3 The Second-Order Phase Transition
If φ falls below φinst before the end of inflation, the false vacuum is destabilized and there is a
possibility of a second-order phase transition, of a kind quite different from the usual thermal phase
transition. In this section we consider the nature of this transition, treating separately the very
different regimes of inflaton domination and vacuum domination.
For simplicity we continue to suppose that ψ is a single real field with the potential Eq. (2.1).
This potential has the discrete symmetry ψ ↔ −ψ , which of course implies that the phase transition
creates domain walls located at surfaces in space where ψ vanishes. One can instead have N real
fields, and replace ψ2 by P ψi 2 in the potential Eq. (2.3), which has O(N ) symmetry. This will
give global strings (N = 2), global monoples (N = 3) or textures (N ≥ 4) if the symmetry is
global, or gauge strings (N = 2) or gauge monopoles (N = 3) if it is local. We expect that in all
cases our discussion of the evolution of φ and ψ should be roughly correct, provided that ψ is taken
to represent the ‘radial’ degree of freedom with respect to which the potential has a maximum as
opposed to the ‘angular’ degrees of freedom with respect to which it is constant.
One other significance of having a local symmetry rather than a global one, emphasised by Linde
[5], is that one might have no defects forming simply because the lowest homotopy groups all vanish.
This takes advantage of there being no such thing as local texture or nontopological texture, due to
the gauge degrees of freedom cancelling the scalar gradients. For global symmetries however, scalar
gradients can still play a harmful role even in the absence of topological constraints.
We will see that the formation of topological defects at the end of a period of inflation offers the
intriguing possibility that we could make use of the inflationary epoch to solve the flatness issues of
our universe, and yet retain the possibility of utilising defects as the source of the density fluctuations
to seed large scale structure.
3.1 The inflaton dominated regime
In Section 2.4, we saw that in the inflaton dominated regime the false vacuum is maintained right up
to the end of inflation unless λ′ is very small. In that case the phase transition to the true vacuum
will take place after inflation, and will presumably be of the usual thermal type, any topological
defects forming by the usual Kibble mechanism [27].
12
If λ′ is sufficiently small, φ can fall below φinst before the end of inflation. Depending on the
regime of parameter space, the transition to the false vacuum and the formation of defects may then
occur before the end of inflation. This case has been treated by several authors7 [11, 12, 13, 14, 15,
16, 17, 18, 19], and we look briefly at the results of these authors because they provide a starting
point for our discussion of the vacuum dominated case, which is our main focus. As we noted in
Section 2.4, the back reaction of ψ on φ is negligible in the vacuum dominated regime. As a result we
can in principle follow the evolution of ψ explicitly, using quantum field theory in curved spacetime
[12, 13, 14, 15, 16, 17, 18, 19]. Depending on the values of the couplings λ, λ′ and the mass scale
M , we can have very different situations, indeed even when considering the same regime the authors
cited above are not always in agreement. Still, a reasonably definite picture emerges provided that
the values of the parameters are not too extreme (discounting the necessarily small λ′ of course),
which we now summarise before mentioning more exotic possibities.
For a rough description of what is going on, we can ignore the spatial gradient of ψ , and treat it
classically. As long as it is small in the sense that
its equation of motion is
with
ψ ≪ ψvac ,
ψ + 3H ψ + M 2
ψ (φ)ψ = 0 ,
(3.1)
(3.2)
(3.3)
ψ (φ) = λ′ (φ2 − φ2
M 2
inst ) .
As long as φ > φinst , the effective mass M 2
ψ is positive and ψ is equal to zero apart from its quantum
fluctuation. When φ first falls below φinst , Mψ is negligible compared with H , and φ remains almost
constant. After some time, which may be either small or large on the Hubble scale depending on
the regime of parameter space, Mψ grows to exceed H , and one can start to use the opposite
approximation of ignoring H
ψ + M 2
(3.4)
ψ (φ)ψ = 0 .
There are now two possibilities, according to whether or not the adiabatic condition Mψ ≪ Mψ 2
is satisfied. If it is, the solution of Eq. (3.2) is
ψ ≃ constant × Mψ −1/2 exp (cid:18)Z t
0 Mψ (t)dt(cid:19) ,
Taking t = 0 to be the epoch when Mψ = H , the exponential becomes large within a Hubble
time, and Eq. (3.1) will be violated more or less independently of the initial value of ψ . When that
happens ψ will quickly roll down to its minimum ψvac . If on the other hand the adiabatic condition
is not satisfied when Mψ first grows to be of order H , there will typically be little change in ψ until
it is satisfied, after which Eq. (3.5) will again hold. Thus the conclusion is that ψ rolls down rapidly
towards its vacuum value at the epoch Mψ ∼ H or the epoch Mψ /Mψ 2 = 1, whichever is later.
(The insufficiency of just the former condition was pointed out in [19].)
Though the spatial gradient of ψ is not crucial initially, it becomes so after roll down, because
domain walls form at the places in space where ψ is trapped with its false vacuum value ψ = 0. As
we already noted, more general defects can form if ψ is replaced by an N -component ob ject. To
determine the stochastic properties of the spatial distribution of the defects, one needs to consider
the spatial variation of ψ . The basic assumption is that during inflation ψ vanishes except for its
quantum fluctuation. Once φ falls below φinst , the fluctuation in ψ can be easily evaluated, because
(3.5)
7 Some of these authors consider a coupling of ψ to the spacetime curvature R rather than to the inflaton field, but
this is equivalent for the present purpose.
13
(3.6)
its Fourier modes decouple, until its rms has grown to be of order ψvac . The classical equation for
each mode is the generalisation of Eq. (3.2) including the spatial gradient
ψ (φ)# ψk = 0 ,
ψk + 3H ψk + "(cid:18) k
a (cid:19)2
+ M 2
where k is the comoving wavenumber of the mode under examination. Now cosmologically interesting
(and smaller) comoving scales presumably leave the horizon many Hubble times after inflation begins,
with the corresponding Fourier modes initially in the vacuum, i. e. containing no ψ particles ([6],
page 46). As a result the initial value of the quantum expectation hψk 2 i is known, and so is its
time dependence which is given simply by the modulus squared of the solution of the classical field
equation. For the nonrelativistic modes k/a ≪ Mψ , Eq. (3.6) reduces to Eq. (3.2), and hψk 2 i
begins to grow as the solution Eq. (3.5) becomes valid. At this point, but not earlier, ψk can be
regarded as a classical quantity in the sense that its quantum state corresponds to a superposition
of states with almost well defined values ψk (t) [29]. Thus after smearing the field over a distance
1/Mψ (i. e. dropping the relativistic modes),8 one has a classical field ψ(x) which has a Gaussian
inhomogeneity whose stochastic properties are specified entirely by the spectrum hψk 2 i. Once the
behaviour Eq. (3.5) sets in the spectrum grows rapidly, and the rms of the smeared field rolls down
to ψvac in accordance with the earlier conclusion.
In order for this simple picture to be self consistent, the smeared field ψ must still be small enough
to satisfy Eq. (3.1), at the epoch when the non-relativistic modes ψk begin to grow according to
Eq. (3.5). This can fail to be true if the parameters are far from their natural values, for instance if
λ is very small, and then one has a more complicated situation which has been looked at by various
authors [13, 15, 16, 18, 19].
In particular, the inhomogeneity in ψ might generate an adiabatic
density perturbation on scales far outside the horizon, which would then survive the subsequent
phase transition.9
Topological defect production in the inflaton dominated case
From the stochastic properties of ψ(x) just before roll down, one can in principle calculate the
stochastic properties of the initial configuration of the defects, since they will form at the places
in space where ψ(x) = 0. In particular one can estimate the typical spacing of the defects. The
smallest possible spacing corresponds to the defect size10 , which at least for couplings of order unity
is of order M −1 . For a thermal phase transition the typical spacing at formation is (λM )−1 [27]. We
would like to know if the same is true in the inflationary case. The different estimates [12, 14, 18, 19]
do not entirely agree but they do seem to indicate that the spacing is still very roughly M −1 , at
least to within a few orders of magnitude.
It does not, however, follow that the cosmological effects of the defects are the same in the two
cases, their subsequent evolution being quite different. In the thermal case, where the defects are
created during a non-inflationary (typically radiation dominated) era, the Hubble distance H −1
increases steadily in comoving distance units. Except in the case of gauge monopoles, the spatial
distribution of the defects typically loses all memory of the initial conditions on scales smaller than
the Hubble distance, exhibiting scaling behaviour whereby the stochastic properties become more
or less fixed in units of the Hubble distance. In particular the typical spacing becomes of order the
Hubble radius H −1 . Only on scales much larger than H −1 does the initial distribution expand with
the universe, remaining fixed in comoving distance units. (The case of gauge monopoles, which do
not scale, is considered in greater detail in Section 3.3.)
8 Since Mψ ≫ H , these modes have yet to leave the horizon and so are still in the vacuum state.
9As has already been pointed out [30, 31, 32] in the somewhat different context of axions, failure to take proper
account of the phase transition has led some authors to draw incorrect conclusions from this type of calculation.
10The thickness of a domain wall or string, or the radius of a monopole, outside which ψ has its vacuum value.
14
The case where the defects are created during inflation is quite different. During inflation, H −1
decreases, typically dramatically, in comoving distance units. As a result the distribution of the
defects is frozen in comoving distance units, and in particular the typical spacing remains roughly
of order the comoving distance scale which left the horizon (became bigger than H −1 ) at the epoch
when the defects form. This remains true until the era, long after inflation ends, when that scale
re-enters the horizon. Only then will the defect distribution become the same as in the thermal case,
as the ‘scaling’ solution is established.
The cosmological significance of this different evolution depends on when the defects form. If
they form after cosmologically interesting scales leave the horizon (50 or 60 Hubble times before the
end of inflation), the scaling solution has been established by the time that these scales enter the
horizon and there should be no significant difference from the thermal case. If they form before, their
typical spacing is still much bigger than the horizon size and we presumably see no defects (unless of
course we are in an atypical region of the universe [12, 18]). Finally, if they form at about the same
time, the configuration of the defects will differ from the scaling solution, as has been discussed at
some length in the case of structure forming gauge strings [12, 14, 18].
3.2 The vacuum dominated regime
Coming now to the regime of vacuum domination, Linde [4, 5] has argued that at least for couplings
of order unity inflation will end promptly (within less than a Hubble time) after φ = φinst . He
demonstrated this by assuming that φ continues to slow-roll down the potential Eq. (2.3) for a
Hubble time, and showing that this inevitably leads to a contradiction. We present below a more
detailed version of this argument, repairing an omission in the original and examining also the case
of small couplings.
To proceed, one has to make the technical assumption that the spatial gradients of both φ and ψ
are negligible. It is hard to see how they could be crucial in maintaining slow-roll, so the contradiction
which we shall establish presumably indicates failure of slow-roll rather than significant gradients in
the presence of slow-roll.
The slow-roll expression for φ is
2
mPl√8π
φ = −
√3λ
= −
where we have used the vacuum dominated value for H ,
V ′
3H
m2
M 2 φ ,
H 2 =
2π
3m2
Pl
λM 4 .
During one Hubble time the change in φ is given by
= −4m2
λM 4
φ
H φ
∆φ
φ
=
m2
Pl
8π ≪ 1 .
It follows that after one Hubble time
(3.7)
(3.8)
(3.9)
and
.
M 2
ψ =
8m2
λM 4 λM 2 m2
8π ≪ M 2 ,
Pl
8π (cid:19)2
λM 6 (cid:18) m2
96m2
Mψ 2
Pl
H 2 =
Using Eqs. (2.43) and (2.44), we see that this is much bigger than unity unless we are in the regime
of Eq. (2.43), which then requires
(3.10)
(3.11)
< 10−9 8π
λ′ ∼
m2
Pl
M 2
< 10−9 .
∼
15
(3.12)
We shall not consider these very small values of λ′ . One also finds after one Hubble time
96m2 (cid:19)1/2 8π
Mψ
2 (cid:18) λM 6
1
Mψ 2 =
m2
Pl
Comparing with Eq. (3.11) one sees that except for the factor 1/2 the right hand side is just H/Mψ .
Thus the adiabaticity condition is satisfied as well (it was not considered by Linde). As a result
Eq. (3.5) shows that ψ will have rolled down to become of order ψvac given by Eq. (2.27).
The next step is to demonstrate that the roll-down of ψ is actually inconsistent with slow-roll
inflation. Since only one Hubble time has elapsed ψ will be oscillating around ψvac rather than
sitting in it, but for a crude estimate we can ignore the oscillation, so that the potential is given by
Eq. (2.28). Using Eqs. (2.29) and (2.30) we find
(3.13)
.
8π (cid:19)3
m4
M 10 (cid:18) m2
λ′
Pl
ǫ ≃ 128
λ3
m2
8λ′
Pl
.
λM 2
8π
,
(3.14)
(3.15)
η ≃ −
For couplings of order unity it is easy to check that the cmb constraint Eqs. (2.43) and (2.44) implies
that ǫ and η are both ≫ 1. The slow-roll solution we started with is therefore presumably invalid.
(Even if valid, it is certainly not inflationary since ǫ ≫ 1). More generally, it follows from Eq. (2.48)
that η > 1 unless λ′ ∼< 10−4λ1/2 , which again presumably means that the slow-roll solution is invalid.
(The condition that ǫ > 1 is too complicated to be worth discussing in the general case, but one
expects slow-roll only if both η and ǫ are less than unity.)
The conclusion is that (unless λ′ is very small) slow-roll inflation ends within a Hubble time
of the epoch φ = φinst . Since the field equations contain a mass scale M ≫ H , it is reasonable
to suppose that in fact inflation ends altogether, giving way to an epoch when the energy density
is dominated by the spacetime gradients of the fields. What happens next is associated with the
question of defect production, to which we now turn.
Defect production in the vacuum dominated case
In contrast with the inflaton dominated case, defect production has not previously been considered
for the vacuum dominated regime. The following discussion assumes that the couplings are of order
unity, or to be more precise that they are not extremely small for in that case a qualitatively different
scenario could ensue. In particular, we assume that λ′ is not small enough to satisfy Eq. (3.12), so
that as argued above the phase transition marks the end of inflation.
Ideally one would like to follow the evolution of the fields using quantum field theory in curved
spacetime, as in the inflaton dominated case, and hence calculate explicitly the typical spacing and
other stochastic properties of the initial distribution of the defects. However, that calculation relies
crucially on the fact that the back-reaction of ψ on φ is negligible which one easily checks is not
the case in the vacuum dominated regime11 , and indeed an estimate using the techniques described
in Section 3.1 indicates rather that the back-reaction hits φ long before ψ has had a chance to roll
down. In the absence of this simplifying feature, it is not even possible to give a qualitative account
of the evolution, let alone follow it in detail. When φ is hit by the back-reaction from ψ , it will
acquire a spatial gradient of order Mψ , which will soon become much bigger than m. As a result of
11 It is important in this connection not to be misled by the evolution discussed in the previous subsection, which was
similar to that seen in the inflaton dominated case. That discussion was a purely hypothetical one, used to establish
a contradiction; the premise that slow-roll continues for one Hubble time leads to that evolution, which in turn leads
to the contradictory result that slow-roll does not continue for one Hubble time. One does not expect anything like
it to actually occur, and in particular there is no reason to suppose that ψ first rolls down to ψvac , after which φ falls
down due to the destabilizing action of ψ. (This ‘waterfall’ sequence of events was suggested in [5].)
16
(3.16)
the backreaction, φ will look more like a collection of interacting plane waves than a homogeneous
field, and cannot be said to ‘roll down’ to its true vacuum. Moreover, ψ will in general still be an
essentially quantum ob ject at this stage (i. e. the state of the system is not a superposition of states
in which it has an almost well defined value over an extended period of time), so φ will become one
as well.
In the absence of an explicit calculation one must rely on order of magnitude arguments, which
as we now see actually point to rather definite conclusions. The crucial point is that the φ and ψ
fields are coupled to each other, and also in general to the quark, lepton etc. fields. Since at least
the scale M is much bigger than H (and even m is not many orders of magnitude less), one expects
the fields to thermalise quickly on the Hubble timescale; reheating in the vacuum dominated case
will occur promptly at the end of inflation. The reheat temperature Treh is therefore given by the
familiar formula Treh = (30/π2g∗ )1/4 ρ1/4 , or
Treh = (30λ/4π2g∗)1/4M ,
where g∗ is the effective number of degrees of freedom at that temperature, presumably at least
of order 102 (e. g. in the minimal supersymmetric standard model g∗ = 229). Thus the reheat
temperature is of order M . The defects that have been formed find themselves effectively in a
thermal bath at a temperature Treh , and may be in thermal equilibrium. If so, then for a string
network, for example, the most likely configuration will be the one which maximises the allowed
density of states. For the case of cosmic strings, such a configuration, below the Hagedorn transition
consists of maximising the number of possible loops that can form, with the long strings being
exponentially suppressed [33]. Since this distribution is similiar to that found soon after a thermal
phase transition has produced strings, it is possible that the effect of such a rapid reheating could
lead to a configuration of defects much the same as if there had been a thermal phase transition.
However, we have not explicitly demonstrated this to be the case here; it would require a detailed
numerical simulation of the reheating to rigorously establish how the network behaves.
Local cosmic strings are perhaps the most interesting defect for cosmology. The primary motiva-
tion for employing them here would be so they could contribute as seeds for the observed large scale
structure and the anisotropies in the microwave background. However, if we do try to make use of
them in the context of this inflation model, we must be cautious; it is expected that strings would
have a very important influence on the cmb if they are to be massive enough to affect structure
formation [34]. Therefore we must reassess the estimates of the allowed model parameters deter-
mined in the previous section, for these were obtained assuming that the inflaton field alone was
responsible for the cmb anisotropies. With two sources (assumed uncorrelated) of anisotropies, the
contributions add in quadrature. As a benchmark figure we reduce the inflation contribution to 10%
of the total anisotropy, corresponding to dropping δH by √10. Numerical calculation shows that for
unit couplings, this only reduces Mmax from 2.4 × 10−3mPl to 1.3 × 10−3mPl .
Recent simulations of cosmic string networks has shown the importance of the small scale struc-
ture on the network. An important effect of this structure is to renormalise the string mass per
unit length µ, relating it to the original mass per unit length µ0 by µ ∼ 1.4µ0 [35]. Recalling that
µ0 ≃ πM 2 (provided scalar and vector masses are not too disparate), we are therefore allowed a
mass per unit length of up to 9 × 10−6mPl , which is comfortably high enough to allow µ to fall in the
favoured range of values for structure formation µ ∼ 2 − 4 × 10−6m2
Pl [36]. An equivalent calculation
indicates that global textures can be similarly reconciled, though more marginally. Indeed, for the
highest values we can obtain, the strings would create excessive microwave fluctuations; the best
current bound on µ arises from cmb anisotropies on scales less than 10’, and yields µ < 3 × 10−6m2
Pl
[37]. Defect production can therefore provide an additional constraint on the viable parameters of
the inflationary theory.
Finally, recall that the above is with the unfavourable assumption of unit couplings; the analysis
of Section 2 indicates that the upper limit on M is yet higher if the couplings are reduced, which to
lowest order does not alter the string tension.
17
3.3 Non-thermal monopole production
We now consider gauge monopole production in a non-thermal phase transition. Some aspects of such
production have already been disussed in [38]. The case of gauge monopoles is particularly interesting
because they do not reach a scaling regime. Let the initial correlation length be some fraction ζ of
the Hubble radius, ξ ≡ ζH −1 . For the thermal case, it is easy to show that ζth ∼ g 1/2
∗ M /λmPl . In
the vacuum dominated case all we can be sure of (for fast roll-down) is that ζvac ≤ 1. Of course the
uncertainties in the initial distribution will be reflected in our lack of knowledge of the form ζvac
should take.
Now in general we can write the initial number density of monopoles and temperature as
1
ζ 3
H 3
T 3
i
.
(3.17)
1
ni
i ∼
i ∼
ξ 3 T 3
T 3
Once we know T we can determine H , and hence the future evolution of n/T 3 for a given value of ζ .
We have demonstrated that in the false vacuum dominated case, reheating is prompt, leading to a
temperature after inflation given by Eq. (3.16). Now this means that in Eq. (3.17), for a given reheat
temperature and assuming ζ ≤ 1, we obtain a similar scenario to the thermal case [41]. Neglecting
the effects of annihilation, which should be valid for monopole masses, Mmon ≤ 1017GeV and ζ ∼ 1,
we obtain
1014 GeV (cid:19)3 (cid:18) Mmon
1011
1016 GeV (cid:19) ,
ζ 3 (cid:18) Treh
Ωmonh2 ≃
where h is the Hubble parameter today in units of 100kms−1Mpc−1 and 0.4 < h < 1 [41].
In
other words, we are unable to differentiate between the thermal production of monopoles and this
particular non-thermal case, for a given initial temperature. The original GUT scale monopole
problem still exists in the non-thermal case.
Demanding Ωmonh2 < 1 constrains us to a region Mmon < 1013GeV. This is not the strongest
bound though, especially for the case of light monopoles. The Parker limit [40] (see [41] for details),
places a constraint on the allowed flux in monopoles of mass below 1017GeV. For consistency we
find that Mmon < 1012 GeV, slightly tighter than the density bound. Thus it appears that at least
under the assumption that ζ ∼ 1, the monopole problem is still very much present in the vacuum
dominated region.
(3.18)
4 Particle Physics Models
No matter how simple it might be, and no matter how well its predictions agree with observation, no
model of inflation can be regarded as satisfactory unless it emerges from a sensible theory of particle
physics. In the present context (Eq. (2.1)) this means that we want to identify φ and ψ with fields
belonging to such a theory, and to show that φ can have a sufficiently flat potential without fine
tuning, in particular to show that the mass m of the φ field can be sufficiently small (m ≪ H ) and
that there is no φ4 term.
We start by considering the case of global supersymmetry [42]. Here it is natural to focus on
the regime 100 GeV to 1 TeV for m, which is the smallest one commonly considered for scalar fields.
The requirement of having no φ4 term means that one cannot identify φ with a Higgs field, but
it might be one of the scalar fields suggested by supersymmetric theories. With m in this range,
the COBE normalisation requires that M is of order 1011 GeV which, as noted earlier [6], suggests
the possibility that the false vacuum is that of Peccei-Quinn symmetry. One is therefore led to ask
whether, by considering Peccei-Quinn symmetry in the context of supersymmetry, there emerges a
field φ with a quadratic coupling to the Peccei-Quinn field and a mass of order 1 TeV, but no φ4
coupling. For global supersymmetry the answer to our question is remarkable; the very first model of
supersymmetric Peccei-Quinn symmetry, proposed by Kim in 1984 [43], indeed has a suitable field φ.
18
If we could stop at this point, we would have fulfilled the wildest dreams of particle cosmologists. A
model motivated purely by particle physics would subsequently be seen (in this case nine years later)
to lead to an observationally viable epoch of inflation without any fine tuning of its parameters!
Unfortunately, there are sound particle physics reasons for rejecting global supersymmetry and
replacing it by supergravity [42]. In a general supergravity theory it is difficult to construct a model
of inflation without fine tuning, because in addition to the global supersymmetric type terms there
is an infinite series of higher order non-renormalisable terms, the first of which usually gives any
would-be inflaton an effective mass of order H .
However, supergravity can only be regarded as an effective theory with a cutoff at the Planck
scale. So in order to get a better handle on the crucial non-renormalisable terms we should consider
a theory of everything. Superstrings provide one possible candidate theory. Here we find another
superstring miracle. For a class of low energy effective supergravity theories derived from super-
strings, they are of precisely the form necessary to cancel the harmful non-renormalisable terms. We
go on to make the first steps towards constructing a truely realistic, superstring derived, model of
inflation.
4.1 A simple supersymmetric model
The simplest superpotential [42] that spontaneously breaks a U(1) symmetry is
W = σ(Ψ1Ψ2 + Λ2 )Φ ,
(4.1)
2
2
2
2
(4.2)
where Φ, Ψ1 and Ψ2 are chiral superfields which we take to have canonical kinetic terms, Λ is a
mass which sets the scale of the spontaneous symmetry breaking, σ is a coupling constant, and the
U(1) symmetry is Ψ1 → eiθΨ1 , Ψ2 → e−iθΨ2 . This superpotential is often used in supersymmetric
model building [42, 43], and in particular was used by Kim [43] to construct the first supersymmetric
realisation of Peccei-Quinn symmetry. We shall now show that for fairly generic initial conditions,
it leads to the false vacuum inflation model of the previous sections with the identifications λ′ =
2λ = σ2/2 and M = 2Λ.
The scalar potential derived from this superpotential is
+ (cid:12)(cid:12)(cid:12)(cid:12)
∂Ψ1 (cid:12)(cid:12)(cid:12)(cid:12)
∂Ψ2 (cid:12)(cid:12)(cid:12)(cid:12)
+ (cid:12)(cid:12)(cid:12)(cid:12)
∂Φ (cid:12)(cid:12)(cid:12)(cid:12)
V = (cid:12)(cid:12)(cid:12)(cid:12)
∂W
∂W
∂W
,
2
+ σ2 (cid:0)Ψ1 2 + Ψ2 2 (cid:1) Φ2 ,
= σ2 (cid:12)(cid:12)Ψ1Ψ2 + Λ2 (cid:12)(cid:12)
where Φ, Ψ1 and Ψ2 now represent just the (complex) scalar component of the respective chiral
superfields. Adding a soft supersymmetry breaking mass, m, of order 1 TeV, for Φ, we obtain
+ σ2 (cid:0)Ψ1 2 + Ψ2 2 (cid:1) Φ2 + m2 Φ2 .
V = σ2 (cid:12)(cid:12)Ψ1Ψ2 + Λ2 (cid:12)(cid:12)
We want to show that Φ can be the inflaton, and so to obtain the effective potential during
inflation we will minimise this potential for fixed Φ. The potential is minimised at arg Ψ1 + arg Ψ2 =
π , and the canonically normalised field corresponding to the phase of the Ψ field with smaller
magnitude has an effective mass ≥ σΛ there. At least if σ is not very small this will anchor
arg Ψ1 + arg Ψ2 at the value π . The potential is independent of the other two angular degrees of
freedom, namely arg Ψ1 − arg Ψ2 (which corresponds to the axion field) and arg Φ, and Hubble
damping will make them practically time independent more or less independently of the initial
conditions. This leaves only the radial degrees of freedom, corresponding to the three canonically
normalised real fields φ = √2 Φ, ψ1 = √2 Ψ1 and ψ2 = √2 Ψ2 . In terms of them, the potential
is
σ2
σ2
1
4 (cid:0)ψ1ψ2 − 2Λ2(cid:1)2
2 (cid:1) φ2 +
4 (cid:0)ψ2
1 + ψ2
m2φ2 .
V (φ, ψ1 , ψ2 ) =
+
(4.5)
2
19
(4.4)
(4.3)
For φ > 0, the degree of freedom orthogonal to ψ1ψ2 (which corresponds to the saxino) has its
minimum at ψ1 = ψ2 , and it is straightforward to show that the effective mass of the canonically
normalised saxino field is everywhere ≥ σφ/√2 and so the saxino will be firmly fixed at its minimum
during inflation12 . Then in terms of φ and the canonically normalised field ψ = √2ψ1ψ2 , the
potential is
σ2
σ2
1
16 (cid:0)ψ2 − 4Λ2 (cid:1)2
2
4
Thus we have the model of Sections 2 and 3, with λ′ = 2λ = σ2/2 and M = 2Λ. From Eq. (2.41) it
follows that the usual axion parameter fa is given by
V (φ, ψ) =
φ2ψ2 +
m2φ2 .
(4.6)
+
2
5
,
5 (cid:16) m
TeV (cid:17)
fa = 2Λ = 8 × 1011GeV σ− 3
which is at the right scale for the axion.
Note that for chaotic initial conditions Λ and m will initially be negligible and so the potential
Eq. (4.4) will initially have the simple form
V = σ2 (cid:16)Ψ1Ψ2 2 + Ψ2Φ2 + ΦΨ1 2(cid:17) .
Thus if initially Φ > Ψ1 , Ψ2 , i.e. one third of the initial condition space, the fields will rapidly
approach the inflating tra jectory Ψ1 = Ψ2 = 0 and Φ ≫ Λ given above.
(4.7)
(4.8)
4.2 Supergravity
+
+
The scalar potential in supergravity [42] has the general form
K (cid:19)
∂ ¯φα ∂φβ (cid:19)−1 (cid:18) ∂W
∂φα (cid:19) (cid:18) ∂ ¯W
Xα,β (cid:18) ∂ 2K
V = exp (cid:18) 8π
∂K
∂ ¯φβ
m2
∂φα
Pl
+ D − term ,
where the Kahler potential K (φ, ¯φ) is a real function of the complex scalar fields φα and their
hermitian conjugates ¯φα , and the superpotential W (φ) is an analytic function of φ. The D-term is
quartic in the charged fields, and we will assume that it is flat along the inflationary tra jectory so
that it can be ignored during inflation. It may however play a vital role in determining the tra jectory
and in stabilising the non-inflaton fields. The term given explicitly is called the F-term.
The kinetic terms are
∂ ¯φβ (cid:19) − 3
∂K
Pl W 2
8π
m2
(4.9)
8π
m2
Pl
8π
m2
Pl
¯W
W
∂ 2K
∂φα∂ ¯φβ
(4.10)
∂µφα∂ µ ¯φβ ,
Xα,β
where µ is a spacetime index. It follows that for canonically normalised fields
φα 2 + . . . ,
K = Xα
where . . . stand for higher order terms. Global supersymmetry corresponds to the case where these
terms are absent, and one has taken the limit mPl → ∞ to obtain the potential V = Pα ∂W/∂φα 2
that we used earlier. Supergravity corresponds to keeping mPl finite. Then the F-term part of the
scalar potential becomes
V = exp 8π
φγ 2 + . . .! ×
Pl Xγ
m2
12 φ > φinst = √2 Λ during inflation (see Eq. (4.6) and Sections 2 and 3).
(4.12)
(4.11)
20
+
.
+
2
,
(4.14)
(4.13)
8π
m2
Pl
Pl W 2
Pl (cid:0) ¯φα + . . .(cid:1) W (cid:21) (cid:20) ∂ ¯W
(δαβ + . . .) (cid:20) ∂W
(φβ + . . .) ¯W (cid:21) − 3
8π
8π
Xα,β
∂ ¯φβ
m2
m2
∂φα
Thus, for any model of inflation, the lowest order (i.e. global supersymmetric) inflationary potential
Vglobal ≡ Xα (cid:12)(cid:12)(cid:12)(cid:12)
∂φα (cid:12)(cid:12)(cid:12)(cid:12)
∂W
will receive corrections13 giving
φα 2 + other terms! + other terms .
V = Vglobal 1 +
8π
Pl Xα
m2
The φα 2 term in this equation gives a contribution 8πVglobal /m2
Pl ≃ 3H 2 to the effective mass
squared of al l scalar fields, therefore, assuming the inflaton is the modulus of a scalar field14 , it gives
a contribution of order unity to η ≡ m2
PlV ′′ /8πV (see Section 2.2). But η ≪ 1 is necessary for
inflation to work (at least in the usual slow-roll form). As a result practically all15 of the supergravity
models of inflation proposed so far [45] have involved unmotivated fine tuning of the Kahler potential
and/or the superpotential in order to cancel the harmful non-renormalisable corrections (i.e. to get
the ‘other terms’ in Eq. (4.14) to cancel the φα 2 term). However, supergravity can only be regarded
as an effective theory with a cutoff at the Planck scale. So in order to get a better handle on the
crucial non-renormalisable terms we should consider a theory of everything. Superstrings provide
the most promising candidate, and in the next section we will find that for the Kahler potential
derived from orbifold compactification of superstrings the cancellation can occur without any fine
tuning.
To end this section we just note that for the special choice of supergravity with minimal kinetic
terms16 (i.e. Eq. (4.11) and Eq. (4.12) without the higher order corrections . . . ) and the superpo-
tential of the previous section, the ‘other terms’ in Eq. (4.14) cancel the φα 2 term for the inflaton,
as we will now show.
Substituting K = Φ2 + Ψ1 2 + Ψ2 2 and the superpotential of the previous section, Eq. (4.1),
into Eq. (4.9) gives
Pl Φ2(cid:19) ×
V (Φ, Ψ1 , Ψ2) ≃ σ2 exp (cid:18) 8π
(4.15)
m2
(8π)2
Pl Φ4(cid:19) + (cid:0)Ψ1 2 + Ψ2 2 (cid:1) Φ2 (cid:21) .
2 (cid:18)1 −
(cid:20)(cid:12)(cid:12)Ψ1Ψ2 − Λ2 (cid:12)(cid:12)
8π
Pl Φ2 +
m2
m4
Minimising with respect to Ψ1 and Ψ2 for Φ > Λ as in the previous section, gives
(8π)2
V (φ) = σ2Λ4 exp (cid:18) 1
φ4(cid:19) ,
φ2(cid:19) (cid:18)1 −
8π
8π
1
1
φ2 +
m2
m4
m2
2
2
4
Pl
Pl
Pl
(8π)2
= σ2Λ4 (cid:18)1 +
φ4 + . . .(cid:19) .
1
m4
8
Pl
Thus the problematic mass term cancels out. However we do not regard this as a realistic model
and so will not pursue it further.
> mPl /√8π then a
13 We assume that the corrections are small as will be the case if φα ≪ mPl /√8π . If φα ∼
glance at the exponential factor in Eq. (4.12) shows that the problems will then be even more severe.
14 Natural inflation [44] avoids this problem because its inflaton is the phase of a complex scalar field.
15 The only exception known to us is ‘natural inflation’ [44], as mentioned above.
16 Although this is in some sense the simplest supergravity theory, it is not well motivated physically and its
adoption must be regarded as fine tuning to some extent.
(4.16)
21
4.3 Superstrings
The inflation that inflated the observable universe beyond the Hubble radius, and could have pro-
duced the seed inhomogeneities necessary for galaxy formation and the anisotropies recently observed
by COBE, must occur at an energy scale V 1/4 ≤ 4 × 1016GeV [23], well below the Planck scale. At
these relatively low energies, superstrings are described by an effective N=1 supergravity theory [42].
The properties of that supergravity theory are known in most detail for orbifold compactification
schemes, and so we will restrict ourselves to such compactifications, although our results may be
more general. Also, for simplicity, we will ignore the twisted sector of the theory. For the remainder
of this section we set mPl /√8π = 1.
Following [46], we will assume the following form for the one-loop corrected Kahler potential K
of the supergravity theory derived from orbifold compactification of superstrings [47, 48, 46, 49]
ln Xi ,
(4.17)
(4.18)
(4.19)
with
and
Y = S + ¯S +
δGS
i
ln Xi ,
K = − ln Y −
3
Xi=1
3
Xi=1
i 2 ,
Xi = Ti + ¯Ti − Xα
φα
where S is the dilaton whose real part gives the tree-level gauge coupling constant (Re S ∼ g−2
gut ), Ti
are untwisted moduli whose real parts give the radii of the three compact complex dimensions of the
orbifold, and φα
i are the untwisted matter fields associated with Ti . The terms with coefficients δGS
i
are one-loop corrections coming from the Green-Schwarz mechanism, whose matter field dependence
is speculative [46] (note that our convention for the sign and magnitude of these coefficients follows
[48, 49], not [46]).
For initial orientation we make the standard assumption that the dilaton and moduli have ex-
pectation values of order one, showing later how this may be achieved. (In our vacuum, hRe S i ∼ 2
and hRe Ti i ∼ 1 [49], but as we note later these values may be different during inflation.) We will
also make the standard assumption that the matter fields have expectation values much less than
one. The fact that this can and does include the inflaton is an important advantage of this model
of inflation. The values of the dimensionless coefficients δGS
depend on the orbifold assumed. Some
i
values for δGS
that have been calculated in [48] are 0, 5 and 15. We will assume δGS
i ≥ 0 as is the
i
case for all orbifolds considered up to now [48, 50].
The F-term part of the scalar potential corresponding to this Kahler potential is [46]
i=1 Xi ((cid:12)(cid:12)(cid:12)(cid:12)
2
∂S (cid:12)(cid:12)(cid:12)(cid:12)
1
∂W
W − Y
Y Q3
3
i (cid:12)(cid:12)(cid:12)(cid:12)
∂ Ti (cid:12)(cid:12)(cid:12)(cid:12)
∂ Ti (cid:12)(cid:12)(cid:12)(cid:12)
+ Xi Xα (cid:12)(cid:12)(cid:12)(cid:12)
Y
1
∂W
∂W
Xi=1
4π2 δGS
W +
i
Y + 1
4π2 δGS
To lowest order in δGS and the matter fields, the kinetic terms given by Eq. (4.10) are
1
1
1
i ∂ µ ¯φα
(cid:0)Ti + ¯Ti(cid:1)2 ∂µTi∂ µ ¯Ti + Xi,α
(cid:0)S + ¯S (cid:1)2 ∂µS ∂ µ ¯S + Xi
∂µφα
i .
Ti + ¯Ti
(4.20)
2!) .
∂W
∂S − Xi
∂W
∂φα
i
+ ¯φα
i
V =
+
− 3W 2
2
1
4π2
22
(4.21)
(4.22)
The superpotential W is composed of a perturbative part, Wpert(φ, T ), and a non-perturbative part,
Wnp (φ, S, T ). To lowest order in the matter fields, Wpert has the general form [47, 46, 49]
Wpert = Xα,β ,γ
where wαβγ = 0 or 1. Wnp is not very well understood. However, it should have an expansion
in powers of eS to reflect its nonperturbative nature [51, 49]. Also, orbifold compactifications of
superstrings are invariant under target-space duality symmetries to all orders of string perturbation
theory and, it is thought, nonperturbatively as well [52, 51, 53]. These duality transformations act
on the moduli as
2 φγ
1 φβ
wαβγ φα
3 ,
aiTi − ibi
aidi − bici = 1 .
Ti →
iciTi + di
The parameters ai , bi , ci , di are in general a discrete set of real numbers. In many cases the duality
group is given by the product of three modular groups, i.e. ai , bi , ci , di ∈ Z, and for simplicity we will
i transform in the same way as 1/[η(Ti)]2
assume this to be the case here. Then the matter fields φα
where
∞
Yn=1 (cid:0)1 − e−2nπTi (cid:1)
is the Dedekind function. It will also be useful to define the modular-invariant dilaton field [48, 46]
η(Ti ) = e−
(4.23)
(4.24)
πTi
12
,
δGS
i
(4.25)
1
4π2
ln [η (Ti )]2
S ′ = S −
3
Xi=1
from which the transformation properties of the dilaton can be deduced. Requiring modular invari-
ance then puts strong constraints on the form of the low-energy supergravity theory and in particular
on Wnp [52, 51, 46, 49].
As we shall see, the Kahler potential of Eq. (4.17) has some very special properties as far as
inflation is concerned. The crucial point is the cancellation of the Xi factor in front of the global
supersymmetric ∂W/∂φα
i term in Eq. (4.20). This has the consequence that if the ∂W/∂φα
i terms for
one value of i, say i = 3, dominate the inflationary potential energy then the canonically normalised
3 fields do not acquire corrections17 of order H to their effective masses as would be expected
T3 and φα
in supergravity in general (see the previous section). This opens up a path to inflation without fine
tuning. Note that the above conditions for inflation are asymmetric in the Ti which means that the
Kahler potential K = − ln (cid:0)S + ¯S (cid:1) − 3 ln (cid:16)T + ¯T − Pα φα 2(cid:17) [55] cannot be regarded as equivalent
to the Kahler potential of Eq. (4.17) and does not share its inflationary properties. We will now go
on to chart this path to inflation for the case of false vacuum inflation.18
i ≪ 1, it is not unreasonable to assume that the ∂W/∂φα
Since we are assuming φα
i terms
dominate the potential during inflation.19 Then
Pα (cid:12)(cid:12)(cid:12)
i (cid:12)(cid:12)(cid:12)
∂W
3
∂φα
Xi=1
(cid:0)Y + 1
4π2 δGS
i (cid:1) Qj 6=i Xj
Next we minimise the potential, with respect to the matter field dependence of the ∂W/∂φα
i
terms,
(4.26)
V =
2
.
17An exception to this statement is the case of a φα
3 field whose ∂W/∂φα
3 terms contribute to the inflationary
potential energy and are likely to pick up masses of order H from the W 2 terms in Eq. (4.20).
18 One of us will consider the case of true vacuum inflation elsewhere [54].
19 We will consider what effect the other terms might have on the inflaton’s potential later.
23
2
,
For a fixed inflationary value (i.e. φ > φinst , see Section 2) of the inflaton (whatever it may turn
out to be), the matter fields have masses much greater than S , so they will settle to their values on the
inflationary tra jectory before S moves significantly. We assume that the ∂W/∂φα
i terms will then be
independent of the inflaton, as will be the case for false vacuum inflation. We also assume that some
of them will be non-zero. Dropping temporarily the superscripts on the matter fields, W transforms
under modular transformations like φ1φ2φ3 , and therefore ∂W/∂φi transforms like Qj 6=i φj which
we noted earlier transforms like Qj 6=i [1/η(Tj )]2 . Since the S dependence arises entirely from non-
perturbative effects, one therefore expects the following functional form [52, 51, 46, 49]
in (T )e−bnS ′
= Pn aα
∂W
(4.27)
,
∂φα
Qj 6=i [η (Tj )]2
i
where the aα
in ’s will in general be arbitrary modular invariant functions of the Ti , but for the most
part we will assume that they are constants as is the case in gaugino condensation scenarios [49].
Also, we will assume that the bn ’s are positive as in such scenarios.
Now, as will soon become clear, in order to get inflation we need the false vacuum energy density,
V0 , to be dominated by one or more ∂W/∂φα
i terms with the same value of i, say i = 3. For the
moment let us ignore the other i values altogether. Then the potential during inflation is given by
3 (cid:12)(cid:12)(cid:12)
Pα (cid:12)(cid:12)(cid:12)
∂W
∂φα
Vinfl =
(cid:0)Y + 1
4π2 δGS
3 (cid:1) Qi6=3 Xi
2
3n e−bnS ′ (cid:12)(cid:12)(cid:12)
= Pα (cid:12)(cid:12)(cid:12)Pn aα
AB
ln
η (Ti )4
2
3
1
i (cid:12)(cid:12)(cid:12)
Ti + ¯Ti − Xβ (cid:12)(cid:12)(cid:12)
φβ
Xi=1
4π2
2
i (cid:12)(cid:12)(cid:12)
η (Ti )4 .
φβ
2
T3 + ¯T3 − ln h(cid:0)T3 + ¯T3 (cid:1) η (T3 )4 i
3 (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
φβ
δGS
3
4π2 (cid:0)S + ¯S (cid:1)
2
Pα (cid:12)(cid:12)(cid:12)Pn aα
3n e−bn S ′ (cid:12)(cid:12)(cid:12)
(cid:0)S ′ + ¯S ′ (cid:1) Qi6=3 (cid:0)Ti + ¯Ti(cid:1) η (Ti )4 .
As pointed out by Brustein and Steinhardt [56] and Carlos et al [57], the dilaton provides the
biggest obstacle to constructing a model of inflation in superstrings.20 Our model helps with the
difficulty pointed out by Brustein and Steinhardt, because V0 can give S ′ a suitable minimum during
inflation in much the same way as double gaugino condensation scenarios do in the true vacuum
[49]. Since we are supposing that the bn ’s are positive, we need for this purpose at least two distinct
1
A ≡ S ′ + ¯S ′ +
4π2 δGS
3 +
Ti + ¯Ti − Xβ (cid:12)(cid:12)(cid:12)
B ≡ Yi6=3
This may be written in the form
2
V = V0
i (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
φβ
1 + Xi6=3 Xβ
Ti + ¯Ti
where
Xβ
,
+ . . .
(4.30)
(4.28)
(4.29)
+
V0 =
where
δGS
i
(4.31)
20At least if we don’t assume something like S -duality [58], which allows some of the bn to be negative.
24
values of n for some α so as to obtain a minimum at a finite value of S ′ , and then at least one more
term with a different value of α to make V0 nonzero. A minimum with V0 > 0 and mass greater
than H can then be obtained for reasonable, but significantly constrained, values of the a’s and b’s.
There is, though, still the problem pointed out by Brustein and Steinhardt that for a potential of
the form of V0 , and for generic initial values, S will tend to roll past the desired minimum and on
to the minimum at S = ∞. As this is also a problem for the true vacuum it should be regarded
as a problem for the assumption of all positive bn ’s rather than of the model of inflation. It might
be solved by anthropic arguments, which in any case seem likely to be needed because of the huge
degeneracy in the superstring vacuum.
There remains the problem that the V0 -induced expectation value for S ′ is likely to be different
from its vacuum expectation value after inflation because V0 disappears at the end of inflation. As
pointed out by Carlos et al [57], this might lead to cosmological problems because S ′ will in general
be left far from its minimum at the end of inflation. We do not address that difficulty here.
Next consider the moduli Ti for i 6= 3. The function (Ti + ¯Ti )η(Ti )4 has its maxima at Ti = eiπ/6
and points equivalent under modular transformations, and
= √3 (cid:12)(cid:12)(cid:12)
η (cid:16)eiπ/6(cid:17)(cid:12)(cid:12)(cid:12)
(Ti + ¯Ti )η(Ti )4 (cid:12)(cid:12)Ti=eiπ/6+√3 ti
(cid:2)1 − ti 2 + O (cid:0)t3
i (cid:1)(cid:3) ,
(4.32)
where (cid:12)(cid:12)η (cid:0)eiπ/6 (cid:1)(cid:12)(cid:12) ≃ 0.8006. Therefore V0 is minimised for Ti = eiπ/6 , i 6= 3, and since, to lowest order
in δGS and the matter fields, the canonically normalised Ti fields (see Eq. (4.21)) have masses √V0
there, they will be firmly anchored during inflation. However, after inflation they will be left sitting
far from their true vacuum minima (which are close to T=1.23 in some models [49]) potentially
giving cosmological problems [57]. Note that if the aα
3n ’s were functions of the Ti for i 6= 3, then
they would merely shift the Ti ’s expectation values during inflation, whilst if they depended on T3
they would fix T3 during inflation, simplifying the following discussion.
Next, the canonically normalised φβ
i matter fields (see Eq. (4.21)), for i 6= 3, acquire masses
√V0 ≃ √3 H (see Eq. (4.30)), making them unviable as inflatons and firmly fixing them during
inflation.
Having argued for the stability of V0 during inflation, let us consider the possible inflatons. The
canonically normalised φβ
3 fields acquire a mass squared (to lowest order in δGS and the matter
fields)
4
,
(4.33)
m2
GS =
δGS
3 V0
4π2 hS + ¯S i
much less than V0 (assuming that we are in the perturbative regime so that the loop corrections
are small). Note that here hS + ¯S i is the expectation value of the dilaton during inflation, which is
probably different from its value in our vacuum. However, it may be reasonable to assume that it
has a similar value during inflation as now because, in both cases, we need to be in the perturbative
> 1, but avoid the potentially runaway behaviour at large Re S [56]. Also, Vinfl is
regime, Re S ∼
minimised for T3 = eiπ/6 . Defining T3 = eiπ/6 + √3 t3 , and assuming t3 ≪ 1 (and (cid:12)(cid:12)(cid:12)
3 (cid:12)(cid:12)(cid:12) ≪ 1, which
φβ
we have been assuming all along), the φβ
3 and T3 dependence of the inflationary potential energy is
given by
Vinfl = V0
4π2 hS + ¯S i
+ t3 2
+ . . .
δGS
2
Xβ (cid:12)(cid:12)(cid:12)
3 (cid:12)(cid:12)(cid:12)
Φβ
3
1 +
(4.34)
,
3 and t3 are the canonically normalised φβ
where Φβ
3 and T3 fields. Then defining the canonically
normalised (inflaton) field φ = √2 rPβ (cid:12)(cid:12)(cid:12)
2
3 (cid:12)(cid:12)(cid:12)
+ t3 2 , and making the reasonable assumption that
Φβ
25
1
2
,
(4.36)
(4.35)
n = 1 +
the orthogonal degrees of freedom are time independent, we get the potential during inflation
δGS
φ2(cid:21) .
Vinfl = V0 (cid:20)1 +
3
4π2 hS + ¯S i
Therefore, from Section 2.5, the density perturbations produced during inflation will have a spectral
index
δGS
3
2π2 hS + ¯S i
directly related to fundamental superstring parameters. For example, taking hS + ¯S i = 4 [49] and
δGS
3 = 5 [48] gives n = 1.06.
3 = 0 [48], T3 and φβ
For δGS
3 receive no contribution to their potential from Vinfl and so either
the terms in Eq. (4.26) neglected in Eq. (4.28) or the terms in Eq. (4.20) neglected in Eq. (4.26)
will dominate. In the first case, if the ∂W/∂φα
i terms for i 6= 3 are non-zero but still much smaller
than the i = 3 terms then they could provide T3 and the φβ
3 with masses ≪ H in the same way as
the i = 3 terms provide the Ti and φβ
i for i 6= 3 with masses of order H . Note that if the ∂W/∂φα
i
terms for i 6= 3 are of the same order as the i = 3 terms then all the fields will have masses of
order H and inflation will not be possible. In the latter case, the neglected terms are the terms
that are thought to provide the soft supersymmetry breaking terms in our vacuum and so, during
inflation, they might also provide the φβ
3 with soft supersymmetry breaking mass terms and give T3
a minimum with a mass of the same order. However, as the expectation value of the dilaton during
inflation is likely to be different from its value in our vacuum, the soft supersymmetry breaking scale
is also likely to be different. Thus for δGS
3 = 0, the results will depend on the specific superpotential.
> 1. Then Eq. (4.30) will no longer be a
Finally let us consider briefly the possibility that φ ∼
good approximation. For example, consider the simplest case of constant T3 and, without loss of
generality, one φβ
3 field which we will call φ3 . Then, to lowest order in δGS and the other matter
fields, φ3 ’s kinetic term is
(4.38)
(4.37)
∂ 2K
∂φ3∂ ¯φ3 ∂µφ3 2 =
T3 + ¯T3
(cid:16)T3 + ¯T3 − φ3 2(cid:17)2 ∂µφ3 2 .
Therefore the canonically normalised real field φ corresponding to φ3 is given by
φ
φ3 = pT3 + ¯T3 tanh
√2
.
Note that φ ∼> 1 corresponds to φ3 ∼ pT3 + ¯T3 but still φ3 < pT3 + ¯T3 . This is the only place
where we will relax the assumption of φ3 ≪ pT3 + ¯T3 . One of the problems of relaxing this
assumption is that we neglected the terms in Eq. (4.20) proportional to 1/X3 = cosh2 (φ/√2)/(T3 +
¯T3 ). 21 This will be reasonable for φ ≪ 1 but is unlikely to be so for φ ≫ 1. However it may just
be acceptable for φ ∼ 1, and so, making the big assumption that the derivation of Eq. (4.28) is still
valid for φ ∼> 1, we get the inflationary potential
δGS
Vinfl = V0 (cid:20)1 + 2
√2 (cid:21) .
φ
3
4π2 hS + ¯S i
3 = 15 [48] and φinst ∼ V 1/4
For example, taking hS + ¯S i = 4 [49], δGS
could give an observable
0
signature of superstrings in the varying spectral index of the density perturbations produced during
inflation. Note that we must be in the vacuum dominated regime for the loop expansion (expansion
in δGS /4π2 hS + ¯S i) to be reliable.
21Note that these terms, although not inflationary, could lead to a scaling a ∝ t of the scale factor of the universe,
taking us down from the Planck scale to the energy scale at which inflation proper starts.
ln cosh
(4.39)
26
A specific model
The arguments that we have given suggest that false vacuum inflation can be achieved, provided
that the superpotential satisfies certain conditions. We have not however demonstrated that such
a superpotential exists, nor have we discussed the instability mechanism which ends inflation (see
Sections 2 and 3). We therefore end by showing how things work out with a specific choice for the
perturbative part of the superpotential. The form we choose is
2 (cid:17) φ3 + (cid:16) φ(1)
Wpert = (cid:16) φ(1)
φ(3)
2 + φ(2)
φ(1)
1
1
1
+ (cid:16) φ → φ(cid:17) ,
where , , and (α) correspond to the generic α label used previously. We also assume that φ(α)
i
and φ(α)
for i = 1, 2 and α = 2, 3 acquire the expectation values
i
2 (cid:17) φ3
φ(1)
2 + φ(4)
φ(4)
1
2 + φ(3)
φ(2)
1
(4.40)
i = Λ(α)
φ(α)
i
(S, T ) =
i e−b(α)
a(α)
i S ′
[η (Ti )]2
,
(4.41)
V =
2
Λ(2)
2
2
φ(1)
2
φ3 + φ(4)
2
φ3 (cid:12)(cid:12)(cid:12)
+ ( → )(cid:21)
2
φ3 (cid:12)(cid:12)(cid:12)
+ (cid:12)(cid:12)(cid:12)
+ ( → )
and similarly for φ. The form of these expectation values is motivated by gaugino condensation
scenarios [49]. These expectation values might be induced by the D-term part of the scalar potential
or, more directly, by a nonperturbative part of the superpotential. We will assume φ3 is D-flat
because it will become (part of ) the inflaton.
Substituting into Eq. (4.26) we get
1 (cid:1) (cid:20)(cid:12)(cid:12)(cid:12)
2
1
φ3 (cid:12)(cid:12)(cid:12)
+ (cid:12)(cid:12)(cid:12)
φ3 (cid:12)(cid:12)(cid:12)
+ (cid:12)(cid:12)(cid:12)
φ(1)
Λ(3)
2
2
X2X3 (cid:0)Y + 1
4π2 δGS
+ (subscript 1 ↔ subscript 2)
3 (cid:1) (cid:20)(cid:12)(cid:12)(cid:12)
2
1
2 (cid:12)(cid:12)(cid:12)
Λ(3)
2 + Λ(3)
Λ(2)
2 + Λ(2)
φ(1)
φ(1)
+
1
1
1
X1X2 (cid:0)Y + 1
4π2 δGS
2(cid:21) .
+ (cid:12)(cid:12)(cid:12)
2 + (cid:16) φ → φ(cid:17)(cid:12)(cid:12)(cid:12)
φ(1)
2 + φ(4)
φ(4)
φ(1)
1
1
For φ3 > 0 this potential is minimised for φ3 = φ3 = φ(4)
2 = φ(4)
1 = φ(4)
1 = φ(4)
2 = 0. Then
2(cid:19) (cid:12)(cid:12)(cid:12)
V = (cid:18)(cid:12)(cid:12)(cid:12)
2
2
2 (cid:12)(cid:12)(cid:12)
φ3 (cid:12)(cid:12)(cid:12)
2 (cid:12)(cid:12)(cid:12)
+ (cid:12)(cid:12)(cid:12)
φ(1)
φ(1)
+ (subscript 1 ↔ subscript 2)
X2X3 (cid:0)Y + 1
4π2 δGS
1 (cid:1)
2
2 (cid:12)(cid:12)(cid:12)
+ (cid:12)(cid:12)(cid:12)
Λ(3)
Λ(2)
2 + Λ(3)
φ(1)
2 + Λ(2)
φ(1)
1
1
1
X1X2 (cid:0)Y + 1
4π2 δGS
3 (cid:1)
Λ(3)
Λ(2)
2 + Λ(3)
Defining the canonically normalised fields Φ ∝ φ3 , Ψ1 ∝ φ(1)
1 , Ψ2 ∝ φ(1)
2 , Λ2 ∝ Λ(2)
2 ,
1
1
and similarly for the checked symbols, then working to lowest order in δGS and the matter fields, we
obtain
S + ¯S (cid:20)(cid:12)(cid:12)(cid:12)
+ (cid:18)(cid:12)(cid:12)(cid:12)
2(cid:19) (cid:12)(cid:12)(cid:12)
+ ( → )(cid:21) .
2
1
Ψ1 Ψ2 + Λ2 (cid:12)(cid:12)(cid:12)
+ (cid:12)(cid:12)(cid:12)
Ψ1 (cid:12)(cid:12)(cid:12)
Φ(cid:12)(cid:12)(cid:12)
Ψ2 (cid:12)(cid:12)(cid:12)
Proceeding as in Section 4.1 then gives
S + ¯S (cid:20) 1
φ2 ψ2 + ( → )(cid:21) .
1
1
16 (cid:16) ψ2 − 4 Λ2(cid:17)2
+
V =
4
27
+ ( → )
2
2
.
(4.43)
(4.42)
(4.44)
(4.45)
V =
For φ > √2 max{ Λ, Λ}, this is minimised for ψ = ψ = 0, giving the false vacuum energy density
V = V0 = Λ4 + Λ4
(4.46)
.
S + ¯S
as discussed above. The higher order terms give the inflaton a potential, also as discussed above.
However, at φ = φinst = √2 max{ Λ, Λ} an instability sets in ending inflation in a manner similar to
that described in Sections 2 and 3.
5 A First-Order Model
The one context in which the dynamical effect of more than one scalar field during inflation has
been considered in some detail in the literature is in models of inflation ended by a first-order
phase transition, where a field must tunnel from the metastable false vacuum, through a classically
forbidden region, to the true vacuum. In the case of a single scalar field (Guth’s old inflation model
[59]) the metric rapidly reaches the static de Sitter metric with a fixed nucleation rate to the true
vacuum and the transition must either complete at once (without sufficient inflation) or not at all.
The critical parameter here is the percolation parameter, p, the average number of bubbles nucleated
per Hubble volume per Hubble time. To complete the transition p must exceed some critical value,
pcr = O(1) [60]. By introducing a second scalar field which can evolve with time, p can grow allowing
sufficient inflation before the transition completes.
To incorporate such a first-order transition into our model we must extend our basic potential,
Eq. (2.1), to include asymmetric terms which can break the degeneracy of the two vacuum states at
low energies. Thus we will consider the more general potential,
αM 2ψ2 −
λ (cid:0)M 4 + ψ4 (cid:1) +
The cubic term spoils the degeneracy, and choosing α greater or less than zero determines whether
ψ = 0, φ = 0 is a local minimum or saddle point respectively. Thus in addition to the two mass
scales M and m we now have four dimensionless coupling constants λ, λ′ , α and γ . Requiring the
energy density of the true vacuum to be zero (V (0, ψtrue ) = 0) can be used to specify γ , say, in
terms of α and λ. Thus we have one more free parameter, α, than in our second-order model (which
corresponds to the particular case α = −λ, γ = 0).
At large values of φ (where λ′ (φ/M )2 > (γ 2/4λ) − α) the potential has only one turning point
with respect to ψ , a minimum at ψ = 0, while for smaller values of φ a second minimum appears,
initially as a point of inflection at ψ = γ /2λ. Although it is this second minimum that develops into
the true vacuum with V = 0 when φ = 0, for γ 6= 0 it initially has an energy density greater than
that of the false vacuum so that if the fields follow the “path of least resistance” they will remain in
the false vacuum for α + λ′ (φ/M )2 > 0.
γM ψ3 +
V (φ, ψ) =
1
4
m2φ2 +
λ′φ2ψ2 .
(5.1)
1
2
1
3
1
2
1
2
5.1 Inflationary dynamics
While ψ is restricted to the false vacuum (ψ = 0) the potential for φ remains that given in Eq. (2.3),
and the dynamics are the same as considered in Section 2, except that the effective mass of the ψ
field is now
ψ = αM 2 + λ′φ2 ≡ α(φ)M 2 ,
M 2
and a second-order transition is not possible for α > 0.
Instead the transition must proceed by
nucleating bubbles of the true vacuum and the end-point φinst is replaced by the critical value φcr
where the percolation parameter reaches pcr .
Notice then that inflation ends at
(5.2)
φend = max{φǫ , φcr} ,
28
(5.3)
where the slow-roll condition may break down at φǫ (defined as in Section 2) before the true vacuum
percolates. If this is the case then we are again in the inflaton dominated limit, the false vacuum
energy density is negligible (λM 4 ≪ m2φ2
60 ) and the constraints are exactly the same as when the
eventual transition to true vacuum is second-order. The precise mechanism of the phase transition
becomes irrelevant as this now occurs after inflation has ended. After passing φǫ the field reaches
φ = 0 within one Hubble time. Unlike the second-order model this does not immediately cause an
instability, and oscillations about φ = 0 could be sufficiently damped to restart inflation if φcr lies
very close to zero. However even in this case we can show that the number of e-foldings, given by
Eq. (2.14), during any subsequent stage of inflation
ln r λ′
16πλ
M ! ,
mPl
(5.4)
N <
1
8
1
4
+
must be very small.
Thus we will consider only the vacuum dominated branch in what follows, where we may take
λM 4 ≫ m2φ2
cr . Another reason for doing this is that we will have to ignore the evolution of the
φ field while calculating the nucleation rate for ψ from the false to the true vacuum. The correct
two–field result is not known so, in common with all other models of first–order inflation, we will
calculate instead the tunnelling rate for the quasi–static potential V (ψ). We would only expect this
to be valid if m ∼< H , which is indeed guaranteed if we are in the vacuum dominated regime.
The percolation parameter is then given by
λM 4
p ≃
4H 4 exp(−SE ) ,
where the term in the exponential is the Euclidean action of the tunnelling configuration [61], recently
given for first-order quartic potentials V (ψ) by Adams [62] as SE = 2π2B4 /λ, with B4 a numerically
calculated monotonically increasing fitting function of the parameter
(5.5)
δ(φ) ≡
9λ α
γ 2 .
(5.6)
.
In our model δ(φ) decreases as φ rolls down its potential during inflation, until SE is sufficiently
small for the percolation parameter to reach unity allowing the first-order transition to complete.
This corresponds to
(5.7)
Scr = ln
λM 4
mPl
4pcrH 4 ≃ 4 ln
M
Figure 3 shows the corresponding value of δcr required for the transition to complete at different
values of the false vacuum energy density. Clearly for a given α there is a lower bound, δ > δ0 =
9λα/γ 2 , and a corresponding bound on the nucleation rate, so the first-order transition cannot
complete when the energy density of the false vacuum is above a given value. For instance, if λ = 1,
and α ∼> 1 then the transition will never complete for M ∼> 1014GeV.
Assuming then that α is sufficiently small (i. e. δ0 < δcr(M )), bubbles of the true vacuum will
percolate at φcr = pλ/λ′eff M , where, to utilise the results of Section 2, we write
9λ2
λ′ .
λ′eff =
γ 2 (δcr − δ0 )
To complicate the matter somewhat, λ′eff is now a function of M through the dependence of δcr
on the energy density, but this dependence is very weak for the energy scales significantly below
the Planck scale which we are interested in. With this proviso then, the results for the vacuum
dominated branch of Section 2 in the second-order model may be carried through to the first-order
model by replacing λ′ by λ′eff .
(5.8)
29
5.2 Big bubble constraints
The production of large true vacuum voids, nucleated early on during inflation and swept up to
astrophysical sizes by the subsequent expansion, can severely constrain some models of first-order
inflation [63, 64]. The isotropy of the microwave background can be used to rule out the possibility
that there are any voids with a comoving size greater than about 20h−1Mpc on the last scattering
surface [64], which corresponds to a filling fraction of less than about 10−5 for bubbles nucleated
around 55 e-foldings before the end of inflation. This means that the percolation parameter at this
point during inflation must be less than 10−5 , requiring
(5.9)
(5.10)
(5.11)
(5.12)
M .
+ 11.5 .
S55 ∼> 4 ln
This gives the second line in Figure 3 showing the minimum permissible value of δ (denoted by δ∗ )
at 55 e-foldings before the end of inflation at different false vacuum energy densities.
> δ∗ , requires α to be greater than a minimum value at this
Obeying this extra constraint, δ55 ∼
point and thus
φ55 ∼> γr δ∗ − δ0
9λλ′
In the vacuum dominated regime the value of φ can be given as a function of the number of e-foldings
before the end of inflation from Eq. (2.14)
φ ≃ φcr exp (cid:18) N m2m2
2πλM 4 (cid:19) ,
Pl
which gives the constraint in Eq. (5.10) as a constraint on the mass scales
m2m2
δcr − δ0 (cid:19) .
ln (cid:18) δ∗ − δ0
Pl
>
M 4 ∼
In other words, the mass of the φ field, m, must be large enough for the decrease in the effective
mass of the ψ field during the last 55 e-foldings of inflation to raise the percolation parameter from
10−5 to unity. The numerical factor on the right-hand side of this equation is fairly small, typically
about 10−2 for λ ∼ 1, so this does not threaten to force us out of the small m limit. Clearly it is
minimised for small α, as δ0 → 0, but can become large if δcr is too close to δ0 .
Normalising the parameters of the model by the observed density perturbations, as described in
Section 2, gives another relation between m and M which, combined with the big-bubble constraint,
provides limits on either m or M alone:
55λ r δcr − δ0
δcr − δ0 (cid:19) ,
ln (cid:18) δ∗ − δ0
M
πγ
> (3 × 10−5)
mPl ∼
9λ′
δcr − δ0 (cid:19)5/2
mPl ∼> (3 × 10−5)2 γ 2
55 (cid:17)5/2 (cid:18)ln
9λ′ (cid:19) (cid:16) π
λ3/2 (cid:18) δcr − δ0
m
δ∗ − δ0
For reasonable values of the coupling parameters, of order unity, we would expect the right-hand
side of Eq. (5.13) to be ∼ 10−6 placing a lower limit on M of around 1013GeV in a first-order model,
unless we have δcr very close to δ0 . The other way to allow first-order models at lower energy scales
would be to introduce a strong coupling λ′ , much larger than unity, between the two fields, which is
clearly always possible as this enables only a small change in φ to effect a large change in the bubble
nucleation rate.
πλ
55
.
(5.14)
(5.13)
mPl
M
30
5.3 Other first-order models
The ability of a second scalar field to allow a first-order inflationary phase transition to complete was
first emphasised by La and Steinhardt [8]. This is the basis of models of extended inflation based on
extensions to the gravitational lagrangian beyond the Einstein-Hilbert action of general relativity
[8, 9, 10]. In Brans-Dicke gravity, for instance, the Ricci scalar appears in the action coupled to a
scalar field rather than Newton’s constant and it is this growing Brans-Dicke field, Φ ≡ m2
Pl , which
triggers the completion of the phase transition in the ψ field. However Linde [2] and Adams and
Freese [3], pointed out that this basic scenario can also be realised in general relativity by coupling
the inflaton to a second scalar field. Linde used the same basic first-order potential V (ψ) as we have,
although he used a Coleman-Weinberg type potential for φ rolling down from +∞ introducing a
minimum at a non-zero value. This would have to be included in the minimum value of α and thus
δ . Adams and Freese considered a specific interaction rather different to ours where as φ rolled down
its potential, the energy of the false vacuum state actually increased relative to the true vacuum,
but their more general discussion was clearly intended to include models such as the one we have
examined here.
The bubble nucleation constraints in terms of δcr and δ55 are independent of the type of first–
order inflation being considered. Extended inflation models consider a first–order potential for the
inflaton which does not change during inflation. Thus δ remains a constant, as does the false vacuum
energy density. The time–varying quantity here is the Planck mass which grows during inflation.
Thus extended inflation models proceed horizontally, from right to left across the parameter space
in Figure 3, completing the phase transition when δcr (M /mPl) = δ . The general relativistic models
considered here, and those considered by Linde and by Adams and Freese, proceed almost vertically
as the false vacuum energy density remains approximately constant as δ decreases with φ. Because
the percolation parameter p is exponentially dependent on the Euclidean action, SE , it is relatively
easy to evade the big–bubble constraints in the general relativistic models varying δ .
In models
where only M or mPl varies, the percolation parameter tends to grow comparatively slowly making
the big–bubble contraint much more severe, especially as V /m4
Pl and V ′/V are already constrained
by density pertubations at 60 e-foldings [25].
This has led other authors [5, 38, 10] recently to consider models of extended inflation where
non-minimal coupling can also change the shape of the ‘effective potential’, making M 2
ψ = ξR − λM 2
for instance. In such cases inflation could again end by a first– or second–order transition. In a de
Sitter metric the Ricci scalar R is a constant (R = 12H 2) so a false vacuum dominated universe
in general relativity does not yield a time-varying mass. But in Brans-Dicke gravity for instance,
where the dominant coupling to the Ricci scalar is via the Brans-Dicke field (rather than a constant)
the expansion is power-law [65] rather than exponential and R ∝ t−2 , triggering an instability when
R ≤ pλ/ξ m. Similar models have been proposed in higher order gravity theories, coupling the ψ
field to R2 terms [10]. These models extending the gravity lagrangian can be re-written in terms
of a general relativistic model with two interacting scalar fields (the defect field and a dilaton field
that acts as the inflaton) using a conformally rescaled metric [66]. But the scalar field lagrangian in
this case is rather different from our model as not only Mψ but all the mass scales are changed by
the dilaton field. These first–order models thus correspond to a more complicated path on Figure
3, and by making δ a function of time can also evade the big–bubble constraint.
6 Discussion and Conclusions
In conclusion, models of inflation based on Einstein gravity, but driven by a false vacuum, offer a
range of new possibilities for both theory and phenomenology.
On the particle physics side, we have shown how false vacuum inflation points to new possibiities
for model building.
In particular, we have shown that it can occur in a class of supergravity
models implied by orbifold compactification of superstrings. One outcome of that discussion was
31
the intriguing possibility of obtaining a handle on the superstring orbifold, through the fact that
one-loop corrections might be the dominant effect determining the spectral index. Much remains to
be done of course. For instance, although we have exhibited a toy model for the scalar field sector of
the string derived supergravity theory, we have made no attempt to put it in the context of a realistic
model involving other fields as well. In particular we have not tried to extend to supergravity the
identification of the false vacuum with that of Peccei-Quinn symmetry, which we found was both
viable and attractive in the context of global supersymmetry.
In terms of direct cosmological phenomenology, false vacuum dominated inflation offers the un-
usual option of a spectral index for the density perturbations exceeding unity, though we have
demonstrated that with the COBE normalisation the deviation can only be rather modest with a
plausible maximum of around n = 1.14. There is however additional interest in that one expects
topological defects to form as the false vacuum decays; because essentially all the energy density is
available to go into the defect fields, the energy available is much greater than in usual models where
reheating is required first, redistributing the energy into a large number of fields. Because of this,
structure-forming defects are comfortably compatible with our inflation model when the masses are
towards the top of their allowed ranges.
We have also made a preliminary investigation of the details of the phase transition in different
regimes, though much remains to be done. For a second-order phase transition, results already exist
in the literature describing the inflaton dominated regime. We have demonstrated that, barring
very weak couplings, the phase transition proceeds very rapidly in the vacuum dominated regime,
but have been unable to develop a solid understanding of the statistics of the defects produced in
such a transition.
In the first-order case, where the transition completes via bubble nucleation,
we have gone on to calculate the bubble distribution and the constraints upon it. We note that
first-order inflation models based on Einstein gravity are generally easier to implement than those
of the extended inflation type.
That one can have both structure-forming topological defects and inflation raises a host of possible
structure formation scenarios, as one could choose to utilise only one of these two or a combination
of the two. It is believed [34] that for a given size of density perturbation (i. e. perturbation in
the gravitational potential), defects give a larger microwave background temperature anisotropy, by
a factor of a few. One could therefore arrange for defects to be the source of a component of the
COBE signal while having only a modest effect on structure formation; alternatively one could aim
to have inflation and defects contributing roughly equally to structure formation in which case the
defects would be predominant in the microwave background. It is conceptually (and calculationally)
preferable to take the option of using only one source, lowering the energy scale of the other to make
its effects negligible, but one should be aware that the required scales of the two are similar, and
should a realistic model along our suggested lines be devised it would not be a particular surprise
should both contributions have a role to play.
Acknowledgements
EJC and DW are supported by the SERC, ARL by the SERC and the Royal Society and EDS by a
JSPS Postdoctoral Fellowship and Monbusho Grant-in-Aid for Encouragement of Young Scientists,
No. 92062. EJC and ARL acknowledge the hospitality of the Aspen Center for Physics, ARL that of
Fermilab and and DHL that of CERN; part of this work was carried out at each of these places. We
would like to thank Mark Hindmarsh and Andrei Linde for helpful discussions, and Kiwoon Choi,
Burt Ovrut and Graham Ross for helpful comments on early drafts of the paper. ARL and DW
acknowledge the use of the STARLINK computer system at the University of Sussex.
References
32
[1] A. D. Linde, Phys. Lett. B129, 177 (1983);
A. D. Linde, Particle Physics and Inflationary Cosmology, Harwood Academic, Switzerland
(1990).
[2] A. Linde, Phys. Lett. B249, 18 (1990).
[3] F. C. Adams and K. Freese, Phys. Rev. D43, 353 (1991).
[4] A. D. Linde, Phys. Lett. B259, 38 (1991).
[5] A. D. Linde, “Hybrid Inflation”, to appear, Phys. Rev. D49, (Jan 15th, 1994).
[6] A. R. Liddle and D. H. Lyth, Phys. Rep. 231, 1 (1993).
[7] S. Mollerach, S. Matarrese and F. Lucchin, “Blue Perturbation Spectra from Inflation”, CERN
preprint, astro-ph/9309054 (1993).
[8] D. La and P. J. Steinhardt, Phys. Rev. Lett. 62, 376 (1989).
[9] P. J. Steinhardt and F. S. Accetta, Phys. Rev. Lett. 64, 2740 (1990);
J. D. Barrow and K. Maeda, Nucl. Phys. B341, 294 (1990).
[10] L. Amendola, S. Capozziello, M. Litterio and F. Occhionero, Phys. Rev. D45, 417 (1992).
[11] L. A. Kofman and A. D. Linde, Nucl. Phys. B282, 555 (1987).
[12] E. T. Vishniac, K. A. Olive and D. Seckel, Nucl. Phys. B289, 717 (1987).
[13] L. A. Kofman and D. Yu. Pogosyan, Phys. Lett. B214, 508 (1988).
[14] J. Yokoyama, Phys. Lett. B212, 273 (1988); Phys. Rev. Lett. 63, 712 (1989).
[15] L. Kofman, G. R. Blumenthal, H. Hodges and J. R. Primack, Proceedings of the Workshop on
LS Structure, Rio (1989), eds D W Latham and L N da Costa.
[16] D. S. Salopek, J. R. Bond and J. M. Bardeen, Phys. Rev. D40, 1753 (1989).
[17] D. H. Lyth, Phys. Lett. B246, 359 (1990).
[18] H. M. Hodges and J. R. Primack, Phys. Rev. D43, 3155 (1991).
[19] M. Nagasawa and J. Yokoyama, Nucl. Phys. B370, 472 (1992).
[20] M. Kawashi and K. Sato, Phys. Lett. B189, 23 (1987).
[21] G. F. Smoot et al, Astrophys. J. Lett. 396, L1 (1992).
[22] E. L. Wright et al, “Comments on the Statistical Analysis of Excess Variance in the COBE
DMR Maps”, to appear, Astrophys. J. (1994).
[23] A. R. Liddle, “The Inflationary Energy Scale”, to appear, Phys. Rev. D49 (Jan 15th, 1994).
[24] A. Linde, Phys. Lett. B284, 215 (1992).
[25] A. R. Liddle and D. H. Lyth, Phys. Lett. B291, 391 (1992).
[26] A. A. Starobinsky, Sov. Astron. Lett. 11, 133 (1985).
[27] T. W. B. Kibble, J. Phys. A9, 1387 (1976).
[28] D. H. Lyth, Phys. Lett. B147, 403 (1984); ibid B150, 465 (E) (1985).
33
[29] A. H Guth and S.-Y. Pi, Phys. Rev. D32, 1899 (1985).
[30] A. D. Linde and D. H. Lyth, Phys. Lett. B246, 353 (1990).
[31] D. H. Lyth, Phys. Rev. D45, 3394 (1992).
[32] D. H. Lyth and E. D. Stewart, Phys. Rev. D46, 532 (1992).
[33] E. Copeland, D. Haws, S. Holbraad and R. Rivers, ‘The Statistical Properties of Strings. I Free
Strings’, in ‘The Formation and Evolution of Cosmic Strings’, eds G.Gibbons, S. Hawking and
T. Vachaspati, (CUP 1990).
[34] A. Stebbins, Ann. N. Y. Acad. Sci. 688, 824 (1993).
[35] T. Vachaspati and A. Vilenkin, Phys. Rev. Lett. 67 1057 (1991).
[36] A. Albrecht and A. Stebbins, Phys. Rev. Lett. 69, 2615 (1992)
[37] M. Hindmarsh, “Small scale microwave background fluctuations from cosmic strings”, Cam-
bridge preprint DAMTP-93-28, astro-ph/9307040 (1993).
[38] A. M. Laycock and A. R. Liddle, “Extended Inflation with a Curvature-Coupled Inflaton”, to
appear, Phys. Rev. D49 (Feb 15th, 1994).
[39] J. Preskill, Phys. Rev. Lett. 43, 1365 (1979).
[40] E. N. Parker, Ap. J. 160, 383 (1970).
M. S. Turner, E. N. Parker and T. Bogdan, Phys. Rev. D 26, 1296 (1982).
[41] E. W. Kolb and M. S. Turner, The Early Universe, Addison Wesley (1990).
[42] For a review of supersymmetry and supergravity, see H. P. Nilles, Phys. Rep. 110, 1 (1984).
[43] J. E. Kim, Phys. Lett. B136, 378 (1984);
J. E. Kim and H. P. Nilles, Phys. Lett. B138, 150 (1984);
K. Ra jagopal, M. S. Turner and F. Wilczek, Nucl. Phys. B358, 447 (1991);
T. Goto and M. Yamaguchi, Phys. Lett. B276, 103 (1992).
[44] K. Freese, J. A. Frieman and A. V. Olinto, Phys. Rev. Lett. 65, 3233 (1990);
F. C. Adams, J. R. Bond, K. Freese, J. A. Frieman and A. V. Olinto, Phys. Rev. D47, 426
(1993).
[45] K. Olive, Phys. Rep. 190, 307 (1990).
[46] D. Lust and C. Munoz, Phys. Lett. B279, 272 (1992).
[47] S. Ferrara, C. Kounnas and M. Porrati, Phys. Lett. B181, 263 (1986).
[48] J.-P. Derendinger, S. Ferrara, C. Kounnas and F. Zwirner, Nucl. Phys. B372, 145 (1992).
[49] B. de Carlos, J. A. Casas and C. Munoz, Nucl. Phys. B399, 623 (1993).
[50] A. Brignole, L. E. Ib´anez and C. Munoz, “Towards a Theory of Soft Terms for the Super-
symmetric Standard Model”, Universidad Aut´onoma de Madrid preprint FTUAM-26/93, hep-
ph/9308271 (1993).
[51] M. Cvetic, A. Font, L. E. Ib´anez, D. Lust and F. Quevedo, Nucl. Phys. B361, 194 (1991).
34
[52] S. Ferrara, D. Lust, A. Shapere and S. Theisen, Phys. Lett. B225, 363 (1989);
S. Ferrara, D. Lust and S. Theisen, Phys. Lett. B233, 147 (1989).
[53] L. E. Ib´anez and D. Lust, Nucl. Phys. B382, 305 (1992).
[54] E. D. Stewart, in preparation.
[55] E. Witten, Phys. Lett. B155, 151 (1985).
[56] R. Brustein and P. J. Steinhardt, Phys. Lett. B302, 196 (1993).
[57] B. de Carlos, J. A. Casas, F. Quevedo and E. Roulet, “Model-Independent Properties and
Cosmological Implications of the Dilaton and Moduli Sectors of 4-D Strings”, CERN preprint
CERN-TH.6958/93, hep-ph/9308325 (1993).
[58] A. Font, L. E. Ib´anez, D. Lust and F. Quevedo, Phys. Lett. B249, 35 (1990).
[59] A. H. Guth, Phys. Rev. D23, 347 (1981).
[60] A. H. Guth and E. J. Weinberg, Nucl. Phys. B212, 321 (1983).
[61] S. Coleman, Phys. Rev. D15, 2929 (1977).
[62] F. C. Adams, Phys. Rev. D48, 2800 (1993).
[63] E. J. Weinberg, Phys. Rev. D40, 3950 (1989);
A. R. Liddle and D. Wands, Phys. Rev. D45, 2665 (1992);
M. S. Turner, E. J. Weinberg and L. M. Widrow, Phys. Rev. D46, 2384 (1992).
[64] A. R. Liddle and D. Wands, Mon. Not. Roy. astr. Soc. 253, 637 (1991).
[65] H. Nariai, Prog. Theor. Phys. 42, 544 (1969).
[66] K. Maeda, Phys. Rev. D39, 3159 (1989).
Figure Captions
Figure 1
The solid line shows the locus of M and m which satisfy the COBE normalisation for λ = λ′ = 1.
The two analytic branches are clearly seen. The dot-dashed line indicates the analytic solution for
the extreme vacuum dominated branch, as utilised by Linde [5]; the deviation of the exact solution
from it is caused by the increasing significance of the exponential term in Eq. (2.37) which is included
in the parametric analytic solution Eq. (2.36), as used in [7] and indicated here by the dotted line.
The dotted line (hidden under the solid for most of its length) terminates when the regime becomes
invalid, though it actually extends somewhat beyond the exact solution because we have interpreted
≪ as ∼< in places.
Figure 2
The spectral index n is shown for the exact COBE normalised models of Figure 1.
Figure 3
δcr and δ∗ plotted as functions of V 1/4/mPl for first-order inflation. 55 e-foldings from the end of
35
inflation δ55 must lie above the dotted line, δ∗ , but then reach the solid line, δcr , to bring inflation
to an end. The two tra jectories, plotted as dashed lines, represent the typical evolution of δ and
M /mPL for (a) extended inflation where δ =constant, and (b) false vacuum inflation in Einstein
gravity where V 1/4/mPl ≃ constant.
36
This figure "fig1-1.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9401011v1
This figure "fig1-2.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9401011v1
This figure "fig1-3.png" is available in "png"(cid:10) format from:
http://arXiv.org/ps/astro-ph/9401011v1
|
astro-ph/9801158 | 1 | 9801 | 1998-01-15T22:42:55 | Gravitational Lens Magnification and the Mass of Abell 1689 | [
"astro-ph"
] | We present the first application of lens magnification to measure the absolute mass of a galaxy cluster; Abell 1689. The absolute mass of a galaxy cluster can be measured by the gravitational lens magnification of a background galaxy population by the cluster potential. The lensing signal is complicated by the variation in number counts due to galaxy clustering and shot-noise, and by additional uncertainties in relating magnification to mass in the strong lensing regime. Clustering and shot-noise can be dealt with using maximum likelihood methods. Local approximations can then be used to estimate the mass from magnification. Alternatively if the lens is axially symmetric we show that the amplification equation can be solved nonlocally for the surface mass density and the tangential shear. In this paper we present the first maps of the total mass distribution in Abell 1689, measured from the deficit of lensed red galaxies behind the cluster. Although noisier, these reproduce the main features of mass maps made using the shear distortion of background galaxies but have the correct normalisation, finally breaking the ``sheet-mass'' degeneracy that has plagued lensing methods based on shear. We derive the cluster mass profile in the inner 4' (0.48 Mpc/h). These show a profile with a near isothermal surface mass density \kappa = (0.5+/-0.1)(\theta/1')^{-1} out to a radius of 2.4' (0.28Mpc/h), followed by a sudden drop into noise. We find that the projected mass interior to 0.24 h^{-1}$Mpc is M(<0.24 Mpc/h)=(0.50+/- 0.09) \times 10^{15} Msol/h. We compare our results with masses estimated from X-ray temperatures and line-of-sight velocity dispersions, as well as weak shear and lensing arclets and find all are in fair agreement for Abell 1698. | astro-ph | astro-ph | For publication in The Astrophysical Journal
The Astrophysical Journal, 000: 000–000, 2011
GRAVITATIONAL LENS MAGNIFICATION AND
THE MASS OF ABELL 1689
8
9
9
1
n
a
J
5
1
1
v
8
5
1
1
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
A.N. TAYLOR, S. DYE
Department of Astronomy, University of Edinburgh, Blackford Hil l, Edinburgh EH9 3HJ, UK; [email protected], [email protected]
T.J. BROADHURST, N. BEN´ITEZ
Department of Astronomy, Campbel l Hal l, University of California, Berkeley, USA; [email protected]; [email protected]
E. VAN KAMPEN
Royal Observatory Edinburgh, Blackford Hil l, Edinburgh EH9 3HJ,
Theoretical Astrophysics Center, Juliane Maries Vej 30, DK-2100 Copenhagen, Denmark; [email protected]
AND
28 December 2011
ABSTRACT
We present the first application of lens magnification to measure the absolute mass of a galaxy cluster; Abell 1689. The
absolute mass of a galaxy cluster can be measured by the gravitational lens magnification of a background galaxy popu-
lation by the cluster gravitational potential. The lensing signal is complicated by the intrinsic variation in number counts
due to galaxy clustering and shot–noise, and by additional uncertainties in relating magnification to mass in the strong
lensing regime. Clustering and shot–noise can be dealt with using maximum likelihood methods. Local approximations
can then be used to estimate the mass from magnification. Alternatively if the lens is axially symmetric we show that
the amplification equation can be solved nonlocally for the surface mass density and the tangential shear.
In this paper we present the first maps of the total mass distribution in Abell 1689, measured from the deficit of
lensed red galaxies behind the cluster. Although noisier, these reproduce the main features of mass maps made using the
shear distortion of background galaxies, but have the correct normalisation, finally breaking the “sheet–mass” degeneracy
that has plagued lensing methods based on shear. Averaging over annular bins centered on the peak of the light distribution
we derive the cluster mass profile in the inner 4′ (0.48h−1Mpc). These show a profile with a near isothermal surface mass
density κ ≈ (0.5 ± 0.1)(θ/1′ )−1 out to a radius of 2.4′ (0.28h−1Mpc), followed by a sudden drop into noise. We find that
the pro jected mass interior to 0.24h−1Mpc is M (< 0.24h−1Mpc) = (0.50 ± 0.09) × 1015 h−1M⊙ .
We compare our results with masses estimated from X-ray temperatures and line-of-sight velocity dispersions, as
well as weak shear and lensing arclets. We find that the masses inferred from X-ray, line-of-sight velocity dispersions,
arclets and weak shear are all in fair agreement for Abell 1698.
Subject Headings : Cosmology: gravitational lensing, clusters, dark matter
1
INTRODUCTION
The magnitude and distribution of matter in galaxy clusters
should in principle provide a strong constraint on cosmologi-
cal models of structure formation and the mean mass density
of the Universe. In addition a direct image of the mass den-
sity will tell us much about the relationship between gas,
galaxies and dark matter and whether light is indeed a fair
– if biased – tracer of mass.
Early techniques for estimating the mass in clusters in-
clude dynamical methods, from the line–of–sight velocity
dispersion of member galaxies, and X-ray temperature mea-
surements. However these estimates make some strong as-
sumptions about equilibrium conditions in the cluster.
Kaiser and Squires (1993) circumvented this by showing
that a more direct method of estimating the mass, with no
underlying assumptions about the dynamical or thermody-
namical state of the cluster, was to measure the shear field
in the source distribution of the cluster background (Kaiser
& Squires 1993; Tyson, Valdes & Wenk 1990, Schneider &
Seitz 1995 ). On average the shear pattern of a population
of unlensed galaxies should be randomly distributed. But in
the presence of a massive gravitational lensing cluster, the
shear field is polarized. Since the shear field is related (non-
locally) to the surface mass density the shear can be used
to estimate the mass distribution – up to an arbitrary con-
stant. The presence of this arbitrary constant, referred to
as the “sheet–mass” degeneracy, means that only differen-
The amplification index, β − 1, accounts for the expansion
of the background image and for the increase in number as
faint sources are lensed above the flux limit.
In the absence of galaxy clustering and finite sampling
effects the background galaxy distribution can simply be
inverted, via equation (1), to find the amplification. One
can then solve equation (2) to find the surface density, with
some realistic assumptions about the shear. In Section 4 we
discuss various approximations that allow us to do this.
However, given a small resolution scale for the surface
amplification, galaxy clustering and finite sampling will in
general be an important effect. In the next Section we discuss
the effects of intrinsic variation in the distribution of the
background galaxy sources.
3 GALAXY CLUSTERING NOISE
2
A.N. Taylor et al.
tial masses can be measured. Shear maps are conventionally
normalised to the edge of the observed field, or such that the
inferred mass–density is everywhere positive, and so repre-
sent a lower limit on the mass.
Soon after, Broadhurst, Taylor & Peacock (1995; here-
after BTP) showed that the sheet–mass degeneracy could be
broken by use of the gravitational lens magnification effect.
The number and magnitude–redshift distribution of back-
ground galaxies is distorted by the gravitational field of the
lensing cluster and in the weak lensing regime this distor-
tion provides a straightforward estimate of the surface mass
density. With calibration from off-set fields the cluster mass
distribution can be properly normalized.
BTP also suggested that a degraded, but much quicker,
estimate of the magnification effect could be made from the
distortion of angular number counts of background sources.
Broadhurst (1995) found evidence for this distortion in the
background counts of the cluster Abell 1689, as did Fort,
Mellier & Dantel-Fort (1997) for Cl0024. In this work we ap-
ply the methods developed by BTP and extended by Taylor
& Dye (1998) to estimate the surface mass density from the
distortion of angular counts, including the effects of shot–
noise and galaxy clustering, and van Kampen (1998) on esti-
mating the surface mass density in the strong lensing regime,
to Abell 1689.
The layout of the paper is as follows. In Section 2 we
describe the magnification effect itself. In Section 3 we de-
scribe the effects of shot noise and clustering on estimates of
the surface mass density. In Section 4 we describe how to es-
timate the surface mass density in the strong lensing regime
using a local approximations, and introduce a new, self–
consistent, nonlocal solution for axially symmetric lenses.
We apply these methods to map out the mass in the cluster
Abell 1689 in Section 5 and find its profile. Our mass esti-
mate is compared to other estimates in Section 6 and our
conclusions are presented in Section 7.
2 THE MAGNIFICATION EFFECT
The observed number of galaxies seen in pro jection on the
sky is (BTP, Taylor & Dye 1998)
n′ = n0Aβ−1(1 + Θ),
(1)
where n0 is the expected mean number of galaxies in a given
area at a given magnitude. Variations in this mean arise
from the angular perturbation in galaxy density, Θ, due to
galaxy clustering, and from gravitational lens magnification.
The lens amplification factor is
A = (1 − κ)2 − γ 2 −1
where
(2)
κ =
Σ
Σcrit
(3)
is the surface mass density in units of the critical surface
mass, Σcrit . The amplitude of the shear field is given by
γ and the background galaxy luminosity function is locally
approximated by
n(L) ∼ L−β .
(4)
P (n) =
=
The main sources of uncertainty in lens magnification are
due to shot-noise, from finite sampling, and the intrinsic
clustering of the background source population which in-
troduce correlated fluctuations in the angular counts. As
we are viewing small–angles the clustering properties of the
background source galaxies are not in general linear, unless
the depth of background is sufficient to wash out the clus-
tering pattern. As a result it is not sufficient to make the
usual assumption that galaxy clustering can be modelled by
a Gaussian distribution.
We can account for the effects of shot–noise and non-
linear clustering by modeling the angular counts by a
Lognormal–Poisson model (Coles & Jones 1991, BTP, Tay-
lor & Dye 1998) – a random point–process sampling of a
lognormal density field. The distribution function of source
counts is then
1
n! hλn e−λ i,
(5)
n! Z ∞
2σ 2 − λ0ex − nx(cid:19) (6)
exp (cid:18)−
λn
x2
dx
0
√2πσ
−∞
where λ = λ0 ex is the local mean density, x is a Gaus-
sian random variable of zero mean and variance σ 2 and
λ0 = n0Aβ−1e−σ2/2 correctly normalises the counts. The
linear clustering variance, σ 2 , is related to the nonlinear
variance by σ 2 = ln[1 + σ 2
nl ]. We have tested this distri-
bution against available data and find that it is an excellent
fit to the distribution of counts in the deep fields. The only
parameters are the observed count per pixel, n, and the vari-
ance of the lognormal field. The amplitude of clustering of
the density field and its dependence of redshift can be esti-
mated from, e.g., the I-band selected galaxies in the Canada–
France Redshift Survey in the range 17.5 < I < 22.5 (Le
F´evre 1996; see section 5.2.3). The quantity required is the
variance in a given area of sky, which can be estimated by
averaging the observed angular correlation function, ω (θ),
over a given area:
Ω(θ) ZΩ
1
where Ω(θ) is the area.
Our method of approach is then that discussed by BTP.
At each pixel in a map of the source counts one uses the
distribution equ. (6) as a likelihood function, L(An, σ ) =
σ 2
nl = ¯ω (θ) =
d2θ ′ ω (θ ′ ),
(7)
P (nσ, A), assuming a uniform prior for the amplification.
The surface density is then found from the amplification
by making some realistic assumption about the shear and
maximizing the likelihood. In the next Section we discuss a
number of ways of transforming from the amplification to κ
in the strong lensing regime.
4 THE STRONG LENSING REGIME
Transforming from amplification to the surface mass density
is potentially non-trivial, as we have no shear information.
One could incorporate this from independent measurements
of the shear field, but for the present discussion we are in-
terested in developing a completely independent lensing ap-
proach. We shall discuss combining shear and magnification
elsewhere. In principle one could generate a first guess for
the surface mass density and iterate the amplification equa-
tion towards a solution of both surface density and shear.
However, given the small field of view and uncertainties in-
troduced by parity changes, this can be an unstable prob-
lem. In addition, as the solutions are in general multivalued,
we would hope to start from as near to the correct solution
as possible. In this section we discuss a number of reason-
able approximations to solve the amplification equation (2).
These can be regarded as solutions in their own right, or
as the first best guess to an iterated solution. We begin by
discussing the local approximation methods suggested and
fully tested on simulated clusters by van Kampen (1998).
Then in Section 4.2 we present a new, self–consistent solu-
tion to the amplification equation, for κ and γ for an axially
symmetric lens.
4.1 Local approximations to the surface mass
density
(9)
(8)
There exist only two local relations between γ and κ that re-
sult in a single caustic solution of the amplification equation
(2) which is easily invertible (van Kampen 1998): γ = 0, cor-
responding to a sheet of matter, and γ = κ, for an isotropic
lens. These two relations have corresponding estimators for
κ as a function of amplification:
κ0 ≡ κ(γ = 0) = 1 − P A−1/2 ,
1
(1 − P A−1 ),
κ1 ≡ κ(γ = κ) =
2
where P = ±1 is the image parity.
Let us assume that the surface mass density of the lens
is smooth over some scale. In this case, for a sufficiently
smooth lens γ ≤ κ (BTP). The equality holds in the case
of an isotropic lens, for instance the isothermal lens. The
inequality holds for any anisotropic lens, with the sheet mass
at the extreme. For a smooth lens these two estimates bound
the true value κ1 ≤ κ ≤ κ0 . Before caustic crossing it can
also be shown that κ1 ≤ κ0 ≤ κweak holds, where A =
1 + 2κweak is the weak lensing limit (BTP). Hence the weak
lensing approximation will overestimate the cluster mass in
the strong regime, usually by a factor of two (van Kampen
1998).
In practice, substructure and asphericity of the clus-
ter will induce extra shear (e.g. Bartelmann, Steinmetz &
Gravitational lens magnification
3
,
(10)
Weiss 1995), especially in the surrounding low-κ neighbour-
hood, where substructure is relatively more dominant and
filaments make the cluster most aspherical. This means that
the lens will not be smooth for small κ, and therefore κ1 is
a lower limit for the true κ only for the central parts of the
cluster, in the case the lens parity is known. Van Kampen
(1998) found it to be a good lower limit only for κ > 0.4 (for
the most massive clusters), while for κ < 0.2, κ1 is usually
fairly close to the true value. For angle–averaged κ-profiles
κ1 is a good lower limit for hκiθ > 0.2. All this has no bear-
ing on κ0 , which remains a strong upper limit until the first
caustic crossing.
An approximation that tries to take these cluster lens
features into account, while still giving an invertible A(κ)
relation, is (van Kampen 1998):
γ = 1 − cq κ
c
which results in an amplification relation that admits the
full four solutions:
A−1 = (κ − c)(κ − 1/c),
with caustics at κ = c and 1/c. The solution for κ is then
1
2c (cid:16)(c2 + 1) − Sp(c2 + 1)2 − 4c2 (1 − P A−1 )(cid:17) .
Solutions are set by choosing the parities P , S = ±1 where P
is the image parity and S is the sign of ((c2 + 1)/2c−κ). Note
that the sheet–like solution is recovered by setting c = 1.
Figure 1 shows a plot of κ versus the inverse amplifica-
tion, A−1 , for the three estimators. Also shown is the weak
field approximation. The points are taken from a simulated
lensing cluster (van Kampen & Kargert 1997) which is of
comparable size to A1689. It is clear that κ0 is a strong
bound, at least until a caustic is crossed, and that κ1 pro-
vides a very good bound for κ > 0.2. The weak field approx-
imation however is extremely bad except in the very weak
regime (κ < 0.1). The two–caustic approximation behaves
as it is designed to do: a good fit between the other two
strong lensing estimators for the central parts of the cluster,
while also modelling the γ > κ behaviour for small κ. These
results are fairly robust over a wide range of clusters and for
all realistic values of the cosmological density parameter Ω0 .
κc =
(11)
(12)
4.2 A nonlocal approximation to the surface
mass density
An alternative approach is to assume axial symmetry for
the lens. Because this fixes a non-local functional relation-
ship between κ and γ (equation 15), we can solve the am-
plification equation (2) for a self–consistent κ and γ profile.
Although we shall apply our results to circularly averaged
data, these results hold for any self-similar embedded set of
contours.
We define a mean surface density interior to a contour
by integration over the interior area, Ω(θ),
Ω(θ) ZΩ
1
The deflection angle for the axi-symmetric lens is
d2θ ′ κ(θ ′ )
¯κ(θ) =
(13)
∆θ = θ ¯κ,
(14)
4
A.N. Taylor et al.
1
-
A
1.0
0.8
0.6
0.4
0.2
0.0
0.0
0.2
0.4
κ
where P , S = ±1 are again the image parities. The only free-
dom we have, for a given amplification profile is the choice
of the shear on the first shell, γ1 = η0 , and the parity. It
should be noted that given the amplification and having
fixed the parities one has to ensure that the first γ satis-
fies γ 2 ≥ P A−1 , to avoid unphysical solutions. The nonlocal
approximation contains both the sheet and isothermal solu-
tions as specific solutions. The uncertainty on κ and γ can
be found by simple error propagation of the uncertainty on
the measurement of the amplification.
Having shown in Sections 2, 3 and 4 how, in principle,
one can measure the surface mass density from angular num-
ber counts, in the next section we exploit these methods to
measure the mass distribution in the lensing cluster Abell
1689.
0.6
0.8
5 APPLICATION TO ABELL 1689
Figure 1. Scatter plot of the surface mass density, κ, versus
the inverse amplification, A−1 , for a simulated cluster in a CDM
universe (see van Kampen 1998 for details). The cluster is at a
redshift of 0.183 and the background population at z = 0.8. The
cluster was selected to look similar to Abell 1689. The solid line
is the γ = 0 strong lensing approximation. Before caustic crossing
this is a hard bound on the locus of points. The dashed line is the
γ = κ approximation, which is a good lower bound for κ > 0.2
for this cluster. The weak–lensing approximation (dotted–line) is
seen to be a very bad approximation for κ > 0.1. The dot-dashed
line is a best fit to the simulation for the 2-caustic approximation
γ = 1 − cpκ/c, with c = 0.7.
and the shear is given by
(15)
γ = γt = κ − ¯κ.
where the tangential term, γt , is the only component of shear
generated. The amplification factor is given by
A−1 = (1 − ¯κ)(1 − 2κ + ¯κ).
One can now simultaneously solve for the surface mass
density, shear and amplification by series solution. Firstly
we divide the surface mass into consecutive shells with equal
separation (any arbitrary separation can be used, we have
chosen a regular separation for convenience). If we split ¯κ
into an interior term, ηn−1 , and a surface term, then for the
nth shell we have
2
(n + 1)
¯κn = ηn−1 +
(16)
(17)
κn
where we have defined
mκm ,
ηn−1 =
2
n(n + 1)
n−1
Xm=1
The surface mass density in the nth shell is then given by
(n + 1)
4n (cid:0)n + 1 − (n − 1)ηn−1 −
S [(n − 1 − (n + 1)ηn−1 )2 + 4nP A−1
n ]1/2 (cid:1),
κn =
(19)
(18)
In this section we apply the methods discussed in sections
2, 3 and 4 to observational data. We begin by describing the
data.
5.1 The Data
5.1.1 Data acquisition and reduction
The data were obtained during a run in February 1994 at
ESO’s NTT 3.6m telescope, with 104 secs integration in the
V and I bands, and covers 70 square arcminutes on the clus-
ter. Seeing was similar in both bands, with FWHM of 0.8′′
and a CCD pixel scale of 0.34′′ . The EMMI instrument was
used throughout. The passbands and exposures were cho-
sen such that the cluster E/S0 galaxies would be bluer than
a good fraction of the background, requiring much deeper
imaging in the bluer passband for detection. The cluster was
observed down to a limiting magnitude of I = 24.
The images were debiased and flattened with skyflats
using standard IRAF procedures. After this there remained
some large scale gradients of a few percent, probably caused
by some rotation of the internal lens. We additionally cor-
rected each separate exposure with a smoothed version of it-
self, obtained after masking out the cluster and other bright
ob jects. Following this we had homogeneous photometry
across the field (A further discussion of the reduction pro-
cedure can be found in Ben´ıtez, et al, 1998, in preparation).
The zero point was found to be good to 0.1 magnitudes.
High humidity on a few nights meant that some of the data
was not photometric, so we calibrated with the photometric
data. The ob ject detection and classification was performed
with SExtractor.
5.1.2 Separation of cluster and background
To measure the distortion in background counts we must
first separate the background from cluster members and
mask off the area they obscure. Cluster galaxies were iden-
tified from the strong cluster E/SO colour sequence, which
forms a horizontal band across the colour–magnitude dia-
gram, shown in figure 2. The sharp upper edge of this band
Gravitational lens magnification
5
Figure 2. Colour–magnitude diagram for A1689, overlaid with
colour cuts used to isolate the cluster members from the back-
ground population; 20 < I < 24, 1.6 < V − I < 3.5 and V < 26.8.
The strong horizontal band of galaxies is the cluster E/SO se-
quence.
represents the reddest galaxies in the cluster. Galaxies red-
der than this are cosmologically redshifted, and hence rep-
resent a background population. As well as isolating cluster
members, this selection should also ensure that any fore-
ground galaxies are removed. Anything redder than V − I =
1.6 was selected as a background galaxy. Further colour cuts
where imposed to ensure completeness of the sample. The
range of magnitudes was restricted to 20 < I < 24 and the
V -band limited to V < 28. Finally we also cut at V −I < 3.5,
where the reddest galaxies cut off.
Since the identification of cluster members is important
to remove contamination of the background sample we also
checked our colour selected candidates with new data from
a photometric redshift survey of the same field (Dye et al,
1998, in preparation). We found general agreement with the
simpler colour selection.
Having identified foreground and cluster members we
produced a mask to eliminate those areas obscured by clus-
ter members that would otherwise bias the mass estimate.
To isolate the cluster members for the mask we selected all
the galaxies in the colour–magnitude diagram lower than
V − I = 1.6 and less than I = 22. This isolated most of
the cluster sequence. The remaining galaxies in the region
V − I < 1.6 and I > 22, V < 26.8, we identified as the faint
blue background population. It is clear from figure 2 that
the distinction between faint cluster member and faint blue
background galaxy is rather vague. However, since the faint
cluster members are also the smallest, the masked area is
fairly insensitive to the exact division. Figure 3 shows the
distribution of cluster galaxies and the red background pop-
ulation. The concentric circles are centered on the peak in
the cluster light distribution and show the position of the
annuli used to calculate the radial profile in section 5.4.
Figure 3. The masked region of A1689. Cluster members where
selected using colour information (see text) and then masked over
so that these regions do not affect the surface density estimate
of background sources. The total region masked is about 10% of
the area. The background galaxies are also shown as open circles.
Superimposed are the concentric bins used to calculate the radial
profile, centered on the peak in the light distribution. North is up
and east is to the left.
5.1.3 Selection by colour
Once the cluster galaxies have been isolated, the background
galaxies may be sub-divided into a red and blue population,
separated by V − I = 1.6. The observed slope of the lu-
minosity function for these two populations for I > 20 is
βR = 0.38 and βB = 1 (Broadhurst 1995; we shall do a
more accurate fit using our colour cuts in section 5.2.2).
From equation (1) we expect that the surface density of red
galaxies will be suppressed due to the dilation effect, while
magnification of the faint blue galaxy population will com-
pensate for the dilation. Hence selecting by colour allows us
to identify a population of galaxies with a very flat luminos-
ity function to boost the lensing signal, at the expense of
a reduction in galaxy numbers. Simple error analysis shows
that the signal–to–noise varies as (Taylor & Dye 1998)
S/N = 2β − 1κA(1 − κ + γ ′/κ′ )√n(1 + nσ 2 )−1/2 ,
(20)
where ′ ≡ ∂ /∂R. While the signal–to–noise is a linear func-
tion of the slope of the luminosity function, it only grows
with the square–root of the galaxy numbers, assuming Pois-
son statistics. Hence one can get a better signal-to-noise by
pre-selection of the red background population to boost the
signal, at the expense of numbers. Equation (20) also shows
that one can get a better signal by observing to fainter mag-
nitudes to simultaneously enhance the surface number den-
sity and reduce the contribution from intrinsic clustering
(see Taylor & Dye, 1998, for a more detailed discussion of
observing strategies).
There is also a practical reason for favouring the red
galaxy population. While the cluster members are unlikely
to be redder than the cluster E/SO sequence the distinc-
tion between faint blue galaxies and cluster members, based
on selection from the colour–magnitude diagram alone, is
6
A.N. Taylor et al.
somewhat vague. There may be blue cluster members that
will contaminate the sample of blue background galaxies.
In the absence of redshift information the blue background
population is clearly harder to isolate.
As we have noted the red population has relatively few
faint counts, so that the expansion term in equation (1) dom-
inates and there is a net underdensity of red galaxies behind
the cluster (see Fig. 4 and Fig. 7). Conversely, faint blue
galaxies are numerous and cancel the expansion effect. As
expected we found the blue galaxies were uniform across the
A1689 field. This is a good indicator that it is the magnifi-
cation effect at work, and not some spurious contaminant,
for example colour gradients across the field, or large scale
variations due to clustering. In addition it also indicates that
the deficit in the red population is not due to dust obscu-
ration or reddening in the cluster, as this would affect both
red and blue populations in equal measure.
5.2 The distribution of background galaxies
In Figure 4 we show the surface distribution of the red popu-
lation behind A1689, Gaussian smoothed on a scale of 0.35′ .
There are 268 background galaxies. The cluster members
have been masked out and the masked areas interpolated
over. The masked region contributes to only ≈ 10% of the
total field. Figure 3 shows the masked region. The cluster
center, identified as the peak of the light distribution, is at
(4.1′ , 3.6′ ).
The angular size of the cluster scales like
R(θ) = 0.87DA (zc )(θ/1′ )h−1Mpc,
(21)
where DA (z ) = 2(1 − (1 + z )−1/2 )/(1 + z ) is the comoving,
dimensionless angular distance in an Einstein–de-Sitter uni-
verse. Hence at the redshift of Abell 1689, zc = 0.183 ± 0.001
(Teague et al 1990), one arcminute is about 0.117h−1Mpc.
Figure 4 clearly shows a deficit of galaxies about the
central peak in the light distribution at (4.1′ , 3.6′ ). At θ =
0.75′ there is an arc of very underdense number counts to
the South–West of the cluster center, marked by a dashed
line (The background is somewhat obscured by the cluster
mask to the North–East of the cluster center). This is clear
indication of a caustic feature in the background number
counts, where the number density drops to zero due to di-
lation. This exactly corresponds to the radius of the blue
arcs observed by Tyson & Fisher (1995) at θ = 0.85′ (see
also the radial number counts in section 5.4). This is strong
evidence that we have detected the magnification effect in
the background counts.
5.2.1 The redshift distribution of background galaxies
The efficiency of lensing varies with the redshift of the back-
ground source (BTP). Therefore it is important to esti-
mate the background redshift distribution. Crampton et al
(1995) find that Canada–France Redshift Survey (CFRS)
has a median redshift of z = 0.56 for galaxies in the range
17.5 < I < 22.5. They also show a colour–redshift diagram
that indicates that the red galaxy population (V − I > 1.6)
has a median redshift of about z ≈ 0.8 (Crampton et
al 1995, their figure 5). More accurately we can integrate
the best fit Schechter function found by Lilly et al. (1995)
Figure 4. The distribution of red I-band background sources for
Abell 1689. Darker grey areas indicate an underdensity of source
counts. The image is Gaussian smoothed with a smoothing scale
of 0.35′ . The peak of the light distribution is at (4.1′ , 3.6′ ). The
maximum density of ob jects is 23.0 per square arcminute and the
minimum is 1.1 per square arcminute. There are 15 contour lines
spaced by ∆n = 1.46 galaxies per square arcminute. A strong
caustic feature is seen 0.75′ from the peak (inner dashed line),
more visible to the south–west, as the other side of the peak is
masked over. A second feature is found in the radial profile at
2.2′ (outer dotted line). The image is orientated with East to the
left and North to the top.
for the CFRS red galaxy population. This has parameters
φ∗ = 0.0031 ± 0.00095, α = 1.03 ± 0.15 and M ∗ (B ) = −21,
where M (B ) = I − 5 log10 (DL /10pc) + 2.5 log10 (1 + z )+k-
correction, where the k-correction is discussed in their paper
and DL = (1 + z )r(z ) is the luminosity distance. Lilly et al.
found no detectable evolution of the luminosity function of
the CFRS red population and we assume no evolution. Ex-
trapolating to the magnitude range 20 < I < 24 we find that
the redshift distribution can be well fitted by the function
αz 2
z 3
∗ Γ[3/α]
φ(z ) =
exp(−(z/z∗ )α )
with α = 1.8 and z∗ = 0.78 to about 5% accuracy over
a redshift range of 0.25 < z < 1.5. The moments of this
distribution are
(22)
.
(23)
Γ[(3 + n)/α]
hzn i = zn
∗
Γ[3/α]
Hence for the red galaxy population we find that hz i = 0.96
and σz = 0.42. To simplify the analysis of the lensing proper-
ties of the cluster, we shall assume hereafter that the back-
ground distribution is at a single redshift of z = 0.8 (the
mode of the distribution) and has an uncertainty of δz = 0.4.
As the caustic indicated by the blue arcs coincides with
the magnification caustic, we can presume that the galaxy
forming the arcs lies at the same redshift as the magnified
red background galaxies, z ≈ 0.8. At present we do not know
the redshift of this arc.
Gravitational lens magnification
7
that log10 n(blue) ≈ 0.35I − 3.49, resulting in βB = 0.88,
close to the lens invariant β = 1, and a count density be-
tween 23 < I < 24 of n0 (blue) = 15.5 galaxies per square
arcminute.
5.2.3 Clustering properties of the background population
The amplitude of clustering of I-band galaxies and its de-
pendence on redshift can be estimated from the CFRS (Le
F´evre et al. 1996). Le F´evre (1996) find that there is little
difference between the clustering properties of red and blue
populations of galaxies for z > 0.5, implying that the pop-
ulations were well mixed at this epoch. We therefore apply
their clustering results directly to our red galaxy population.
They fit their results to a power–law model for the evolving
correlation function, ξ (r) = (r/r0 )−γ , where
r0 (z ) = r0 (0)(1 + z )−(3+ε)/γ
(26)
where ε = 1 ± 1 and r0 (z = 0.53) = 1.33 ± 0.09h−1Mpc and
γ = 1.64 ± 0.05 is in this section the slope of the correlation
function.
The quantity we require is the variance in a given area
of sky, which can be estimated by averaging the observed an-
gular correlation function, ω (θ), over a given area (equation
7). The clustering variance for I-band galaxies then scales
roughly like (Taylor & Dye 1998)
nl = 10−2 z−2.8 (θ/1′ )−0.8
σ 2
(27)
where the sampled area is a circle of radius θ and we have
assumed unbiased, linear evolution of the density field. The
background galaxies are assumed to all lie at z ≈ 1.
5.3 Reconstructing the surface mass density
In Figure 6 we plot the reconstructed surface mass density
of Abell 1689 using the nonlinear local sheet approximation,
κ0 (see section 4.1), changing parity on the caustic line at
θ = 0.75′ (see Figure 4). The uncertainty on the peak of
the mass distribution is somewhat large (see Section 5.2),
but significant features can be seen around the cluster core.
There appears to be an extension to the south–west not seen
in the cluster galaxy distribution. Interestingly there also ap-
pears to be a loosely connected ridge, about 2.4′ from the
peak. We shall discuss this feature further below, but note
that the shear mass map derived by Kaiser (1996; figure 2)
shows similar extensions and ridge, although the extension
to the west is not apparent in the shear map. Two under-
dense regions are also seen to the south and to the east in
both maps. While the comparison is only qualitative and the
maps are noisy, we find this very encouraging as these maps
are derived from completely independent methods, although
the underlying data set is the same.
5.4 The Mass Profile of Abell 1689
While the mass maps are suggestive, a more quantitative
measure can be made by angle averaging the counts and cal-
culating the mass profile. Figure 7 shows the radial counts
about the peak in the light distribution, normalised to the
Keck data. The plotted error bars are only due to Pois-
son statistics, although in the mass analysis below we shall
Figure 5. Magnitude distribution of all I-band galaxies (solid
dots), the red selected galaxies (grey dots) and the blue back-
ground galaxies (open dots). The lines are the best fits to the
data.
5.2.2 Number counts of the background galaxy population
Of ma jor importance to the lens magnification method is
the normalisation of the background galaxy population. The
CFRS is not adequate for this, since their colour cuts were
in the rest frame U − V , rather than the observed V − I .
Instead we have used the Keck data of Smail et al (1995),
who observed deep V RI images down to a limiting magni-
tude of R ≈ 27. The total differential galaxy count rate in
the I-band can be approximated by
(25)
(24)
log10 n = (0.271 ± 0.009)I − 1.45
over the range 20 < I < 24, where n is per magnitude
per square degree. We have applied our colour criteria (see
section 5.1) to the Keck data and find that the red galaxy
population V − I > 1.6, can be well approximated by
log10 n(red) = (0.0864 ± 0.0187)I + (2.12 ± 0.41)
over the range 20 < I < 24. Figure 5 shows the mag-
nitude distribution for the full dataset and for the red–
selected galaxy population and the best–fit lines. Integrat-
ing the fit for the red galaxies yields a total count rate of
n = 12.02 ± 3.37 galaxies per square arcminute in the range
20 < I < 24. Since β = 2.5 d log10 n/dm, we find that the
Keck data implies βR = 0.216 ± 0.047. This is the value of
β we shall use in the subsequent analysis.
An alternative, although less exact, method of normal-
isation is to assume negligible cluster mass at the edge of
the field and normalise the cluster to this. In general this
would put a lower limit on the mass, and is similar to the
method used to normalise shear mass maps. In fact if we do
this for A1689 we find a background count rate very sim-
ilar to that given by the Keck data. The error introduced
into the final mass estimate by uncertainties in β scale as
δκ/κ ≈ δβ/1 − β , which for the Keck data results in a
fractional error of around 5%.
We have also fitted the blue counts in the Keck sample
(Figure 5). Over the same range as the red counts we find
8
A.N. Taylor et al.
Figure 6. Reconstruction of the surface mass density of Abell
1689 from the red background galaxy population, using the non-
linear local sheet approximation and a full likelihood analysis in
2-D. Light regions are high density. Only one caustic line is as-
sumed, at θ = 0.75′ from the peak of the light distribution. The
maximum surface density is κ = 1.35, at (4.02′ , 3.41′ ), consistent
with the peak in the light distribution. The minimum surface
mass density is κ = −0.47. There are 15 linearly spaced contours,
separated by ∆κ = 0.12 and the map is Gaussian smoothed with
a smoothing length of θS = 0.35′ . North is up and east is to the
left.
r(arcmin) N
N/N0 Annulus area Obscured area
0.33
0.67
1.01
1.35
1.69
2.03
2.36
2.70
3.04
3.38
3.72
4.06
2
0
4
13
20
23
11
26
32
34
45
31
1.19
0.00
0.24
0.51
0.59
0.57
0.28
0.77
0.94
0.95
1.25
1.06
0.35
1.08
1.79
2.51
3.23
3.62
3.46
3.08
3.01
3.14
3.18
2.49
0.21
0.25
0.40
0.39
0.40
0.26
0.23
0.26
0.16
0.17
0.20
0.06
Table 1. Table of angular radius (r in arcminutes), number of red
galaxies (N ), ratio of galaxies to background (N/N0 ), the total
area of the annuli (areas are in arcmin2 ) and area obscured by
the mask. The unobscured area is total area − obscured area. The
expected number of galaxies in an annuli is N0 = n0× unobscured
area
be discussed in more detail in section 5.4.4. The increase in
counts at θ = 3.7′ is likely to be due to a clustering effect.
Figure 7. Radial profile of red counts behind Abell 1689. The
background count density is n0 = 12 ob jects/arcmin2 . Superim-
posed is the profile for an isothermal model, normalised at the
caustic radius, θ = 0.75′ (dashed line).
take into account the effects of clustering. A general trend
is clear, and lies close to the prediction for an isothermal
lens normalised to the blue arc caustic. This has a surface
mass density of κ = 0.375(θ/1′ )−1 , corresponding to a virial
velocity of 1600 kms−1 . Again it is worth emphasizing that
the zero of the number counts at θ = 0.75′ corresponds to
the caustic inferred from the blue arcs. The second dip will
Figure 8. Radial profile of blue counts behind Abell 1689. The
background count density is n0 = 22 ob jects/arcmin2 . As ex-
pected from the nearly lens invariant slope βblue = 0.88, the
number counts are nearly flat, and at large radii tend towards
n/n0 = 1. The slight increase towards the cluster center is prob-
ably due to contamination of the counts by blue cluster members
In Figure 8 we show that radial profile for the blue
galaxy population, normalised with the Keck data in sec-
tion 5.2.2. As expected there is no lensing signal. The slight
increase towards the cluster center is due to contamination
from the blue cluster members.
5.4.1 Local approximations for the surface mass density
Fig. 9 shows the radial mass profile of the cluster Abell 1689
assuming a single caustic at θ = 0.75′ . The two solid lines
Gravitational lens magnification
9
Figure 9. Radial profile of surface mass density of cluster Abell
1689. The dark solid region shows the uncertainty due to the
strong lensing estimators. The lighter shaded region is due to the
clustering and shot-noise uncertainty of the background popula-
tion. The solid line is a singular isothermal profile, normalised to
the caustic feature at θ = 0.75′ .
Figure 10. Radial profiles of surface mass density, κ, for A1689
(solid line with dots), calculated by solving the axially symmetric
lens equation (19). The shaded regions are 1σ errors calculated via
error propagation from the uncertainty on the measured amplifi-
cation profile. The solid dark line is a singular isothermal profile
normalised to the caustic feature at θ = 0.75′ . The lighter solid
line is the local best fit estimator, κc
are calculated using the Lognormal–Poisson likelihood esti-
mator (equation 6) with each of the two strong lensing ap-
proximations (equations 9 and 8 ). The light shaded region
indicates the 1σ uncertainty due to both shot noise and the
effects of clustering. The dark shaded region indicates the
region between the two extreme estimators. Away from the
cluster center these agree and are equal to the weak lensing
estimator, but noise effects become dominant. Closer to the
cluster center the uncertainty due to the shear increases and
becomes dominant at θ < 1′ . However the cluster mass pro-
file is significantly detected between 1′ < θ < 2.6′ . We also
appear to see a deviation from an isothermal profile, which
is also plotted. When the procedure was repeated with the
center of the annuli offset from the peak of the light distribu-
tion the mass profile was weaker and less significant, as one
would expect if the peak of the mass density was associated
with that of light.
5.4.2 Nonlocal approximation for the mass density and
shear
In Figures 10 and 11 we assume axisymmetry and equation
(19) to calculate the surface mass–density and shear simul-
taneously. We set γ1 = 0.3 for the first shell. The resulting
profile is fairly insensitive to this choice, only affecting the
first 2 shells. The uncertainty on the shear in the first shell
is small because this must be chosen a priori. However av-
eraging over shells means that the errors do not strongly
propagate through to higher radii. Again a mass detection
is found between 1′ and 2.8′ , this time with the shear ac-
counted for. In this region κ ≈ 0.4 ± 0.15, which is somewhat
higher than that found by the shear estimate of κ = 0.2± 0.1
(Kaiser 1996; note that we quote Kaiser’s colour selected
sample, where cluster members that may contaminate the
shear estimate have been removed. This corresponds to com-
Figure 11. Radial profiles of tangential shear, γt , for A1689 (solid
line with dots), calculated by solving the axially symmetric lens
equation (19). The shaded regions are 1σ errors calculated via
error propagation from the uncertainty on the measured amplifi-
cation profile. The solid dark line is a singular isothermal profile
normalised to the caustic feature at θ = 0.75′ . The lighter solid
line is the local best fit estimator, κc .
10
A.N. Taylor et al.
bining our red and blue background populations. This will
change the redshift distribution of the background and in-
clude some residual blue cluster contamination which may
account for the discrepancy). Also, for the single caustic so-
lution, we see a large spike at 2.2′ , which is not seen in the
Kaiser (1996) results. However the shear method correlates
points, which may lead to both the suppression of features,
and underestimation of the errors.
Our estimate of the shear field is far more uncertain,
with γt = 0.2 ± 0.3 over most of the range. There is a slight
increase beyond 2.4′ due to the spike in the surface mass pro-
file at that radius, but the profile is dominated by noise. This
increase is not reflected in the angle averaged measurements
of Kaiser (1996), where the mean shear is γ = 0.15 ± 0.05.
5.4.3 Local approximation for the surface mass density
and shear
Figures 10 and 11 also show κ and γ estimated from the
best-fit solution of section 4.1. We find good agreement be-
tween the local and nonlocal approximations for κ, but the
the shear profiles are somewhat different, reflecting that one
estimator is local and one nonlocal. However, the large un-
certainties produced by each estimator means that we can-
not predict the shear profile with much certainty from the
available data.
5.4.4 Two background populations ?
An interesting feature of the counts in Figure 7 is the ap-
pearance of two pronounced dips, one at 0.75′ and another
one at 2.2′ . While the inner dip has already been identified
with a caustic line, the outer dip is somewhat anomalous. A
number of possibilities could account for this. The feature
was noted in the mass plot as a low signal-to-noise ridge in
the density and can be seen in the number counts as a loosely
connected ring about the cluster center. One possibility is
that this is due to clustering in the background population,
combined with a large mass concentration to the south–east
of the peak in the light distribution. There are few cluster
members in the region of the ridge, or the bump, so the
effect is not due to masking.
An alternative is that this is the first glimpse of a second
caustic line. In principle a second caustic can be created by
placing the background galaxies at two redshifts, one at low
redshift, one at high redshift (eg, Fort, Mellier & Dantel-Fort
1996). The observed number counts would then be given by
2 − Aβ−1
1 + ν (cid:0)Aβ−1
n′ /n0 = Aβ−1
(cid:1)
(28)
1
where Ai = A(fi ) with fi = κ(zi )/κ∞ = (√1 + zi −
√1 + zL )/(√1 + zi − 1) (BTP), and i = 1, 2 for the two
galaxy populations. ν is the fraction of galaxies at redshift
z2 . An outer caustic line must be produced by the high red-
shift population. If we make this population lie at z = 0.8,
then the low redshift population must lie at z = 0.3. Both
populations are reflecting the same arc, the difference in
pro jected radii is wholly due to their relative redshifts.
However this would double the predicted mass from lens
magnification, making Abell 1689 a very extreme cluster.
In addition it seems hard to make a caustic line from the
high redshift population for such a massive cluster without
forming a second, inner radial caustic. As the strongest arc
is tangential and is seen near the inner arc, one would have
to conspire to have a nearby galaxy, at z = 0.3, lensed and
lying at the same pro jected radii as the radial arc produced
by the high redshift population. This seems highly unlikely.
One could also keep the mass roughly constant and
place a second population at z > 0.8. This is a possibil-
ity, but does not strongly affect our mass estimate assuming
a single caustic solution. In the absence of further evidence
for a second high redshift population, we shall only consider
the single caustic model.
5.5 Mass estimate of Abell 1689
5.5.1 From κ to mass surface density
Σ =
Σ0 κ
Assuming that the background galaxies all lie at the same
redshift of z = 0.8, and given that the surface density scales
as
(√1 + z − 1)(1 + zL )2
1
(√1 + zL − 1)(√1 + z − √1 + zL )
3
where Σ0 = 8.32 × 1014 hM⊙Mpc−2 is the mean mass per
unit area in the universe, then we find that the surface mass
density is
Σ = 5.9 × 1015 κ [hM⊙Mpc−2 ].
Although we have assumed an Einstein–de-Sitter universe,
these results only depend weakly on cosmology (BTP).
,
(29)
(30)
5.5.2 Uncertainty in the redshift distribution
The error introduced by assuming the background galaxies
lie at the same redshift can be estimated by error propa-
gation and assuming δz = 0.4 (see section 5.2.1). Hence
δΣ = ∂Σ/∂ z δz and the fractional uncertainty on the sur-
face mass density due to the uncertainty in redshift distribu-
tion of the background galaxies is δΣ/Σ = 0.37δz = 0.148.
The same error is also found in mass estimates based on the
shear pattern.
5.5.3 Uncertainty due to normalisation of background
counts
Assuming a sheet mass solution (κ0 in section 4.1) we find
that the uncertainty due to the normalisation of the back-
ground counts is δκ = (1 − κ/2β − 1)δn0 /n0 . For A1689
and the red galaxy population this is δκ = 0.151 − κ. For
an average κ = 0.5 the uncertainty is around δκ = 0.07.
5.5.4 The cumulative mass distribution
Figure 12 shows the cumulative mass interior to a shell, cal-
culated from both the nonlocal approximation (section 4.2),
and the best–fit local approximation allowing only a single
caustic solution (section 4.1). The uncertainties are treated
by error propagation. We find that the 2d pro jected mass
interior to 0.24h−1Mpc is
M2d (< 0.24h−1Mpc) = (0.50 ± 0.09) × 1015 h−1M⊙ ,
(31)
Gravitational lens magnification
11
2d lens mass to other cluster characteristics, such as the
line–of–sight velocity dispersion (section 6.3) and the X-ray
temperature (section 6.4). Discrepancies that arise between
these predicted characteristics and the actual measurements
can be used to infer information about the mass distribution
along the line of sight (Bartelmann & Kolatt, 1997). We find
that while there is fair agreement between all of the mass
estimates when pro jection effects are taken into account, the
agreement is better if the cluster A1689 is composed of two
clusters superimposed along the line of sight and separated
by about ∆z = 0.02.
The transformation from a 2-d pro jected lensing mass
to a 3-d mass, line–of–sight velocity dispersion and X-ray
temperature can be made using either the isothermal model,
or by using relations found in N-body simulations of clusters.
While the former provides a simpler method, one has more
freedom with simulations to include or exclude the various
pro jection effects that contaminate measurements of these
quantities. In this section we shall use the relations found
by van Kampen (in preparation) from an ensemble of CDM
cluster simulations, all with Ω0 = 1 and σ8 = 0.54. These
relations are model dependent, but serve to aid comparison
between the various mass measurements. We have also pro-
vided a table of quantities (Table 1) where the uncertainties
have been calculated by combining the error on the cluster
mass with the dispersion found in the depro jection relations.
We begin by comparing the lens magnification mass
with the mass determined from the shear field around
A1689.
6.1 Comparison with arclets and weak shear
(33)
Tyson & Fischer (1995) provide mass profiles of A1689 from
arclets, another independent estimator of the mass, normal-
ized to the caustic line indicated by the blue arcs. They find
that the 2d pro jected mass within R = 0.1h−1Mpc is
M2d (< 0.1) = (0.18 ± 0.01) × 1015 h−1M⊙ .
They also find that the mass scales like an isothermal sphere
out to 0.4h−1Mpc, before turning over to an R−1.4 profile.
This implies that in the regime we probe with the magnifi-
cation the cumulative mass scales like
M (< R) = (1.8 ± 0.1) × 1015 (R/h−1Mpc)h−1M⊙ .
This is very close to the profile we find from lens magnifica-
tion (equation 32). Using this we scale their results giving
M2d (< 0.24) = (0.43 ± 0.02) × 1015 h−1M⊙ ,
in good agreement with the mass from magnification.
Kaiser (1996) has also calculated κ based on the weak
shear method (Kaiser & Squires, 1993), using the same data
we have used here for A1689. We noted above that there
are qualitative similarities between the weak shear maps
and those presented by Kaiser, which is significant since the
methods are independent. The mass–density profile found
from the shear pattern is also well fitted by an isothermal
profile;
M2d (< R) = 1.8 × 1015 (R/h−1Mpc)h−1M⊙ ,
with a 10% statistical uncertainty and further 10% system-
atic error due to the uncertainty in the redshift distribution
(36)
(34)
(35)
Figure 12. Cumulative mass profile of Abell 1689. The solid
dark line and shaded uncertainties are estimated using the axi-
symmetric nonlocal estimator described in section 4.2. The lighter
grey line is the cumulative mass estimated from the best–fit local
approximation, κc described in section 4.1. Also plotted is the
isothermal fit to the blue arc caustic (dotted line), similar to the
shear results of Kaiser (1996) and Tyson & Fischer (1995).
and that the two estimators are in good agreement. We find
that the pro jected mass scales like
M2d (< R) ≈ 3.5 × 1015 (R/h−1Mpc)1.3 h−1M⊙ ,
for R < 0.32h−1Mpc, similar to that for an isothermal
sphere; M ∼ R. Hence it appears that A1689 has a near–
isothermal core. Beyond R = 0.32h−1Mpc the lensing signal
is lost in background noise, and we can only say that κ ≤ 0.1
Including the uncertainty from the background redshift
distribution and the normalisation of background counts in-
creases the error to about 30%.
(32)
6 COMPARISON WITH OTHER MASS
ESTIMATES OF A1689 AND INFERRING
THE 3D MASS DISTRIBUTION
In this section we compare the mass derived from lens mag-
nification with that found from a number of other, indepen-
dent measurements. Firstly we compare our results with esti-
mates of the mass based on the shear pattern found around
A1689 (section 6.1). The magnification and shear comple-
ment each other in that the shear pattern has a higher
signal–to–noise, since it is not affected by clustering noise
(although with redshift information the magnification can
also be measured free from clustering noise – see BTP), but
suffers from the “sheet–mass” degeneracy. We shall combine
the magnification and shear pattern elsewhere.
While the lens magnification mass is vital for fixing the
total 2d pro jected mass distribution independently of any
assumptions about the dynamical state of the cluster, much
information can be gained by combining this with other
mass estimates, assuming that these are not strongly bi-
ased by their reliance on thermodynamical equilibrium. In
this section we describe a method for transforming from the
12
A.N. Taylor et al.
(section 5.5.2). Compared with our 2-d mass Kaiser’s anal-
ysis suggests that
M2d (< 0.28) = (0.43 ± 0.04) × 1015 h−1M⊙ ,
again in good agreement with that found by the magnifica-
tion method.
(37)
6.2 The 3d mass estimated from lensing alone
(38)
The 3-d mass inferred from the 2-d pro jected mass inside a
sphere of radius r = 0.5h−1Mpc is
M3d (< 0.5) = (0.72 ± 0.25) × 1015 h−1M⊙ ,
while the mass inside an Abell radius, r = 1.5h−1Mpc, is
M3d (< 1.5) = (1.6 ± 0.6) × 1015 h−1M⊙ .
These estimates are probably an overestimate of the true
3d mass since the dispersion in the simulations includes the
effect of the alignment of the clusters’ principle axis along
the line of sight. Given that the inferred 3d mass is so high
A1689 is probably lying at the extreme of such a distribu-
tion. In such cases the 3d mass may be much lower than mass
inferred from a 2d pro jection. We discuss this possibility in
the next few sections.
(39)
6.3 Velocity dispersion of Abell 1689
(40)
The predicted line–of–sight velocity dispersion estimated
from the simulations also includes the effects of superposi-
tion of clusters, infall along filaments and interlopers and so
predict larger velocities than isolated clusters would. Given
this we find that the observed line–of–sight velocity disper-
sion inferred from the 2d lensing mass is
σv (< 1.5h−1Mpc) = 3400 ± 900 kms−1
for A1689. Again this is very much on the high side for a
cluster in a typical CDM universe, and much higher than
observational data suggests.
The velocity dispersion measured in a similar manner
by Teague et al (1990) is 2355+238
−183 kms−1 , nearly 1σ lower
than our estimate. Hence the predicted velocity is not too
different from the measured value, so long as both are mea-
sured in the same way. The large mass implied by lensing
suggests that A1689 is not a single cluster, but a superposi-
tion of two smaller clumps. Miralde-Escud´e & Babul (1995)
have modeled A1689 by two adjacent clumps with velocity
dispersions 1450 kms−1 and 700 kms−1 , suggesting that one
clump has a mass 4 times that of the other. Taken together
and including the relative velocity of the merging clumps,
this accounts for the larger velocities measured earlier. den
Hartog & Katgert (1996) have tried to take into account
interlopers and find that v = 1861 kms−1 .
If, like Miralde-Escud´e & Babul, we assume a double
cluster model but place one cluster at z = 0.18 with a ve-
locity dispersion of 1500 kms−1 and the second at z = 0.20
with a velocity dispersion of 750 kms−1 , we find that we can
reproduce a total pro jected velocity dispersion of around
2300 kms−1 . Figures 4 and 5 of Teague et al (1990) also pro-
vide marginal evidence for a second concentration of galaxies
at z = 0.2. Hence it seems plausible that the high lensing
mass of A1689 can be explained by a superposition of clus-
ters.
Quantity
This work
Other
M2d (< 0.24)
0.50 ± 0.09
M3d (< 0.5)
M500
σv (< 1.5)
0.72 ± 0.25
1.6 ± 0.65
3400 ± 900
0.43 ± 0.02
0.43 ± 0.04
(Tyson & Fischer)
(Kaiser)
0.94 ± 0.16
2355+238
−183
(Yamashita)
(Teague et al)
Table 2. Mass estimates for A1689 based on lens magnification
(second column) and from other measurements (third column).
Masses are given in units of 1015 h−1M⊙ and velocities are quoted
in units of kms−1 . Distance are given in h−1Mpc. The other mea-
surements are based on arclets (Tyson & Fischer 1995), the shear
pattern (Kaiser 1996), X-ray temperatures (Yamashita 1994) and
line–of–sight velocity dispersion (Teague et al 1990). Also given
are the 3d mass estimates from lens magnification.
6.4 X–ray mass estimates of Abell 1689
(41)
h−1M⊙ ,
Evrard, Metzler & Navarro (1996) have found that the mass
within the radius defined where the mean cluster density is
500 times the critical density is strongly correlated with the
cluster temperature. They fit this relation from simulations
by
10keV (cid:17)3/2
M500 = 1.11 × 1015 (cid:16) TX
where TX is the broad–beam temperature and M500 is the
3d mass within a radius defined by an overdensity 500ρcrit .
This radius is roughly given by r500 = 1.175h−1Mpc.
X-ray temperatures of A1689 have been measured by
both GINGA and ACSA. Yamashita (1994) has analysed
this data and finds T = 9 ± 1keV while Mushotzky and
Scharf (1997) find T = 9.02+0.4
−0.3 keV. Daines et al (1997)
have also recently re-analyzed ROSAT PSPC observations
and find a mean temperature of TX = 10.2±4keV. Note that
we are quoting the mean temperature, and incorporated the
40% uncertainty in the error estimate, rather than quot-
ing upper limits as Daines et al do. The ma jor uncertainty
in measuring X-ray temperatures here is instrumental, as
10keV is approaching the limit of ROSAT’s sensitivity.
Taking the result of Yamashita and the relation found
by Evrard et al, we find that
M500 = (0.95 ± 0.16) × 1015 h−1M⊙ .
Using the simulated scaling relations we find
M500 = (1.6 ± 0.65) × 1015 h−1M⊙
for Abell 1689, implying an X-ray temperature of TX =
12.7 ± 3.4keV, within the 1σ uncertainty of the measured
X-ray temperature. Again, if we consider A1689 as a double
cluster, the nearer, larger mass concentration would be de-
tected in X-ray, lowering the expected X-ray temperature.
From the velocity dispersions we can infer a temperature
nearer to TX = 0.7keV, slightly below, but again in agree-
ment with observations.
In conclusion, although we find a high mass, there is a
general consistency between the mass of A1689 estimated
from lens magnification and shear. In addition we find a
fair agreement between the lens mass and the line of sight
(42)
(43)
velocity dispersion if we take into account pro jection effects.
Modelling A1689 as a double cluster we find that the velocity
dispersion can be much lower, implying two smaller clusters,
with the lensing mass a superposition of cluster masses. This
hypothesis might also help explain the marginal discrepancy
with X-ray temperature.
However, if A1689 is a double cluster, one would expect
that the measured velocity dispersion would be higher than
that inferred by the lensing mass, due to the cluster sepa-
ration, while the X-ray mass is lower, since only the more
massive cluster dominates the X-ray emission. Our results
indicate that Abell 1689 has a larger lensing mass than that
implied by both velocity dispersion and X-ray temperature,
although given the large uncertainties it is hard to be con-
clusive.
Finally, A1689 is in pro jection a highly spherical clus-
ter, in contrast with the ma jority of clusters which appear
extended. While this may be be a result of its high mass, it is
also possible that A1689 has its ma jor axis aligned along the
line of sight, pointing towards a second cluster. While much
of the evidence on the mass distribution along the line–of–
sight is circumstantial, all of these effects would conspire to
give A1689 its impressively massive appearance.
7 DISCUSSION
The absolute surface mass density of a galaxy cluster can
be estimated from the magnification effect on a background
population of galaxies, breaking the “sheet–mass” degener-
acy. To apply this in practice, we have taken into account
the nonlinear clustering of the background population and
shot–noise, both of which contribute to uncertainties in the
lensing signal. A further complication is the contribution
of shear to the magnification in the strong lensing regime,
where the magnification signal is stronger. We have argued
that this can be circumvented by approximate methods that
can be either local, where a relationship between surface
mass and shear is assumed, or by a nonlocal approxima-
tion where only the shape of the cluster is assumed. Both
approximations seem to work well on simulated data.
We have applied these methods to the lensing cluster
Abell 1689, using Keck data of Smail et al (1995) to nor-
malise the background counts, and the CFRS results to in-
fer the redshift distribution and clustering properties of our
data. Using a γ = 0 approximation of the surface density in
the strong lensing regime, we have reconstructed a 2-d mass
map for A1689 in the innermost 27 square arcminutes, where
a substantial part of the lensing signal comes from. The 2d
map has general features similar to those seen from shear
maps (Kaiser 1996). This is encouraging for both methods,
as they are independent determinations of the mass distri-
bution.
For a more quantitative measure we have binned the
data in annuli around the peak in the light distribution and
found a significant (5σ ) drop in the number counts, drop-
ping to zero where a caustic is inferred from arcs. Local and
nonlocal approximations were used to find the κ profile from
the number counts, and estimate the shear field. We found
these to be quantitatively similar to that found by the shear
method.
We have also discussed the possibility of a second pop-
Gravitational lens magnification
13
(44)
ulation of background galaxies, creating a second dip in the
radial number counts and a spike in the mass profile. How-
ever we argued that is is unlikely that there is a second low-z
population as the cluster mass would be improbably high,
and if there is a high-z population it has little effect on our
results.
We have calculated a cumulative mass profile for A1689
and find a pro jected 2-d cumulative mass of
M2d (< 0.24h−1Mpc) = (0.50 ± 0.09) × 1015 h−1M⊙ .
Such a large mass is very rare in a CDM universe normalised
to the observed cluster abundance, and may indicate that
A1689 is composed of two large masses along the line of
sight, and/or filaments connected to the cluster and aligned
along the line–of–sight. This is also implied by the high
line of sight velocity dispersion which would be enhanced
by merging clusters (Miralde-Escud´e & Babul 1995) or by
infall from aligned filaments.
We have compared our mass estimates with other esti-
mates available in the literature and find that the lens mag-
nification, shear, arclets, line–of–sight velocity dispersions
and the X-ray temperature mass estimates are all in reason-
able agreement, to within the uncertainties at this time.
The results presented here are from 3 hrs integration on
the 3.6m NTT. Longer integration times have the combined
benefit of increasing the number of background galaxies, and
so reducing shot noise, and of reducing the contribution from
cosmic variance (equation 20, Section 5.1.3). Hence by in-
creasing the exposure time we can expect to reduce the un-
certainty from lens magnification by a factor of 2 or so.
One drawback of this analysis is the contribution of
clustering noise to the background counts. This can be re-
moved using redshift information, either from spectroscopy,
or more efficiently using photometric redshift information
(BTP). We shall explore this elsewhere (Dye et al, in prepa-
ration).
If our results are extended to other clusters we can hope
to have a good representation of the total mass distribution,
gas and galaxy contents with which to make strong statisti-
cal arguments about the matter content of the largest grav-
itationally collapsed structures in the universe.
Acknowledgments
ANT thanks the PPARC for a research associateship and
the University of Berkeley and the Theoretical Astrophysics
Center, Copenhagen for their hospitality during the writ-
ing of this paper. EvK acknowledges an European Commu-
nity Research Fellowship as part of the HCM programme
and thanks the University of Edinburgh and ROE for their
hospitality. This work was supported in part by Danmarks
Grundforskningsfond through its funding of the Theoreti-
cal Astrophysics Center. SD thanks the PPARC for a stu-
dentship. NBL thanks the Spanish MEC for a Ph.D. schol-
arship, the University of Berkeley for their hospitality and
financial support from the Spanish DGES, pro ject PB95-
0041. We thank Ian Smail who kindly provided us with a
copy of his Keck data. We also thank John Peacock, Alan
Heavens, Jens Hjorth, Adrian Webster and an anonymous
referee for comments and useful suggestions.
14
A.N. Taylor et al.
REFERENCES
Bartelmann M., Steinmetz M., Weiss A., 1995, AA, 297, 1
Bartelmann M., Kolatt T.S., 1997, MNRAS, submitted,
astro-ph/9706184
Broadhurst T.J., Taylor A.N., Peacock J.A., 1995, ApJ, 438,
49
Broadhurst T.J., 1995, in AIP Conf. Proc. 336, “Dark Mat-
ter”, eds S.S. Holt, C.L. Bennett (New York)
Coles P., Jones B., 1991, MNRAS, 248, 1
Crampton D., Le F`erve O., Lilly S.J., Hammer F., 1995,
ApJ, 455, 96
Daines S., Jones C., Forman W., Tyson J.A., 1997, ApJ,
submitted
den Hartog R., Katgert P., 1996, MNRAS, 279, 349
Evrard et al., 1996, ApJ, 469, 494
Fort B., Mellier Y., Dantel-Fort M., 1997, AA, 321, 353
Kaiser N., 1996, in “Gravitational Dynamics”, Proc. of the
36thh Herstmonceux Conf., eds O. Lahav, E. Terlevich,
R.J. Terlevich (Cambridge University Press)
Kaiser N., Squires G., 1993, ApJ, 404, 441
Lilly S., Tresse L., Hammer F., Crampton D., Le F´evre O.,
1995, ApJ, 455, 108
Le F´evre O., Hudon D., Lilly S.J., Crampton D., Hammer
F., Tresse L., 1996, ApJ, 461, 534
Miralda-Escud´e J., Babul A., 1995, ApJ, 449, 18
Mushotzky R.F., Scharf C.A., 1997, ApJ 482, L13
Navarro, J.F., Frenk, C.S., White, S.D.M., 1996, ApJ, 462,
563
Schneider, P. & Seitz, C., 1995, A&A, 294, 411
Smail I., Hogg D.W., Yan L., Cohen J.G., 1995, ApJLett,
449,L105
Taylor A.N., Dye S., 1998, submitted MNRAS
Teague P.F., Carter D., Gray P.M., 1990, ApJSupp, 72, 715
Tyson J.A., Valdes F., Wenk R.A., 1990, ApJ, 349, L1
Tyson J.A., Fischer P., 1995, ApJLett, 446, L55
van Kampen E., 1998, submitted MNRAS
van Kampen E., Katgert P., 1997, MNRAS, 209, 327
Yamashita, 1994, in Recontr. de Moriond on “Clusters of
Galaxies”, eds F. Durret
|
0810.2967 | 1 | 0810 | 2008-10-16T17:23:01 | The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER | [
"astro-ph"
] | TRACER ('Transition Radiation Array for Cosmic Energetic Radiation') is a balloon borne instrument that has been developed to directly measure the composition and energy spectra of individual heavy elements up to 10^15 eV per particle. TRACER achieves a large geometric factor (5 m^2 sr) through the use of a Transition Radiation Detector utilizing arrays of single wire proportional tubes. TRACER has measured the energy spectra of the elements O, Ne, Mg, Si, S, Ar, Ca, and Fe. The energy spectra reach energies in excess of 10^14 eV per particle and exhibit nearly the same spectral index (2.65 +/- 0.05) for all elements. | astro-ph | astro-ph | May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
8
0
0
2
t
c
O
6
1
]
h
p
-
o
r
t
s
a
[
1
v
7
6
9
2
.
0
1
8
0
:
v
i
X
r
a
Modern Physics Letters A
c(cid:13) World Scientific Publishing Company
The Elemental Composition of High-Energy Cosmic Rays:
Measurements with TRACER
Enrico Fermi Institute, The University of Chicago, 933 E. 56th Street
P.J. Boyle
Chicago, Illinois 60637, USA.
[email protected]
Received (Day Month Year)
Revised (Day Month Year)
TRACER ("Transition Radiation Array for Cosmic Energetic Radiation") is a bal-
loon borne instrument that has been developed to directly measure the composition and
energy spectra of individual heavy elements up to 1015 eV particle−1. TRACER achieves
a large geometric factor (5 m2 sr) through the use of a Transition Radiation Detector
utilizing arrays of single wire proportional tubes. TRACER has measured the energy
spectra of the elements O, Ne, Mg, Si, S, Ar, Ca, and Fe. The energy spectra reach
energies in excess of 1014 eV particle−1 and exhibit nearly the same spectral index (2.65
± 0.05) for all elements.
Keywords: Cosmic rays; diffusive shock acceleration; energy spectrum; composition; tran-
sition radiation detector
PACS Nos.: 95.55.Ym; 95.85.Ry; 98.70.Sa
1. Introduction
Cosmic rays arriving at Earth span an energy range from 108 eV to 1020 eV. Mea-
surements of the cosmic-ray flux from experiments, both on the ground and above
the atmosphere, are presented in Figure 1 and show the cosmic-ray flux falling as a
near featureless power law over 30 decades. Accurate measurements of the elemen-
tal composition and individual energy spectra has, however, been an experimental
challenge for many years. To date, few measurements of the energy spectra above
1012 eV particle−1 have been made for individual elements. In this paper we review
an effort by the TRACER group to directly measure the elemental composition of
cosmic rays up to 1015 eV particle−1.
The TRACER ("Transition Radiation Array for Cosmic Energetic Radiation")
instrument is a balloon borne detector and probably the largest cosmic-ray detector
ever flown above the atmosphere. TRACER has had three successful balloon flights
since 1999, yielding an exposure of ∼70 m2 sr days. The observational goal of the
first two flights of TRACER was to measure the individual energy spectra of the
heavy elements O, Ne, Mg, Si, S, Ar, Ca, and Fe. For the third flight the instrument
1
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
2 P.J. Boyle
was upgraded to be sensitive to the light-medium elements B, C, and, N. Future
flights of TRACER are proposed and would include measurements of the the sub-Fe
elements Sc, Ti, V, Cr, and Mn.
This review begins with an overview of the TRACER concept and a description
of the instrument in Section 2. The analysis of the TRACER data is detailed in
Section 3 and the resulting energy spectra are presented and compared with mea-
surements from other experiments in Section 4. The measurement of the elements
B,C, and sub-Fe is discussed in Section 5.
1
-
)
V
e
G
s
r
s
2
m
(
x
u
F
l
10 2
-1
-4
-7
10
10
10
-10
-13
-16
-19
-22
-25
-28
10
10
10
10
10
10
10
Satellites
Fluxes of Cosmic Rays
(1 particle per m-2 second-1)
Knee
(1 particle m-2 year-1)
Balloons
Ankle
(1 particle km-2 year-1)
Air-showers
10
10
12
10
14
10
16
10
18
20
10
10
Energy (eV)
Fig. 1. All particle cosmic-ray spectrum. The hashed regions indicates the energy regions sensitive
to satellites, balloon based detectors, and ground based air-shower detectors. [adapted from Cronin
et al.]
2. The TRACER Concept
A successful cosmic-ray experiment must determine for each nucleus both the nu-
clear charge Z and the energy E or Lorentz factor γ ≈ E/mc2. TRACER realizes
this goal by combining scintillation and Cherenkov counters to measure the charge,
together with a dE/dx counter and a Transition Radiation Detector (TRD) to mea-
sure the energy of individual particles up to 1015 eV per particle (see Figure 2).
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER 3
The key feature on TRACER is the employment of a TRD to measure the energy
or Lorentz factor of cosmic-ray nuclei up to 1015 eV per particle. This application
of TRD's is more challenging than their use as threshold devices in accelerator
or cosmic-ray measurements [2,3,4], where one just wishes to discriminate between
particles of the same energy but different mass (e.g. electrons and pions or protons).
Key features which distinguish TRD's from conventional energy measuring de-
vices such as magnet spectrometers or hadronic calorimeters include: (a) a favorable
area-to-weight ratio, which is important for weight-restricted-instruments on bal-
loons or spacecraft (b) the possibility of accelerator calibrations with beams of
electrons or pions at large Lorentz factors, for which beams of nuclei are not avail-
able (c) the fact that the TR-signal scales strictly with Z 2. Therefore, relative signal
fluctuations decrease as 1/Z, making possible increasingly precise measurements for
particles with higher charge. On the other hand, large signal fluctuations at low Z
restrict this use of TRD's to cosmic-ray nuclei heavier than helium (Z ≥ 3) (d) a
good match between the Lorentz-factor range of a TRD (400 - 105) and the range
of energies that should be covered in direct measurements (e) the possibility of re-
dundant measurements in a layered radiator/detector configuration (f) good energy
resolution, typically of the order of 10% or better for the heavier nuclei at γ ≈ 1000.
Traditionally, a TRD uses multiwire proportional chambers (MWPC's) as x-ray
detectors. MWPC's require the use of a pressurized container, as employed in the
Cosmic Ray Nuclei (CRN) experiment which was the first TRD flown in space [5].
The weight of such a container is large, and the need for it can be avoided by
the approach taken by TRACER. For the TRACER instrument the MWPC's are
replaced with layers of thin-walled single wire proportional tubes which can easily
withstand internal overpressure.
2.1. The TRACER Instrument
Balloon borne experiments are limited in size and weight. Payloads are presently
limited to ∼2700 Kg, of which ∼700 Kg is reserved for satellite telemetry interfaces,
ballast and landing gear (parachute and crush pad). TRACER is at the upper limit
in terms of size and weight for a balloon payload. The employment of a TRD,
without the need of a pressurized container allows for the construction of a large
area detector (2.06 x 2.06 m2), resulting in an overall geometric factor of 5 m2sr.
The individual components of TRACER are shown in Figure 2 and are summarized
here. From top to bottom the instrument consists of:
1 Scintillator Counter : A plastic counter (BICRON-408) that has an active
area of 2 m x 2 m and is 0.5 cm thick. The counter is read out via wavelength
shifter bars (BICRON 482) by 24 photomultiplier tubes (Photonis XP1910).
2 dE/dx Counter : An array of 800 single wire proportional tubes. Each tube
is 2 m long and 2cm in diameter and the walls, which are made of mylar,
are 127 µm thick. The tubes are arranged in a total of four double layers,
with each layer consisting of 100 tubes. Each double layer is orientated in
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
4 P.J. Boyle
Fig. 2. Schematic diagram of the TRACER detector as flown during the Antarctic balloon cam-
paign in 2003. [From Ave et al.]
an orthogonal direction to the one above.
3 Transition Radiation Detector : A second array of 800 single wire propor-
tional tubes identical to the dE/dx counter except that each double layer
is preceded by blankets of plastic fiber material which act as a radiator to
generate transition radiation.
4 Scintillator Counter : A second counter of identical design as the top scin-
tillator counter is placed below the TRD.
5 Cherenkov Counter : An acrylic counter with refractive index = 1.49 is
placed at the bottom of the stack. The active area of the counter is 2 m x
2 m and the thickness is 1.27 cm. The counter is also read out, via wave-
length shifting bars, by 24 PMTs.
The scintillator counters act as a trigger for the instrument and together with
the Cherenkov detector measure the nuclear charge Z of each individual cosmic-
ray particle traversing the instrument. The array of all 1600 proportional tubes
determines the trajectory of each cosmic-ray particle through the instrument. The
energy response of the TRACER detector spans 4 decades in energy and is shown in
Figure 3. This extensive energy range is accomplished by combining three comple-
mentary measurements: Cherenkov light measurement (∼ 1011 eV), the relativistic
rise of the ionization signal in gas (∼ 1011 − 1013 eV) and the measurement of
transition radiation (> 1013 eV).
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER 5
)
s
t
i
n
u
b
r
a
(
l
a
n
g
S
i
0.3
0.2
0.1
0
1
CER
TRD
DEDX
10
2
10
3
10
Lorentz Factor (g )
10
4
Fig. 3. Energy scale of TRACER. TRACER combines three energy measurements to cover more
than 4 decades in energy. The relative amplitude of the response curves is representative of the
signal/Z 2 in each detector [From Ave et al.]
2.2. Balloon Flights
TRACER has had three successful flights on high altitude balloon thus far. The
instrument and data have been recovered intact after each flight. A summary of the
flights is as follow:
Flight I: Launch from Ft. Sumner, USA on September 20, 1999 on a 39 million
cubic-foot balloon. The flight was 28 hours in duration at a float altitude between
34 and 38 km, corresponding to a residual atmospheric depth of 4-6.5 g cm−2. This
flight served as a test flight for the subsequent longer duration flights.
Flight II: Launch from Antarctica on December 12, 2003 on a 39 million cubic-
foot balloon. The flight lasted 14 days at a float altitude between 36 and 39 km,
corresponding to an average residual atmospheric depth of 3.9 g cm−2. A total of
5 × 107 cosmic-ray particles were collected.
For the first two flights, the trigger threshold was set such that the instrument
had full efficiency for the elements oxygen to iron.
Flight III: Launch from Kiruna, Sweden on July 8, 2006 on a 39 million cubic-
foot balloon. The flight was limited to 4.5 days due to the lack of permission to fly
over Russian territory. The float altitude was between 36 and 40 km, corresponding
to an average residual atmospheric depth of 3.5 g cm−2. For this flight, the trigger
threshold was set at a signal level corresponding to a nuclear charge between lithium
and beryllium. This allowed for full efficiency for the lighter elements boron, carbon,
and nitrogen in addition to the heavier elements. A total of 3 × 107 cosmic-ray
particles were collected.
The results of Flight I have been published by Gahbauer et al. (2004) [6]. The
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
6 P.J. Boyle
subsequent discussion of the analysis and results from TRACER presented in this
paper pertain to Flight II. These results have been published by Ave et al. [7], where
the analysis is discussed in greater detail. Analysis of data collected in Flight III is
currently ongoing.
3. Constructing an Energy Spectrum with TRACER
The analysis of the TRACER data begins with the reconstruction of the trajectory
of each cosmic-ray particle through the instrument. The accurate knowledge of
the particle trajectory is essential to the analysis for two reasons; first, it permits
corrections of the scintillator and Cherenkov signals due to spatial non-uniformities
and zenith-angle variations, second, it makes it possible to determine accurately the
energy per unit pathlength deposited by each cosmic ray traversing the proportional
tubes.
3.1. Trajectory Reconstruction
The trajectory of cosmic rays through the instrument is reconstructed for signals
recorded in the entire proportional tube array. The reconstruction follows a two step
procedure. As a first estimate, the trajectory is obtained by constructing a straight
line fit to the center of each of the tubes hit in an event. All possible combinations
of these tubes are fit in the x- and y-projections, and the combination with the
minimum χ2 per degree of freedom and maximum number of tubes hit is kept. This
procedure reconstructs the trajectories of over 95% of all cosmic-ray particles, with
a lateral accuracy in the track position of 5 mm. The trajectory is then refined using
the fact that the energy deposited in each tube is proportional to the track length
in that tube. Taking this fact into account, an accuracy of 2 mm in lateral track
position is achieved, which corresponds to a 3% uncertainty in the total path length
through all tubes.
3.2. Charge Reconstruction
The charge is determined for each cosmic-ray particle using the signals from the
scintillation and Cherenkov counters. The signals are corrected for spatial non-
uniformities in the counter responses using the trajectory information together with
response maps recorded with muons on the ground. Figure 4 shows a correlation of
the signals from the top scintillator and Cherenkov counters. Particles of constant
charge are clustered along the lines, with the position on the line dependent on the
primary energy. A charge histogram is then constructed through the summation
along these lines of constant charge.
The resulting charge resolution is both charge and energy dependent. For oxygen
nuclei (Z=8) the charge evolves from 0.25 at energies below 3 GeV amu−1 to 0.30
charge units at higher energies, while for iron nuclei (Z=26) the resolution is 0.5
and 0.6 charge units respectively.
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER 7
·
90
80
70
60
50
40
30
20
10
0
3
10
1000
800
600
400
200
0
16
18
20
22
24
26
28
30
4
6
8
10
Charge Z
12
14
16
Fig. 4. Scatter plot of the signal from the scintillator and Cherenkov counters (left). A charge
histogram for all events is obtained by summing along lines of constant charge (right). [From Ave
et al.]
3.3. Energy Reconstruction
The energy of each cosmic-ray particle is obtained from the combined signals of the
Cherenkov counter and of the proportional tubes. The Cherenkov counter identifies
particles below minimum ionization energy (3 GeV amu−1), and also provides a good
energy measurement for these low-energy particles (see Figure 3). Between minimum
ionization energy and the onset of TR (400 GeV amu−1), the signals in the dE/dx
counter and the TRD are the same and increase logarithmically with energy (see
Figure 3). Above 400 GeV amu−1, the signals from the dE/dx counter and TRD
diverge, and the rapid increase in the TR signals with particle energy provides an
excellent energy measurement. However, cosmic-ray particles in this energy region
are extremely rare, being less abundant than particles in the minimum ionization
energy region by more than 4 orders of magnitude. To uniquely identify these rare
cosmic-ray particles, it is required that particles with low-energy are identified and
removed on the basis of their Cherenkov signals. To select the highest energy events,
it is required that the measurement in the dE/dx counter places these particles at
an energy level well above minimum ionization energy and the presence of TR is
detected. Thus the combination of the Cherenkov counter, the dE/dx counter and
the TRD is crucial for the success of the TRACER measurement at the highest
energies.
Figure 5 shows a cross-correlation plot between the TRD and dE/dx counter
signals for neon nuclei (Z=10). Cosmic-ray particles below minimum ionization en-
ergy have been removed with the aid of the Cherenkov counter. The small black
points represent particles with energies below the onset of TR. As the signals from
the TRD and dE/dx counter are similar in this energy regime, the correlation of the
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
8 P.J. Boyle
)
s
t
i
n
u
.
b
r
a
(
R
T
+
n
o
i
t
a
z
i
n
o
I
c
i
f
i
c
e
p
S
0.4
0.3
0.2
0.1
0
0
6 x 1014 eV
NEON
0.1
0.2
0.3
0.4
Specific Ionization (arb. units)
Fig. 5. Scatter plot of the signal from the TRD and the DEDX counters for neon nuclei. The
highlighted points represent the highest energy events as measured with the TRD. As expected
the TR events have signals in the DEDX counter which are well above the minimum ionization
energy (white circle). [From Ave et al.]
signals lies along the diagonal. At the onset of TR, where the TRD and the dE/dx
signals diverge, the highest energy events show up away from the diagonal line. The
off-diagonal position of these highest energy particles defines them uniquely as TR
events for this selected charge. The highest energy neon nucleus in this data sample
has an energy of 6 × 1014 eV particle−1.
3.4. Absolute Intensities
Once each cosmic-ray particle has been assigned an energy it is sorted into energy
bins of width ∆Ei and a differential energy spectrum is constructed for each elemen-
tal species. To convert from the number of events ∆Ni in a particular energy bin
∆Ei to an absolute flux dN/dE(i) one must compute the exposure factor, effective
aperture of the instrument, and the efficiency of the analysis selection and instru-
ment response. The overall efficiency is high, with a tracking efficiency of ∼95%,
livetime ∼94%, and charge selection efficiency 70-80% (for the more abundant ele-
ments). Further details of the reconstruction efficiencies can be found in Ave et al.
[7].
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER 9
1
-
)
V
e
G
s
r
s
2
m
(
x
u
F
l
10 4
1
-4
-8
10
10
-12
10
-16
10
-20
10
-24
10
-28
10
-32
10
H
He x 10-2
C x 10-4
O x 10-6
Ne x 10-8
Mg x 10-10
Si x 10-12
S x 10-14
Ar x 10-16
Ca x 10-18
Fe x 10-21
Heavy Nuclei
TRACER
HEAO-3
CRN
-1
10
1
10
p + He
AMS
BESS
ATIC-2
JACEE
CAPRICE
RUNJOB
2
3
10
10
Kinetic Energy per Particle (GeV)
10
4
10
5
10
6
Fig. 6. Flux as a function of energy for the major components of the primary cosmic radiation.
The measurements by the TRACER experiment are represented by the solid squares for the
elements O, Ne, Mg, Si, S, Ar, Ca, and, Fe. The dashed line represents a power-law fit to the
TRACER data above 20 GeV amu−1. For references to the data presented in this plot see [7]-[16]
and references therein.
4. Resulting Energy Spectra
The energy spectra, in terms of absolute intensities at the top of the atmosphere,
for the elements O, Ne, Mg, Si, S, Ar, Ca, and Fe are plotted as solid squares in
Figure 6. For clarity, the intensity of each element is scaled by a factor shown on the
left. Existing data from measurements in space with HEAO-3 (open diamonds [8])
and CRN (open crosses [9]) are shown for comparison. For completeness, data on the
light primary cosmic ray components (protons, helium, and carbon) which are not
measured with TRACER are also included. These data come from measurements
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
10 P.J. Boyle
in space (AMS [10]) and on balloons (ATIC [11], BESS [12, 13], CAPRICE [14],
JACEE [15], RUNJOB [16]).
Note the large range in intensity (10 decades) and particle energy (4 decades)
covered by TRACER. As can be seen, the energy spectra for O, Ne, Mg, and Fe
extend up to and beyond 1014 eV particle−1. The energy range is limited by the
current exposure and not by the saturation of the TRD. No significant change
in spectral slope is evident at the highest energies. The energy spectrum of each
element, as measured by TRACER, can be fit to a power law above 20 GeV amu−1
(dashed line in Figure 6). The resulting spectral indices are remarkably similar with
an average exponent of -2.65 ± 0.05.
1
-
)
V
e
G
s
r
s
2
m
(
x
u
F
l
-3
-4
-5
-6
-7
-8
-9
10
10
10
10
10
10
10
-10
-11
-12
10
10
10
Acrylic Cerenkov
Iron
Aerogel Cerenkov
Gas Cerenkov
Specific Ionization
Thin Calorimeter
TRACER
HEAO-3
CRN
ATIC-2
RUNJOB
HESS-QGSJET
HESS-SYBILL
Transition Radiation
Direct Cerenkov
2
10
3
10
4
10
5
10
Kinetic Energy per Particle (GeV)
Fig. 7. Energy spectrum for iron nuclei highlighting the complementarity between detection tech-
niques employed in cosmic-ray composition measurements.
Figure 7 compares the TRACER results for iron nuclei with results from previous
experiments in space (HEAO-3 [8] and CRN [9]), and on balloons (ATIC-2 [11] and
RUNJOB [16]). Figure 7 also illustrates the variety of detection techniques used in
measuring the energy of heavy nuclei. Within the statistical uncertainties (which
in some measurements are quite large), the data indicate fairly consistent results.
Also presented in Figure 7 are recent results from the ground based HESS Imaging
Air Cherenkov Telescope using the Direct Cherenkov Technique [17]. Here, two
flux values are presented for each energy arising from ambiguities from different
nuclear interaction models used in the data analysis [18]. Again, the HESS results
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER 11
are consistent with TRACER.
)
5
6
.
1
V
e
G
1
-
s
1
-
r
s
2
-
m
(
y
t
i
s
n
e
t
n
I
x
5
6
2
E
.
10 5
10 4
10 3
10 2
10 5
10 4
10 3
10 2
oxygen
iron
2
10
3
10
4
10
5
10
6
10
7
10
8
10
Kinetic Energy per Particle (GeV)
Fig. 8. Energy spectra multiplied by E2.65 from TRACER (solid squares) compared with the in-
terpretation of air-shower data of KASCADE (filled triangles, for two different interaction models:
Antoni et al. (2005)) and of EAS-TOP (open crosses, two data points for each energy represent
upper and lower limits: Navarra et al. (2003). [From Ave et al.]
4.1. Comparison with Air-shower Data
Data for oxygen and iron nuclei from TRACER are compared in Figure 8 with
energy spectra derived from indirect air-shower observations of the EAS-TOP col-
laboration [19] and of the KASCADE group [20]. These experiments report results
for groups of elements and not individual elements; therefore the fluxes for the
"CNO group" probably have about twice the intensity than oxygen alone while the
"iron group" is probably dominated by iron. Furthermore, these results remain am-
biguous as they depend strongly on the choice of the nuclear interaction model used
in the analysis. The TRACER results do not yet overlap with the energy region of
the air-shower data. Additional measurements with TRACER should help to close
the gap and to provide significant constraints on the interpretation of air-shower
results.
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
12 P.J. Boyle
5. Future Measurements with TRACER
Extending the energy spectra of the heavy elements towards 1015 eV particle−1 is
an ongoing goal of TRACER. However, to determine the energy spectra and relative
abundances for each nuclear species at the cosmic-ray source, the mode of galactic
propagation must be understood. The simplest model of cosmic-ray propagation
assumes an equilibrium between the production of cosmic rays, from acceleration
at the source or from spallation of heavier nuclei, and their loss from competing
actions of escape from the Galaxy and spallation on the interstellar medium [21].
This model can be described as:
Ni(E) =
1
Λ(E)−1 + Λ−1
s
(
Qi(E)
βcρ
+ X
k>i
Nk
Λk→i
)
(1)
where Ni(E) is the observed intensity of element i, Λ(E) the propagation path-
length, Λs the spallation pathlength, Qi(E) the rate of production in the cosmic-ray
source, β = v/c, ρ the average density of the interstellar medium, Nk the intensity
of element k, and Λk→i the spallation mean free path for an element k to spallate
to element i.
o
i
t
a
R
x
u
F
n
o
b
r
a
C
l
/
n
o
r
o
B
HEAO-3
CRN
-1
10
-2
10
(E)= C E-0.3
(E)= A E-0.6
(E)= A E-0.6 + 0.3 g/cm2
1
10
2
10
3
10
4
10
Kinetic Energy (GeV/AMU)
Fig. 9. Boron to carbon flux ratio with data from HEAO-3 and CRN. The solid line represent a
parameterization given by Yanasak. The addition of a residual pathlength of Λ0 = 0.3 g/cm2 to
the Yanasak parameterization is presented by the dash-dot line. The dashed line represents a fit
to Λ(E) ∝ E−0.3. The hashed area represents the energy region attainable from the 5-day flight
of TRACER in 2006.
L
L
L
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER 13
The energy dependent propagation pathlength Λ(E) can be derived from the
flux ratio of secondary spallation produced particles to primary accelerated par-
ticles. This flux ratio has been observed to decrease with increasing energy [22],
indicating that the mean lifetime or propagation pathlength of cosmic rays in the
Galaxy decreases with energy. As an example, the boron to carbon flux ratio as
measured by HEAO-3 [8] and CRN [9] is shown in Figure 9. The flux ratio has been
parameterized with Λ(E) ∝ E−0.6 [8]. However, this flux ratio has been measured
accurately only to about 1012 eV particle−1. It may well be that the flux ratio will
eventually reach a finite asymptotic value Λ0 (i.e. Λ(E) = AE−0.6 + Λ0). The
residual pathlength Λ0 would represent the minimum amount of matter high-energy
particles must encounter during propagation from the cosmic-ray source to Earth.
Alternatively, some diffusion propagation models expect a Kolmogorov spectrum
with Λ(E) = CE−0.3 (see [24] for a comprehensive review, and references therein).
Figure 9 illustrates that none of these scenarios cannot be fully ruled out with the
present data. An extension of the boron-carbon flux ratio was a major science goal
of Flight III of TRACER in 2006. It is expected that even with the limited exposure
obtained in Flight III, the measurement of the boron-carbon flux ratio will reach
into the 1013-1014 eV particle−1 range. This energy region is represented as the
shaded area in Figure 9.
The flux ratio of the secondary sub-Fe elements (Z=21-25) to iron (Z=26) can
also be used to determine the propagation parameters Λ(E) and Λ0. The sub-Fe
elements are produced by the spallation of iron nuclei during propagation. Figure 10
shows the currently available energy spectra for the sub-Fe elements along with B,
C, Ar, and Ca spectra. The energy spectra for the sub-Fe elements do not yet reach
1012 eV particle−1.
TRACER is unable to measure the energy spectra for the sub-Fe elements with
the present detector configuration. The main reasons for this are the limited charge
resolution, due to nonlinearity of the scintillator response at high charges, and the
limited photo-electron statistics of the Cherenkov counter. It is therefore proposed
to replace one of the acrylic Cherenkov counters with a counter containing an aerogel
radiator (refractive index 1.04), and to use a more efficient readout system for the
second acrylic Cherenkov counter. For relativistic particles the charge would then
be reconstructed using the acrylic and the aerogel Cherenkov counters. Contrary
to the behavior of the scintillators, both signals are strictly proportional to Z 2 and
the charge resolution would be independent of charge. This technique has been
successfully utilized in the balloon borne Trans Iron Galactic Recorder (TIGER)
instrument [25], which achieved a charge independent resolution of 0.22 charge units.
Monte-Carlo simulations have shown that a charge resolution of 0.2 charge units
could be achieved with the proposed Cherenkov counter system [26,27]. Figure 11
shows the results from a simulated charge reconstruction for such a setup. The
left-hand panel shows a cross correlation plot of the simulated signals from the
aerogel and acrylic counters. The vertical lines represent cosmic rays of constant
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
14 P.J. Boyle
10 9
10 8
10 6
10 4
10 2
1
-2
-4
-6
-8
10
10
10
10
)
5
6
1
.
)
u
m
a
V
e
G
/
(
1
-
s
1
-
r
s
2
-
m
l
(
x
u
F
x
5
6
.
2
E
-10
-12
-14
-16
10
10
10
10
B x 107
N x 105
Ar x 103
Ca
Sc x 10-3
Ti x 10-6
V x 10-8
Cr x 10-10
Mn x 10-12
-1
10
1
10
3
4
2
10
10
10
Kinetic Energy (GeV/AMU)
10
5
Fig. 10. Energy spectra multiplied by E2.65 of purely secondary elements (B and sub-Fe) and
mixed elements (primary and secondary: N, Ar, and Ca) from the HEAO-3 (open diamonds),
CRN (open crosses) and TRACER (solid squares). The hashed region represents the energy region
attainable with a 30 day flight of the upgraded TRACER.
charge. The resulting charge histogram, shown in the right-hand panel of Figure 11,
indicates that the sub-Fe elements are clearly resolved. As Antarctic flights of 30
days duration are now becoming a reality [28], such an exposure of the upgraded
TRACER instrument would extend the energy spectra for all secondary elements to
1014 eV particle−1. This energy region is indicated by the shaded region in Figure 10.
6. Conclusions
The TRACER program has pioneered the development of a TRD using proportional
tube arrays to directly measure the energy spectra of cosmic-ray nuclei. The coupling
of the TRD with the Cherenkov and DEDX counters enables an energy measurement
of individual nuclei across 5 decades in energy and provides a technique to uniquely
identify the rare high-energy particles with a discrimination power of > 104.
TRACER has measured the energy spectra of the elements O, Ne, Mg, Si, S,
Ar, Ca, and Fe up to 1014 eV particle−1. This data set currently represents the
most comprehensive measurements to date of heavy nuclei with individual charge
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
The Elemental Composition of High-Energy Cosmic Rays: Measurements with TRACER 15
)
4
0
.
1
=
n
(
l
e
g
o
r
e
A
9
8
7
6
5
4
3
2
1
0
S
Si
Mg
Ne
O
N
C
B
Be
Fe
Cr
Ti
Ca
Ar
12000
10000
8000
6000
C
O
1800
1600
1400
1200
1000
800
600
400
200
0
Fe
S
Ca
Ar
Ti
Cr
16
18
20
22
24
26
28
Charge Z
B
4000
N
2000
Mg
Ne
Si
0
4
6
8
10
Charge Z
12
14
16
2
4
6
8
10
12
14
18
16
22
Acrylic (n=1.49)
20
Fig. 11.
(Left-hand panel) Cross correlation plot of simulated signals from two Cherenkov coun-
ters; one containing an aerogel radiator (refractive index = 1.04) and the second an acrylic plastic
radiator (refractive index = 1.49). (Right-hand Panel) Resulting charge histogram with charge
resolution of 0.2 charge units.
resolution. The measured energy spectra reach similar power laws at high energies
and no change in slope is evident. The data indicate a common origin and mode of
propagation for all elements and support the SNR theory of the origin of galactic
cosmic rays.
TRACER has been upgraded to include a measurement of the secondary to
primary flux ratio. Data analysis of the boron to carbon flux ratio is currently
underway. Future upgrades of TRACER will include measurements of the sub-Fe
elements.
Acknowledgments
This work has been made possible by the enormous contributions of the TRACER
team: Maximo Ave, Florian Gahbauer, Christian Hoppner, Jorg Horandel,
Masakatsu Ichimura, Jesse Marshall, Dietrich Muller, Andreas Obermeier, and An-
drew Romero-Wolf. We gratefully acknowledge the services of the University of
Chicago Engineering Center in particular Gary Kelderhouse, Gene Drag, Casey
Smith, Richard Northrop and Paul Waltz. We thank the staff of the Columbia
Scientific Balloon Facility, the NSF Antarctic Program and the Swedish Space Cor-
poration for support during the balloon campaigns. This work was supported by
NASA grants NAG55305, NN04WC08G and NNG06WC05G. Numerous students
have participated in the construction and refurbishment of the instrument under
support from the Illinois Space Grant Consortium.
May 30, 2018 5:4 WSPC/INSTRUCTION FILE
MPLA-BOYLE2008
16 P.J. Boyle
References
1. J. Cronin, T.K. Gaisser & S.P. Swordy, Scientific American 276, 44 (1997)
2. G. Hartmann, D. Muller, T. Prince, Physical Review Letters 38, 1368 (1977)
3. R. Bellotti et al., Nucl. Phys. B, Proc. Suppl. 54B, 375 (1997)
4. S.W. Barwick et al., AP J 482L, 191B (1997)
5. J. L'Heureux, J. Grunsfeld, P. Meyer, D. Muller, & S.P. Swordy, NIM A 295, 246
(1990)
6. F. Gahbauer et al., AP J 607, 333 (2003)
7. M. Ave et al., AP J 678, 262 (2008)
8. J.J. Engelmann et al., Astronomy & Astrophysics 233, 96 (1990)
9. D. Muller et al., AP J 374, 356 (1991)
10. AMS Collaboration, Physics Letters B490, 27 (2000)
11. H.S. Ahn et al., Advances in Space Science Reviews 37, 1950A (2006)
12. T. Sanuki et al., AP J 545, 1135 (2000)
13. S. Haino et al., Physics Letters B594, 35H (2004)
14. M. Boezio et al., Astroparticle Physics 19, 583 (2003)
15. K. Asakimori et al., AP J 502, 278 (1998)
16. V. Derbina et al., AP J 628L, 41D (2005)
17. D. Kieda et al., Proc. 27th ICRC 1, 147 (2003)
18. F. Aharonian et al., Physical Review D75, 042004 (2007)
19. G. Navarra et al., Proc. 28th ICRC 1, 147 (2003)
20. T. Antoni et al., Astroparticle Physics 24, 1 (2005)
21. S. Swordy et al., AP J 403, 658S (1993)
22. E. Juliusson et al., Phys. Rev. Lett. 29, 445 (1972)
23. N. E. Yanasak et al., AP J 563, 768
24. A. Strong, I. Moskalenko, V. Ptuskin, Annu. Rev. Nucl. Part. Sci. 57, 285-327 (2007)
25. J. T. Link et al., Proc. 28th ICRC 4, 1781L (2003)
26. J. Marshall & A. Obermier, Private communication (2008)
27. T. Hams, Private communication (2008)
28. W.V. Jones, Inter. Cosmic-ray Conf 10, 173J (2005)
|
astro-ph/0503567 | 1 | 0503 | 2005-03-25T21:08:51 | On the Absorption and Emission Properties of Interstellar Grains | [
"astro-ph"
] | Our current understanding of the absorption and emission properties of interstellar grains are reviewed. The constraints placed by the Kramers-Kronig relation on the wavelength-dependence and the maximum allowable quantity of the dust absorption are discussed. Comparisons of the opacities (mass absorption coefficients) derived from interstellar dust models with those directly estimated from observations are presented. | astro-ph | astro-ph |
invited talk for "The Spectral Energy Distribution of Gas-Rich Galaxies:
Confronting Models with Data" (Heidelberg, Germany/4-8 October
2004), edited by C.C. Popescu & R.J. Tuffs, AIP Conf.
Ser., in press
On the Absorption and Emission Properties of
Interstellar Grains
Aigen Li
Department of Physics and Astronomy, University of Missouri, Columbia, MO 65211, USA
email: [email protected]
Abstract. Our current understanding of the absorption and emission properties of interstellar grains
are reviewed. The constraints placed by the Kramers-Kronig relation on the wavelength-dependence
and the maximum allowable quantity of the dust absorption are discussed. Comparisons of the
opacities (mass absorption coefficients) derived from interstellar dust models with those directly
estimated from observations are presented.
1. INTRODUCTION
Interstellar dust reveals its presence in astrophysical environments and its (both positive
and negative) role in astrophysics mainly through its interaction with electromagnetic
radiation (see Li [38] for a recent review):
• obscuring distant stars by the absorption and scattering of starlight by dust (the
combined effects of absorption and scattering are called extinction);
• reddening starlight because the extinction is stronger for blue light than for red;
• generating "reflection nebulae" by the scattering of starlight by dust in interstellar
clouds near one or more bright stars;
• generating the "diffuse Galactic light" seen in all directions in the sky by the diffuse
scattering of starlight of stars located near the Galactic plane;
• generating X-ray halos by the small-angle dust scattering of X-ray sources;
• polarizing starlight as a result of preferential extinction of one linear polarization
over another by aligned nonspherical dust;
• heating the interstellar gas by ejecting photoelectrons created by the absorption of
energetic photons;
• and radiating away the absorbed short-wavelength radiation at longer wavelengths
from near infrared (IR) to millimeter (mm) in the form of thermal emission, with a
small fraction at far-red wavelengths as luminescence.
In order to correct for the effects of interstellar extinction and deredden the reddened
starlight, it is essential to understand the absorption and scattering properties of inter-
stellar grains at short wavelengths (particularly in the optical and ultraviolet [UV]). The
knowledge of the optical and UV properties of interstellar dust is also essential for in-
terstellar chemistry modeling since the attenuation of UV photons by dust in molecu-
lar clouds protects molecules from being photodissociated. The knowledge of the dust
emission properties at longer wavelengths are important (i) for interpreting the IR and
submillimeter (submm) observations of emission from dust and tracing the physical con-
ditions of the emitting regions, (ii) for understanding the process of star formation for
which the dust is not only a building block but also radiates away the gravitational energy
of collapsing clouds (in the form of IR emission) and therefore making star formation
possible, and (iii) for understanding the heating and cooling of the interstellar medium
(ISM) for which interstellar dust is a dominant heating source by providing photoelec-
trons (in the diffuse ISM) and an important cooling agent in dense regions by radiating
in the IR (see Li & Greenberg [45] for a review).
Ideally, if we know the size, shape, geometry and chemical composition (and therefore
the dielectric function) of an interstellar grain, we can calculate its absorption and
scattering cross sections as a function of wavelength. If we also know the intensity of the
illuminating radiation field, we should be able to calculate the equilibrium temperature
or temperature distribution of the grain from its absorption cross section and therefore
predict its IR emission spectrum.
However, our current knowledge of the grain size, shape, geometry and chemical
composition is very limited; the nature of interstellar dust itself is actually mainly
derived from its interaction with radiation (see Li [38] for a review):
• from the interstellar extinction curve which displays an almost linear rise with
inverse wavelength (l −1) from the near-IR to the near-UV and a steep rise into the
far-UV one can conclude that interstellar grains must span a wide range of sizes,
containing appreciable numbers of submicron-sized grains as well as nanometer-
sized grains;
• from the wavelength dependence of the interstellar polarization which peaks at
l ∼ 0.55 m m, one can conclude that some fraction of the interstellar grains must be
nonspherical and aligned by some process, with a characteristic size of ∼ 0.1 m m;
• from the scattering properties measured for interstellar dust which are characterized
by a quite high albedo (∼0.6) in the near-IR and optical and a quite high asymmetry
factor (typically ∼0.6 -- 0.8 in the optical) one can infer that a considerable fraction
of the dust must be dielectric and the predominantly forward-scattering grains are
in the submicron size range;
• from the IR emission spectrum of the diffuse ISM which is characterized by a
modified black-body of l −1.7 Bl (T=19.5 K) peaking at ∼ 130 m m in the wave-
length range of 80 m m ∼< l
∼< 1000 m m, and a substantial amount of emission at
l
∼< 60 m m which far exceeds what would be expected from dust at T ≈ 20 K, one
can conclude that in the diffuse ISM, (1) the bulk interstellar dust is in the submi-
cron size range and heated to an equilibrium temperature around T ∼ 20 K, respon-
sible for the emission at l
∼> 60 m m; and (2) there also exists an appreciable amount
of ultrasmall grains in the size range of a few angstrom to a few nanometers which
are stochastically heated by single UV photons to high temperatures (T > 50 K),
responsible for the emission at l
• from the spectroscopic absorption features at 9.7, 18 m m and 3.4 m m and emission
features at 3.3, 6.2, 7.7, 8.6 and 11.3 m m which are collectively known as the "UIR"
(unidentified IR) bands, one can conclude that interstellar dust consists of appre-
ciable amounts of amorphous silicates (of which the Si-O stretching mode and the
∼< 60 m m (see Li [39]);
O-Si-O bending mode are respectively responsible for the 9.7 and 18 m m features),
aliphatic hydrocarbon dust (of which the C-H stretching mode is responsible for
the 3.4 m m feature), and aromatic hydrocarbon molecules (of which the C-H and
C-C stretching and bending vibrational modes are responsible for the 3.3, 6.2, 7.7,
8.6 and 11.3 m m "UIR" features), although the exact nature of the carriers of the
3.4 m m feature and the "UIR" features remain unknown.
The inferences from observations for interstellar dust summarized above are quite gen-
eral and model-independent. But these inferences are not sufficient to quantitatively
derive the absorption and emission properties of interstellar grains. For a quantitative
investigation, one needs to make prior specific assumptions concerning the grain size,
shape, geometry and chemical composition which are still not well constrained by the
currently available observational data. To this end, one needs to adopt a specific grain
model in which the physical characteristics of interstellar dust are fully specified. While
a wide variety of grains models have been proposed to explain the interstellar extinction,
scattering, polarization, IR emission and elemental depletion, so far no single model can
satisfy all the observational constraints (see Li [39] and Dwek [22] for recent reviews).
In view of this, in this article I will first try to place constraints on the absorption
and emission properties of interstellar dust based on general physical arguments; these
constraints are essentially model-independent.
In astrophysical literature, the most frequently used quantities describing the dust
absorption and emission properties are the mass absorption coefficient (also known as
"opacity") k abs with a unit of cm2 g−1, and the emissivity e l , defined as the energy
emitted per unit wavelength per unit time per unit solid angle per unit mass, with
a unit of erg s−1 sr−1 cm−1 g−1. The Kirchhoff's law relates e l
to k abs through e l =
k abs(l ) Bl (T ) if the dust is large enough to attain an equilibrium temperature T when
exposed to the radiation field, or e l = k abs(l )R
0 dT Bl (T )dP/dT if the dust is so small
that it is subject to single-photon heating and experiences "temperature spikes", where
is the Planck function, dP is the probability for the dust to have a temperature
Bl
in [T, T + dT ]. Other often used quantities are the absorption cross section Cabs and
absorption efficiency Qabs, with the latter defined as the absorption cross section Cabs
divided by the geometrical cross sectional area Cgeo of the grain projected onto a plane
perpendicular to the incident electromagnetic radiation beam (Bohren & Huffman [8]).
For spherical grains of radii a, Cgeo = p a2 so that Qabs = Cabs/p a2. By definition,
k abs = Cabs/m = Cabs/ (Vr ), where m, V and r are respectively the dust mass, volume,
and mass density; for spherical grains, k abs = 3Qabs/ (4ar ).
(the
wavelength dependence exponent index of k abs) and an upper limit on the absolute value
of k abs. The state of our knowledge of interstellar grain opacity will be presented in §3
(with a focus on b ) and in §4 (with a focus on the absolute value of k abs), followed by a
summary in §5.
In §2 I will apply the Kramers-Kronig relation to place a lower limit on b
2. CONSTRAINTS FROM THE KRAMERS-KRONIG RELATION
As already mentioned in §1, with specific assumptions made concerning the grain size,
shape, geometry and composition, in principle one can calculate the absorption cross
¥
section Cabs and the opacity k abs as a function of wavelength. But this is limited to
spherical grains; even for grains with such simple shapes as spheroids and cylinders, the
calculation is complicated and limited to small size parameters defined as 2p a/l . As a
result, astronomers often adopt a simplified formula
k abs(l ) =(cid:26) k 0 ,
k 0 (l /l 0)−b
,
l < l 0 ,
l ≥ l 0 ,
(1)
where l 0 and the exponent index b are usually treated as free parameters, while k 0 is
often taken from experimental measurements of cosmic dust analogs or values predicted
from interstellar dust models; or
(Qabs/a) (l ) =(cid:26) (Qabs/a)0 ,
(Qabs/a)0 (l /l 0)−b
,
l < l 0 ,
l ≥ l 0 ,
(2)
where again l 0 and b are free parameters and (Qabs/a)0 is usually taken from model
calculations.
The Kramers-Kronig dispersion relation, originally deduced by Kronig [33] and
Kramers [32] from the classical Lorentz theory of dispersion of light, connects the
real (m′; dispersive) and imaginary (m′′; absorptive) parts of the index of refraction
(m[l ] = m′ +i m′′) based on the fundamental requirement of causality. Purcell [55] found
that the Kramers-Kronig relation can be used to relate the extinction cross section inte-
grated over the entire wavelength range to the dust volume V
Cext(l ) dl = 3p 2FV ,
Z
0
(3)
where Cext
is the extinction cross section, and F, a dimensionless factor, is the
orientationally-averaged polarizability relative to the polarizability of an equal-volume
conducting sphere, depending only upon the grain shape and the static (zero-frequency)
dielectric constant e 0 of the grain material (Purcell [55]; Draine [18]). Since Cext is the
sum of the absorption Cabs and scattering Cabs cross sections both of which are positive
numbers at all wavelengths, replacing Cext by Cabs in the left-hand-side of Eq.(3) would
give a lower limit on its right-hand-side; therefore we can write Eq.(3) as
k abs(l ) dl < 3p 2F/r
.
Z
0
(4)
It is immediately seen in Eq.(4) that b should be larger than 1 for l → ¥
since F is a
finite number and the integration in the left-hand-side of Eq.(4) should be convergent
(also see Emerson [23]), although we cannot rule out b ≤ 1 over a finite range of
wavelengths.
Astronomers often use the opacity k abs of the formula described in Eq.(1) to fit the
far-IR, submm and mm photometric data and then estimate the dust mass of interstellar
clouds:
,
(5)
d2Fl
mdust =
k abs(l ) Bl (T )
¥
¥
where d is the distance of the cloud, T is the dust temperature, Fl
is the measured flux
density at wavelength l . By fitting the photometric data points, one first derives the
best-fit parameters b , l 0, and T . For a given k 0, one then estimates mdust from Eq.(5).
In this way, various groups of astronomers have reported the detection of appreciable
amounts of very cold dust (T < 10 K) both in the Milky Way and in external galaxies
(Reach et al. [56]; Chini et al. [14]; Krügel et al. [35]; Siebenmorgen, Krügel, & Chini
[59]; Boulanger et al. [10]; Popescu et al. [54]; Galliano et al. [24]; Dumke, Krause, &
Wielebinski [20]).
As can be seen in Eq.(5), if the dust temperature is very low, one then has to invoke
a large amount of dust to account for the measured flux densities. This often leads to
too large a dust-to-gas ratio to be consistent with that expected from the metallicity of
the region where the very cold dust is detected (e.g. see Dumke et al. [20]), unless the
opacity k abs is very large. It has been suggested that such a large k abs can be achieved
by fractal or porous dust (Reach et al. [56]; Dumke et al. [20]). Can k abs be really so
large for physically realistic grains? At a first glance of Eq.(4), this appears plausible if
the dust is sufficiently porous (so that its mass density r
is sufficiently small). However,
one should keep in mind that for a porous grain, the decrease in r will be offset by a
decrease in F because the effective static dielectric constant e 0 becomes smaller when
the dust becomes porous, leading to a smaller F factor (see Fig. 1 in Purcell [55] and
Fig. 15 in Draine [18]).
Similarly, if one would rather use Qabs/a of Eq.(2) instead of k abs of Eq.(1), we can
also apply the Kramers-Kronig relation to place (1) a lower limit on b -- b cannot be
smaller than or equal to 1 at all wavelengths, and (2) an upper limit on (Qabs/a)0 from
(Qabs/a) (l ) dl < 4p 2F .
Z
0
(6)
Finally, the best-fit parameters b , l 0, and T should be physically reasonable. This can be
checked by comparing the best-fit temperature T with the dust equilibrium temperature
Td calculated from the energy balance between absorption and emission
Z
0
k abs(l ) cul dl = Z
0
k abs(l )4p Bl (Td) dl
,
(7)
where c is the speed of light, and ul
is the energy density of the radiation field.
Alternatively, one can check whether the strength of the radiation field required by
Td ≈ T is in good agreement with the physical conditions of the environment where
the dust is located.
3. OPACITY: WAVELENGTH-DEPENDENCE EXPONENT INDEX
It is seen in §2 that the Kramers-Kronig relation requires b > 1 for l → ¥
. For the
Milky Way diffuse ISM, the 100 m m -- 1 mm dust emission spectrum obtained by the Dif-
fuse Infrared Background Experiment (DIRBE) instrument on the Cosmic Background
Explorer (COBE) satellite is well fitted by the product of a single Planck curve of
T ≈ 17.5 K and an opacity law characterized by b ≈ 2 (i.e. k abs (cid:181)
l −2; Boulanger et
¥
¥
¥
Smaller b
al. [10]), although other sets of T and b are also able to provide (almost) equally good
fits to the observed emission spectrum (e.g. T ≈ 19.5 K and b ≈ 1.7; see Draine [17]).
But we should emphasize here that it is never flatter than b ≈ 1.65 (Draine [17]).
in the submm and mm wavelength range has been reported for cold
molecular cores (e.g. Walker et al. [63]: b ≈ 0.9 − 1.8), circumstellar disks around
young stars (e.g. Beckwith & Sargent [4]: b ≈ −1 -- 1; Mannings & Emerson [47]:
b ≈ 0.6; Koerner, Chandler, & Sargent [28]: b ≈ 0.6), and circumstellar envelopes
around evolved stars (e.g. Knapp, Sandell, & Robson [27]: b ≈ 0.9) including the
prototypical carbon star IRC + 10126 (Campbell et al. [13]: b ≈ 1).
However, these results are not unique since the dust temperature and density gradients
in the clouds, disks or envelopes have not been taken into account in deriving b --
the b exponent was usually estimated by fitting the submm and mm spectral energy
distribution by a modified black-body l −b Bl (T ) under the "optically-thin" assumption,
with b and T as adjustable parameters. If the dust spatial distribution is not constrained,
the very same emission spectrum can be equally well fitted by models with different b
values.
l −2) is expected for
both dielectric and conducting spherical grains: in the Rayleigh regime (where the
wavelength is much larger than the grain size)
As a matter of fact, an asymptotic value of b ≈ 2 (i.e. k abs (cid:181)
k abs ≈
18p
lr
e 2
(e 1 + 2)2 + e 2
2
,
(8)
where e (l ) = e 1 + ie 2 is the complex dielectric function of the grain at wavelength l .
For dielectric spheres, k abs (cid:181)
since e 1 approaches a constant (≫
e 2) while e 2 (cid:181)
l −2
as l → ¥
l and e 1 ≪ e 2. Even for dielectric grains with such an
extreme shape as needle-like prolate spheroids (of semiaxes l along the symmetry axis
and a perpendicular to the symmetry axis), in the Rayleigh regime we expect k abs (cid:181)
l −2:
l −1; for metallic spheres with a conductivity of s , k abs (cid:181)
since e 2 = 2ls
l −2 as l → ¥
l −1e −1
2
l −1e 2 (cid:181)
/c (cid:181)
(9)
k abs ≈
2p
3lr
e 2
(cid:2)Lk(e 1 − 1) + 1(cid:3)2
+(cid:0)Lk
e 2(cid:1)2
where Lk ≈ (a/l)2 ln(l/a) is the depolarization factor parallels to the symmetry axis;
e 2, therefore k abs (cid:181)
since for dielectric needles e 1 → constant and Lk(e 1 − 1) + 1 ≫ Lk
l −2 (see Li [36] for a detailed discussion). Only for both conducting and
l −1e 2 (cid:181)
extremely-shaped grains k abs can still be large at very long wavelengths. But even
for those grains, the Kramers-Kronig relation places an upper limit on the wavelength
range over which large k abs can be attainable (see Li & Dwek [43] for details). The
Kramers-Kronig relation has also been applied to interstellar dust models to see if the
subsolar interstellar abundance problem can be solved by fluffy dust (Li [40]) and to TiC
nanoparticles to relate their UV absorption strength to their quantities in protoplanetary
nebulae (Li [37]).
The inverse-square dependence of k abs on wavelength derived above applies to both
crystalline and amorphous materials (see Tielens & Allamandola [61] for a detailed dis-
cussion). Exceptions to this are amorphous layered materials and very small amorphous
(cid:181)
grains in both of which the phonons are limited to two dimensions and their phonon
spectrum is thus proportional to the frequency. Therefore, for both amorphous layered
materials and very small amorphous grains the far-IR opacity is in inverse proportional
to wavelength, i.e. k abs (cid:181)
l −1 (Seki & Yamamoto [58]; Tielens & Allamandola [61]).
Indeed, the experimentally measured far-IR absorption spectrum of amorphous carbon
shows a k abs (cid:181)
l −1 dependence at 5 m m < l < 340 m m (Koike, Hasegawa, & Manabe
[29]). If there is some degree of cross-linking between the layers in the amorphous lay-
ered grains, we would expect 1 < b < 2 (Tielens & Allamandola [61]). This can explain
the experimental far-IR absorption spectra of layer-lattice silicates which were found to
have 1.25 < b < 1.5 at 50 m m < l < 300 m m (Day [15]). For very small amorphous
grains, if the IR absorption due to internal bulk modes (for which the density of states
frequency spectrum is proportional to l −2) is not negligible compared to that due to
surface vibrational modes (for which the frequency spectrum is proportional to l −1),
we would also expect 1 < b < 2 (Seki & Yamamoto [57]).
If there exists a distribution of grain sizes, ranging from small grains in the Rayleigh
regime for which b ∼ 2 and very large grains in the geometric optics limit for which
b ∼ 0, we would expect b
to be intermediate between 0 and 2. This can explain the
small b values of dense regions such as molecular cloud cores, protostellar nebulae and
protoplanetary disks where grain growth occurs (e.g. see Miyake & Nakagawa [50]).
is temperature-dependent, as measured by Agladze et al. [1]
for silicates at T = 1.2 -- 30 K and l = 700 -- 2900 m m, and by Mennella et al. [49] for
silicates and carbon dust at T = 24 -- 295 K and l = 20 -- 2000 m m. Agladze et al. [1] found
that at T = 1.2 -- 30 K, b first increases with increasing T , after reaching a maximum at
T ∼ 10 -- 15 K it starts to decrease with increasing T . Agladze et al. [1] attributed this
to a two-level population effect (Bösch [9]): because of the temperature-dependence of
the two-level density of states (i.e. the variation in temperature results in the population
change between the two levels), the exponent index b
is also temperature dependent. In
contrast, Mennella et al. [49] found that b
increases by ∼10% -- 50% from T = 295 K
to 24 K, depending on the grain material (e.g. the variation of b with T for crystalline
silicates is not as marked as for amorphous silicates). The increase of b with decreasing
T at T = 24 -- 295 K is due to the weakening of the long wavelength absorption as T
decreases because at lower temperatures fewer vibrational modes are activated. Finally,
it is noteworthy that the inverse temperature dependence of b has been reported by
Dupac et al. [21] for a variety of regions.
The exponent index b
4. OPACITY: ABSOLUTE VALUES
In literature, one of the widely adopted opacities is that of Hildebrand [26]:
k abs (cm2 g−1) ≈(cid:26) 2.50 × 103 (l /m m)−1 , 50 m m < l ≤ 250 m m ,
l > 250 m m .
6.25 × 105 (l /m m)−2 ,
the above values from first estimating k abs(125m m)
Hildebrand [26] arrived at
and then assuming b ≈ 1 for l < 250 m m and b ≈ 2 for l > 250 m m. He
estimated the 125 m m opacity from k abs(125m m) = 3Qabs(125m m)/ (4ar ) =
(10)
3/ (4ar ) [Qabs(125m m)/Qext(UV)] Qext(UV) by taking a = 0.1 m m,
r = 3 g cm−3,
Qext(UV) = 3, and Qext(UV)/Qabs(125m m) = 4000, where Qext(UV) is the ultraviolet
extinction efficiency at l ∼0.15 -- 0.30 m m.
The most recent silicate-graphite-PAHs interstellar dust model for the diffuse ISM (Li
& Draine [41]) gives
k abs (cm2 g−1) ≈(cid:26) 2.92 × 105 (l /m m)−2 ,
20 m m < l ≤ 700 m m ,
3.58 × 104 (l /m m)−1.68 , 700 m m < l ≤ 104 m m ,
(11)
while k abs(l ) ≈ 4.6 × 105 (l /m m)−2 cm2 g−1 for the classical Draine & Lee [19]
silicate-graphite model. The fluffy composite dust model of Mathis & Whiffen [48]
has k abs(l ) ≈ 2.4 × 105 (l /m m)−1.6 cm2 g−1 in the wavelength range of 100 m m <
l < 1000 m m. The silicate core-carbonaceous mantle dust model of Li & Greenberg
[44] gives k abs(l ) ≈ 1.8 × 105 (l /m m)−2 cm2 g−1 for 30 m m < l < 1000 m m. We see
that these k abs values differ by over one order-of-magnitude; e.g., the 1350 m m opacity
calculated from the composite model [k abs(1350m m) ≈ 2.35 cm2 g−1; Mathis & Whif-
fen [48]] is higher than that from the silicate-graphite-PAHs model [k abs(1350m m) ≈
0.20 cm2 g−1; Li & Draine [41]] by a factor of ∼12.
While the Mathis & Whiffen [48] composite dust model predicts an IR emission
spectrum too flat to be consistent with the COBE-FIRAS observational spectrum, and
the dust IR emission has not been calculated for the Li & Greenberg [44] core-mantle
model which focuses on the near-IR to far-UV extinction and polarization, the silicate-
graphite-PAHs model has been shown successful in reproducing the infrared emission
spectra observed for the Milky Way (Li & Draine [41]), the Small Magellanic Cloud (Li
& Draine [42]), and the ringed Sb galaxy NGC 7331 (Regan et al. [57], Smith et al. [60]).
Therefore, at this moment the dust opacity calculated from the silicate-graphite-PAHs
model is preferred.
It has recently been suggested that the long wavelength opacity can be estimated
from the comparison of the visual or near-IR optical depth with the (optically thin)
far-IR, submm and mm dust emission measured for the same region with high angular
resolution, assuming that both the short wavelength extinction and the long wavelength
emission are caused by the same dust (e.g. see Alton et al. [2,3], Bianchi et al. [5,6],
Cambrésy et al. [12], Kramer et al. [30,31])
t V
S(l )
=
Qext(V )
Qabs(l )
2.2 × 10−18
Bl (T )
(12)
where t V is the visual optical depth, S(l ) is the surface brightness at wavelength l ,
and Qext(V ) is the extinction efficiency in the V -band (l = 5500 Å). With the dust
temperature T determined from a modified black-body l −b Bl (T ) fit to the far-IR dust
emission spectrum (b
is not treated as a free parameter but taken to be a chosen number),
Qabs(l ) can be calculated from the far-IR, submm, or mm surface density S(l ), and the
measured visual optical depth t V [if what is measured is the near-IR color-excess, say
E(H − K), instead of t V , one can derive t V from t V ≈ 14.6E(H − K)]. In so doing,
Qext(V ) is usually taken to be ≈ 1.5.
The long wavelength k abs values recently determined using this method (Eq.[12]) are
generally higher than those predicted from the dust models for the diffuse ISM. Although
this can be explained by the fact that we are probably looking at dust in dense regions
where the dust has accreted an ice mantle and coagulated into fluffy aggregates for which
a higher k abs is expected (e.g. see Krügel & Siebenmorgen [34], Pollack et al. [52],
Ossenkopf & Henning [51], Henning & Stognienko [25], Li & Lunine [46]), the method
itself is subject to large uncertainties: (1) the grains responsible for the visual/near-IR
extinction may not be the same as those responsible for the far-IR, submm and mm
emission; the latter is more sensitive to large grains while the former is dominated
by submicron-sized grains; (2) the dust temperature T may have been underestimated
if the actual b
is larger than chosen; and (3) the fact that in many cases the IRAS
(Infrared Astronomical Satellite) 60 m m photometry was included in deriving the dust
temperature T results in appreciable uncertainties since the 60 m m emission is dominated
by stochastically heated ultrasmall grains; ignoring the temperature distributions of
those grains would cause serious errors in estimating the dust mass (see Draine [16])
and therefore also in deriving the long wavelength opacity k abs. These problems could
be solved by a detailed radiative transfer treatment of the interaction of the dust with
starlight (e.g. Popescu et al. [53], Tuffs et al. [62]) together with a physical interstellar
dust model (e.g. the silicate-graphite-PAHs model; see Li & Draine [41,42]).
Based on the laboratory measurements of the far-IR and mm absorption spectra of
both amorphous and crystalline silicates as well as disordered carbon dust as a function
of temperature, Agladze et al. [1] and Mennella et al. [49] found that not only the
wavelength dependence exponent index b but also the absolute values of the absorption
are temperature dependent: the far-IR and mm opacity systematically decreases (almost
linearly) with decreasing temperature to T ∼10 -- 15 K and then increases with decreasing
temperature at very low temperature. While the linear dependence of k abs on T at
T > 10 -- 15 K was interpreted by Mennella et al. [49] in terms of two-phonon difference
processes, the inverse-temperature dependence of k abs on T at very low temperature was
attributed to a two-level population effect (Agladze et al. [1]). Agladze et al. [1] and
Mennella et al. [49] also found that the far-IR and mm opacity of amorphous materials
are larger than that of their crystalline counterparts. This is because for amorphous
materials, the loss of long-range order of the atomic arrangement leads to a relaxation of
the selection rules that govern the excitation of vibrational modes so that all modes are
infrared active, while for crystalline solids, only a small number of lattice vibrations are
active.
5. SUMMARY
The wavelength-dependent mass absorption coefficient (opacity k abs) is a critical param-
eter in the determination of the total dust mass of IR-emitting dusty regions. The dust
opacity shows marked variation with local conditions. Due to the incomplete under-
standing of the size, shape, composition and structure of dust grains, our knowledge of
the long wavelength dust opacity is subject to large uncertainties. We apply the Kramers-
Kronig relation to place a lower limit on the exponent index b and an upper limit on the
absolute value of the opacity (§2), if the dust opacity is described as a power-law func-
tion of wavelength (k abs ∼ l −b ). Our current knowledge of the wavelength dependence
exponent index b
(§3) and the absolute values of the opacity (§4) of interstellar dust is
summarized in the context of interstellar grain models, laboratory measurements, and
direct comparison of the short-wavelength extinction with the long wavelength thermal
emission.
ACKNOWLEDGMENTS. I am grateful to C.C. Popescu and R.J. Tuffs for inviting me to give
an invited talk at this stimulating conference from which I learned much. I also thank E. Dwek,
C.C. Popescu, R.J. Tuffs, A.N. Witt, and E.M. Xilouris for helpful discussions.
REFERENCES
1. Agladze, N.I., Sievers, A.J., Jones, S.A., Burlitch, J.M., & Beckwith, S.V.W. 1996, ApJ, 462, 1026
2. Alton, P.B., Xilouris, E.M., Bianchi, S., Davies, J., & Kylafis, N. 2000, A&A, 356, 795
3. Alton, P.B., Xilouris, E.M., Misiriotis, A., Dasyra, K.M., & Dumke, M. 2004, A&A, 425, 109
4. Beckwith, S.V.W., & Sargent, A.I. 1991, ApJ, 381, 250
5. Bianchi, S., Davies, J.I., & Alton, P.B. 1999, A&A, 344, L1
6. Bianchi, S., Gonçalves, J., Albrecht, M., Caselli, P., Chini, R., Galli, D., & Walmsley, M. 2003, A&A,
7. Block, D.L., Puerari, I., Frogel, J.A., Eskridge, P.B., Stockton, A., & Fuchs, B. 1999, Ap&SS, 269, 5
8. Bohren, C.F., & Huffman, D.R. 1983, Absorption and Scattering of Light by Small Particles (New
399, L43
York: Wiley)
9. Bösch, M.A. 1978, Phys. Rev. Lett., 40, 879
10. Boulanger, F., Bourdin, H., Bernard, J.P., & Lagache, G. 2002, in EAS Publ. Ser., Vol. 4, Infrared and
Submillimeter Space Astronomy, ed. M. Giard, J.P. Bernard, A. Klotz, & I. Ristorcelli (Paris: EDP
Sciences), 151
11. Boulanger, F., Abergel, A., Bernard, J.P., Burton, W.B., Désert, F.X., Hartmann, D., Lagache, G., &
Puget, J.L. 1999, A&A, 312, 256
12. Cambrésy, L., Boulanger, F., Lagache, G., & Stepnik, B. 2001, A&A, 375, 999
13. Campbell, M.P., et al. 1976, ApJ, 208, 396
14. Chini, R., Krügel, E., Lemke, R., & Ward-Thompson, D. 1995, A&A, 295, 317
15. Day, K.L. 1976, ApJ, 210, 614
16. Draine, B.T. 1990, in The Interstellar Medium in Galaxies, ed. H.A. Thronson & J.M. Shull (Dor-
17. Draine, B.T. 1999, in 3 K Cosmology, ed. L. Maiani, F. Melchiorri, & N. Vittorio (Woodbury: AIP),
18. Draine, B.T. 2003, in The Cold Universe, Saas-Fee Advanced Course Vol. 32, ed. D. Pfenniger
drecht: Kluwer), 483
283
(Berlin: Springer-Verlag), 213
19. Draine, B.T., & Lee, H.M. 1984, ApJ, 285, 89
20. Dumke, M., Krause, M., & Wielebinski, R. 2004, A&A, 414, 475
21. Dupac, X., et al. 2003, A&A, 404, L11
22. Dwek, E. 2005, in AIP Conf. Proc., The Spectral Energy Distribution of Gas-Rich Galaxies: Con-
fronting Models with Data, ed. C.C. Popescu & R.J. Tuffs (Woodbury: AIP), in press
23. Emerson, J.P. 1988, in Formation and Evolution of Low-Mass Stars, ed. A.K. Dupree & M. T. Lago
24. Galliano, F., Madden, S.C., Jones, A.P., Wilson, C.D., Bernard, J.-P., & Le Peintre, F. 2003, A&A,
(Dordrecht: Kluwer), 21
407, 159
25. Henning, Th., & Stognienko, R. 1996, A&A, 311, 291
26. Hildebrand, R.H. 1983, QJRAS, 24, 267
27. Knapp, G.R., Sandell, G., & Robson, E.I. 1993, ApJS, 88, 173
28. Koerner, D.W., Chandler, C.J., & Sargent, A.I. 1995, ApJ, 452, L69
29. Koike, C., Hasegawa, H., & Manabe, A. 1980, Ap&SS, 67, 495
30. Kramer, C., Alves, J., Lada, C., Lada, E., Sievers, A., Ungerechts, H., & Walmsley, M. 1998, A&A,
329, L33
31. Kramer, C., Richer, J., Mookerjea, B., Alves, J., & Lada, C. 2003, A&A, 399, 1073
32. Kramers, H.A. 1927, Atti. Congr. Intern. Fisici. Como, 2, 545
33. Kronig, R. 1926, J. Opt. Soc. Amer., 12, 547
34. Krügel, E., & Siebenmorgen, R. 1994, A&A, 288, 929
35. Krügel, E., Siebenmorgen, R., Zota, V., & Chini, R. 1998, A&A, 331, L9
36. Li, A. 2003a, ApJ, 584, 593
37. Li, A. 2003b, ApJ, 599, L15
38. Li, A. 2004a, in Penetrating Bars Through Masks of Cosmic Dust, ed. D.L. Block, I. Puerari, K.C.
Freeman, R. Groess, & E.K. Block (Dordrecht: Kluwer), 535
39. Li, A. 2004b, in ASP Conf. Ser. 309, Astrophysics of Dust, ed. A.N. Witt, G.C. Clayton, & B.T.
Draine (San Francisco: ASP), 417
40. Li, A. 2005, ApJ, 622, 000
41. Li, A., & Draine, B.T. 2001, ApJ, 554, 778
42. Li, A., & Draine, B.T. 2002, ApJ, 576, 762
43. Li, A., & Dwek, E. 2005, to be submitted to ApJ
44. Li, A., & Greenberg, J.M. 1997, A&A, 323, 566
45. Li, A., & Greenberg, J.M. 2003, in Solid State Astrochemistry, ed. V. Pirronello, J. Krelowski, & G.
Manicó (Dordrecht: Kluwer), 37
46. Li, A., & Lunine, J.I. 2003, ApJ, 590, 368
47. Mannings, V., & Emerson, J.P. 1994, MNRAS, 267, 361
48. Mathis, J.S., & Whiffen, G. 1989, ApJ, 341, 808
49. Mennella, V., Brucato, J.R., Colangeli, L., Palumbo, P., Rotundi, A., & Bussoletti, E. 1998, ApJ, 496,
50. Miyake, K., & Nakagawa, Y. 1993, Icarus, 106, 20
51. Ossenkopf, V., & Henning, Th. 1994, A&A, 291, 943
52. Pollack, J.B., Hollenbach, D., Beckwith, S., Simonelli, D.P., Roush, T., & Fong, W. 1994, ApJ, 421,
1058
615
53. Popescu, C.C., Misiriotis, A., Kylafis, N.D., Tuffs, R.J., & Fischera, J. 2000, A&A, 362, 138
54. Popescu, C.C., Tuffs, R.J., Völk, H.J., Pierini, D., & Madore, B.F. 2002, ApJ, 567, 221
55. Purcell, E.M. 1969, ApJ, 158, 433
56. Reach, W.T., et al. 1995, ApJ, 451, 188
57. Regan, M.W., et al. 2004, ApJS, 154, 204
58. Seki, J., & Yamamoto, T. 1980, Ap&SS, 72, 79
59. Siebenmorgen, R., Krügel, E., & Chini, R. 1999, A&A, 351, 495
60. Smith, J.D., et al. 2004, ApJS, 154, 199
61. Tielens, A.G.G.M., & Allamandola, L.J. 1987, in Interstellar Processes, ed. D. Hollenbach & H.A.
Thronson, Jr. (Dordrecht: Reidel), 397
62. Tuffs, R.J., Popescu, C.C., Völk, H.J., Kylafis, N.D., & Dopita, M.A. 2004, A&A, 419, 821
63. Walker, C.K., Adams, F. & Lada, C.J. 1990, ApJ, 349, 515
|
astro-ph/9812389 | 1 | 9812 | 1998-12-21T17:02:20 | Tidal disruption rates of stars in observed galaxies | [
"astro-ph"
] | We derive the rates of capture, Ndot, of main sequence turn off stars by the central massive black hole in a sample of galaxies from Magorrian et al. 1998. The disruption rates are smaller than previously believed with Ndot ~ 10^-4 - 10^-7 per galaxy. A correlation between Ndot and black hole mass, M, is exploited to estimate the rate of tidal disruptions in the local universe. Assuming that all or most galaxies have massive black holes in their nuclei, this rate should be dominated by sub-Lstar galaxies. The rate of tidal disruptions could be high enough to be detected in supernova (or similar) monitoring campaigns---we estimate the rate of tidal disruptions to be 0.01 - 0.1 times the supernova rate. We have also estimated the rates of disruption of red giants, which may be significant (Ndot ~> 10^-4 y^-1 per galaxy) for M ~> 10^8 Msun, but are likely to be harder to observe---only of order 10^-4 times the supernova rate in the local universe. In calculating capture rates, we advise caution when applying scaling formulae by other authors, which are not applicable in the physical regime spanned by the galaxies considered here. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- 9 (1998)
Printed 14 August 2018
(MN LATEX style file v1.4)
Tidal disruption rates of stars in observed galaxies
D. Syer and A. Ulmer ⋆
Max-Planck-Institut fur Astrophysik Karl-Schwarzschild-Strasse 1, 85748 Garching, Germany
Accepted ........ Received .......; in original form .......
8
9
9
1
c
e
D
1
2
1
v
9
8
3
2
1
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
ABSTRACT
We derive the rates of capture N of main sequence turn off stars by the central massive
black hole in a sample of galaxies from Magorrian et al . 1998. The disruption rates are
smaller than previously believed with N ∼ 10−4
− 10−7 per galaxy. A correlation be-
tween N and black hole mass M is exploited to estimate the rate of tidal disruptions
in the local universe. Assuming that all or most galaxies have massive black holes
in their nuclei, this rate should be dominated by sub-L∗ galaxies. The rate of tidal
disruptions could be high enough to be detected in supernova (or similar) monitor-
ing campaigns -- we estimate the rate of tidal disruptions to be 0.01 − 0.1 times the
supernova rate. We have also estimated the rates of disruption of red giants, which
may be significant ( N >
∼ 108M⊙, but are likely to be
harder to observe -- only of order 10−4 times the supernova rate in the local universe.
In calculating capture rates, we advise caution when applying scaling formulae by
other authors, which are not applicable in the physical regime spanned by the galaxies
considered here.
∼ 10−4y−1 per galaxy) for M >
Key words: galaxies: nuclei
1
INTRODUCTION
A number of early type galaxies and spiral bulges are
now thought to contain massive black holes in their nuclei
(Magorrian et al . 1998). Direct evidence is also now available
which supports the idea that active galaxies are powered by
massive black holes (Reynolds & Fabian 1997, Fabian et al .
1998). It can be argued that many if not all galaxies could be
expected to have undergone an active phase and to possess a
massive black hole (Haehnelt & Rees 1993). Magorrian et al .
measure black hole masses in the range M = 106 − 109M⊙,
and report that M is correlated with the mass M of the
underlying hot stellar system, with M/M = x ≈ 0.006.
A main sequence star of roughly solar type can be
tidally disrupted by a black hole with mass M <∼ 2 × 108M⊙
(Hills 1975). Larger black holes swallow main sequence stars
whole, but red giants are susceptible to tidal disruption be-
cause of their lower density. Frank & Rees (1976) (FR),
Young, Shields & Wheeler (1977) considered a system com-
posed of a black hole and an isothermal sphere of stars and
derived analytic expressions for the rate of capture of stars.
Cohn & Kulsrud (1978) (CK) were able to calibrate FR
using a more detailed numerical calculation. Estimates of
capture rates in the literature are often based on equation
(66) of CK (e.g. Cannizzo, Lee & Goodman 1990, Kormendy
& Richstone 1995). We argue that such estimates are often
⋆ E-mail: (syer, aulmer)@mpa-garching.mpg.de
c(cid:13) 1998 RAS
wrong because the physical conditions in galactic nuclei are
different from those anticipated by CK (who were in any
case more concerned with globular clusters).
The next section summarises briefly what is known
about the observational signatures of tidal disruption. Sec-
tion 3 establishes notation by reviewing the analytic theory
of FR and sets out our method for calculating capture rates
in the Nuker galaxies. We calibrate our calculation against
the result of CK. In Section 4 we describe the data and list
the quantities we derive from them, including capture rates.
We discuss our results in Section 5.
2 OBSERVABILITY
The observable consequences of such an event have been dis-
cussed by Rees (1988), Evans & Kochanek (1989), Canizzo,
Lee & Goodman (1990) and Ulmer (1997). The general ex-
pectation for the disruption of a main sequence star is a
flare with bolometric luminosity near the Eddington limit.
In comparison, the luminosity from the central r < 1 arcsec
of all the Nuker galaxies is small compared to the Edding-
ton luminosity of the black hole (<∼ 10−4LEdd, Table 1). The
bolometric luminosity should be close to the Eddington limit
because of the predicted mass accretion rates are above the
Eddington accretion rate for black holes with mass up to
about 3 × 107 M⊙. Above that the predicted mass accretion
rate divided by the Eddington mass accretion rate falls as
M 3/2 (Ulmer 1997). There is much uncertainty as to how
2
optically bright a tidal disruption event will be, in optical,
but the minimum luminosity expected for typical disrup-
tions of main sequence stars is 10−3−10−2LEdd for the U
and V bands (Ulmer 1997). Typical durations of the bright-
est phase will be 0.1−2 years depending on the black hole
mass. Main sequence tidal disruption events should there-
fore be relatively easy to detect as a byproduct of supernova
searches (Section 4.1) or other searches which observe the
same galaxies many times over (e.g. Southern strip of the
SDSS).
Red giant disruption events are generally much longer
lasting, and therefore fainter, than main sequence events.
Red giants typically are disrupted by growing onto the loss
cone (Section 3.2), so at disruption, they have pericenter
equal to the tidal disruption radius. Consequently, an aver-
age red giant (with R ∼ 100R⊙) disruption has a approxi-
mate duration of 100M 1/2
years and a produces a peak mass
accretion rate no larger than ∼ 10−3 M −1/2
6 M⊙ yr−1 (see
Ulmer 1997, eqs 3,5). The Eddington accretion rate, which
would supply enough mass to provide the Eddington lumi-
nosity, is 3 × 10−2M6ǫ−1
.1 , where 0.1 × ǫ.1 is the efficiency of
rest mass to light conversion in the accretion process. This
means that red giant disruption could keep a small, 106 M⊙
black hole illuminated at (1/30th) its Eddington rate for up
to a thousand years.
6
Just how bright a red giant disruption would appear
in optical bands is difficult to determine, but a simple es-
timate is as follows. If most of the energy is released from
within a few Schwarzschild radii for both the main sequence
and red giant disruptions, and the bolometric luminosity of
the red giant disruptions is 1/30 that of the main sequence
disruptions for a 106 M⊙ black hole, then
1
30
∼
LRG
LMS
∼
T 4
T 4
eff,RG
eff,MS
.
(1)
Therefore, the ratio of temperatures is (1/30)1/4 ∼ 0.4. Be-
cause the optical bands lie in the Rayleigh-Jeans tail of the
spectrum where the luminosity scales linearly with T , we
would at first guess expect that red giant disruptions would
be nearly (i.e. ∼ 0.4 times) as bright as main sequence
events. Around more massive black holes with M ∼ 108M⊙,
red giants may be significantly less luminous if the accre-
tion becomes advection dominated as appears to be the case
for many extra-galactic black holes with accretion rates far
below the Eddington accretion rate (e.g. NGC 4258 (La-
sota et al . 1996), M87 (Reynolds et al . 1996)). Around very
massive black holes with M ∼ 1010M⊙, the timescale for
return of the disruption material to the black hole becomes
comparable with the interval between successive tidal dis-
ruptions (see Figure 4). Red giant disruptions would then
provide a small, nearly constant flow of material onto the
black hole of ∼ 10−3M⊙yr−1 corresponding to a luminosity
of ∼ 10−5ǫ.1LEdd
3 THEORY OF CAPTURE RATES
Consider a spherical cluster of stars of mass m∗ with density
ρ(r), and isotropic velocity dispersion σ(r). Typically ρ(r) is
in the form of two power laws, separated by a break radius
rb. The velocity dispersion is then given by Jeans' equations:
dρσ2
dr
+
GρM(r)
r2
= 0
(2)
where M(r) is the mass enclosed within radius r, including
the black hole, and G is Newton's constant. The black hole
exerts an influence on the stars inside a sphere of influence
with radius ra given by
GM
r
− σ(r)2 = 0,
(3)
where M is the black hole mass. Within the sphere of influ-
ence (r < ra)
σ2
≈
GM
(1 + p)r
,
(4)
where the logarithmic slope of the density is −p (i.e. ρ ∼
r−p, with p possibly dependent on radius). An important
derived quantity is the two-body relaxation timescale, given
by
tr =
σ3
ΛG2ρm∗
,
(5)
where m∗ is the typical stellar mass and Λ includes dimen-
sionless factors of order unity as well as the Coulomb loga-
rithm. We use the numerical value of tr given by Binney &
Tremaine 1987, equation (8-71). Since tr is not very sensitive
to Λ we do not attempt to include the dependence of Λ on
r. Instead we take
tr =
1.8 × 1012y
ln(0.4N ) (cid:16)
σ
100km s−1(cid:17)3(cid:18) M⊙
m∗ (cid:19)(cid:18) 104M⊙pc−3
ρ
(cid:19) ,(6)
where N is the number of stars within the characteristic
radius rb.
3.1 The loss cone
Suppose that stars are removed from the system if they
come within a distance q of the centre of the cluster. If
there is a massive black hole at the centre, q will be the
larger of the tidal radius rT ≈ R⋆(M/M⋆)1/3 (Hill 1975) or
the Schwarzschild radius of the black hole. For stars at the
main sequence turn off, q = rT when M <∼ 2 × 108M⊙. A
star at radius r will be removed if its angular momentum is
small enough. Such stars are said to populate a 'loss-cone'
(Lightman & Shapiro 1977, Young, Shields & Wheeler 1977,
FR), the angular size of which is given at radius r by
θ2
lc =
q
r2
GM
σ2
(7)
(provided σ2 ≪ M/q). Even at the edge of the sphere of
influence, the loss cone is quite small with θlc ∼ 10−7 −10−5.
The loss process is usually divided into two regimes ac-
cording to the angle θd through which a star is scattered in
a dynamical time
td = r/σ.
(8)
In the 'diffusive' regime θd < θlc, and on timescales much
longer than td the loss cone is empty, because a star which
finds itself in the loss cone is removed within a dynamical
time. In this case, the capture rate is limited by diffusion
into the loss cone, and is given, per star, by
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
d N<
dN
=
1
ln(2/θlc)tr
(Lightman & Shapiro 1977), where N (r) is the number of
stars contained within a radius r.
In the other regime, θd > θlc, the loss cone is always full.
Since we assume isotropic velocity dispersions, the fraction
2. Thus in this
of stars in the loss cone at any time is just θlc
case the capture rate per star is
(9)
and in case (ii)
N ≈ Ncrit =
N (rcrit)
tr(rcrit)
.
(10)
3.2 Red giant capture
d N>
dN
=
2
θlc
td
.
The rates (9) and (10) are equal at a radius r = rcrit, where
Λlc
M
m∗
Gρr3
σ2 = q
where we have written Λlc = ln(2/θlc)/Λ. The radial depen-
dence of Λlc is weak, so we set it to a constant equal to its
value at rb.
(11)
The first step in calculating the capture rate is to solve
equation (11) to find rcrit. Then we can find the total loss
N , by integrating equations (9) and (10) over the clus-
rate,
ter:
N = N< + N>,
with
N< = 4π
N> = 4π
rcrit
Z
0
∞
Z
rcrit
ΛlcG2ρ2r3
σ3
dr
r
GM qρ
m∗σ
dr
r
.
(12)
(13)
(14)
Before moving on, let us examine the various contribu-
tions to N in some more detail. Let us define f (r) to be the
r-dependence of the integrand in N<, so
f (r) =
G2ρ2r3
σ3
(15)
Now at small r, σ ∼ r−1/2, so f (r) → 0 as r → 0 provided
p < 9/4 (which is true in all the galaxies we consider). The
lower limit in the first integral in principle should be finite,
but for p < 9/4 we may set it to zero. At large radius,
σ2
σ2
∼ r2−p
∼ r−1/2
,
,
p < 3
p > 3
(16)
(17)
so f (r) → 0 as r → ∞ provided p > 0. Thus generally f (r)
will reach a maximum at some radius rmax. Quite generally
we expect that rmax will be approximately equal to ra (for
ra < rb or 0 < p < 9/4 at large r) or rb (otherwise). We can
identify two distinct regimes according to whether rcrit ><
rmax .
(i) If rcrit ≫ rmax then N> will be insignificant compared
to N<.
with N<.
(ii) If rcrit <∼ rmax then N> will be at most comparable
In case (i)
N ≈ Nmax =
N (rmax)
tr(rmax)
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
(18)
3
(19)
Equations (18) and (19) constitute a poor man's recipe for
calculating capture rates, and within a factor of order unity
are interchangeable with the full integral version in equation
(12).
For black hole masses >∼ 2 × 108M⊙, the Schwarzschild ra-
dius exceeds the tidal disruption radius for main sequence
stars. Red giants, as a result of their low densities, have much
much larger tidal radii, and can therefore be disrupted. Red
giants may enter the loss cone in the same manner as main
sequence stars: they diffuse towards an empty loss cone for
orbits below a critical radius, and above that radius, they
scatter onto a full loss cone. There is an additional contribu-
tion to the red giant capture rate. Red giants also grow onto
the loss cone as they expand. The loss cone for red giants is
larger than for main sequence stars (q is larger because they
are less dense), so even when the main sequence loss cone is
empty, red giants are being born inside their own loss cone.
While the total fraction of stars which are red giants at any
given time may be very small, the number of red giants sus-
ceptible to capture may be significant. We discuss each of
these capture channels below. In addition, throughout this
discussion we use the term red giant loosely to describe stars
on both the RGB and AGB.
3.2.1 Radius-time relationship
In order to calculate the capture rates, it is necessary to
understand how red giants evolve in radius as a function of
time because the loss cone, θlc ∝ R⋆ (7). A good estimate of
the radius evolution can be easily determined as the radius
and luminosity of a red giant are largely independent of the
stellar mass, and depend only on the core mass. Functional
forms for L(mc) and R(mc) are given in Joss, Rappaport &
Lewis (1987) , where mc is the core mass, which are good
approximations for both the red giant and ascending red
giant branches. Because the dominant energy source in red
giants is the p−p chain with ∼ 0.7% efficiency and the core
mass is increased as hydrogen burns to helium, we may write
L(mc) ≈ 0.007M⊙c2 ∂µ
∂t
105.3µ6
(20)
≈
1 + 100.4µ4 + 100.5µ5 L⊙,
(21)
where µ = mc/M⊙, and the right hand side comes from
Joss, Rappaport & Lewis (1987). Solving this equation for
core mass as a function of time, gives the luminosity-time
dependence. A similar expression to equation (21) yields the
radius-time dependence.
Applying this model between an initial core mass of
0.17 M⊙ and a final core mass of 0.8 M⊙ we find an ini-
tial radius of 3.09 R⊙ and a maximum radius, Rmax, of
679 R⊙ after 7 × 108 years. This maximum radius is in fact
larger than the radius predicted by stellar evolution codes,
because mass loss significantly alters the evolution, espe-
cially when in the star reaches the thermally pulsing AGB
4
of stars that are red giants (with radii between ∼ 3 and
200R⊙) is
fRG =
d Nnew
dN
tRG ≈ 0.05.
(26)
We therefore expect that the red giant capture rate will be
lower than the main sequence rate by no more than a fac-
tor of 20. Another important quantity is the time averaged
radius, <R> ∼ 12R⊙. The evolution above 100R⊙ makes
only a small (<10%) contribution to the time averaged ra-
dius.
3.2.3 Maximum radius limits
The red giant capture rates depend on the maximum radius
reached by the red giant. We consider how this radius may
be truncated by stellar collisions, and how this radius can
effectively be truncated when the red giant evolution time
becomes shorter than the dynamical time.
First, consider the frequency of collisions between a red
giant and a main sequence star:
Figure 1. Time as a function red giant radius (solid line). The
dashed line shows an approximation adopted in eq. 22.
tcoll
tr
≈
3.7 ln 0.4N θ2
1 + 2θ
,
(27)
phase (TPAGB). We therefore limit the red giant evolution
at 200R⊙ which corresponds to the onset of the TPAGB
(e.g. Bressan et al . 1992).
For our level of approximation of the red giant capture
rates, a simple analytical expression for the radius-time de-
pendence is sufficient. A course approximation (figure 1) is:
t
7 × 108years
≈ 0.38 ln(cid:18) R
R⊙
− 3(cid:19) + .37.
(22)
3.2.2 Red giant birth rate and number density
In order to calculate the red giant capture rates, we must
determine their birth rate and their number density.
The main sequence lifetime of a star of mass m is given
approximately by
tms ∼ 1010(cid:18) m
M⊙(cid:19)−2.5
y
(23)
and the total number of stars, assuming a Salpeter mass
function, is
dN
dm
≈ 1.6(cid:18) m
M⊙(cid:19)−1.35
,
(24)
where the total mass has been normalised to unity in the
range (0.3, 0.8)M⊙. Therefore the birth rate of red giants
(per star) is
d Nnew
dN
=
dN
dm (cid:12)(cid:12)(cid:12)
dm
dt (cid:12)(cid:12)(cid:12)
≈ 0.65 × 10−10(cid:18) m
M⊙(cid:19)2.15
y−1,
(25)
where θ is the Safronov number, Gm∗/2σ2R (see Binney &
Tremaine 1987, equation 126). for σ ∼ 100km s−1, θ varies
between about 5 and 0.01 as a red giant evolves. The col-
lisions are important if a red giant typically collides during
its lifetime:
Z Rmax
Rmin
dt
tcollR
≈ 1,
(28)
where Rmin ≈ 3R⊙. Using the relations from section 3.2.1
and setting ln(0.4N ) ∼ 20, we find that collisions are im-
portant when
tr <∼ 1.1 × 107(cid:18) 250km s−1
σ
(cid:19)3
years.
(29)
For the systems we consider here, the shortest relaxation
times (at ra) are many tens of times too large for collisions
to be important.
Second, at large system radii, the most extended red gi-
ants may evolve on a time faster than the dynamical time, so
that extended red giants have a reduced chance of encoun-
tering the black hole. Using the red giant evolution model
from Section 3.2.1, we calculate the characteristic evolution
time, R/ R as a function of radius (Figure 2). We find that
over the radii range 3−200R⊙, the relationship is well ap-
proximated by
tR ≡ R/ R ∼ 6.5 × 106(cid:18) R
200R⊙(cid:19)−1.12
.
(30)
The maximum radius for a red giant is therefore the smaller
of 200R⊙ and the radius at which tR = td. In practice for
the Nuker galaxies in the regions of interest tR is always
much greater than td, and we take Rmax = 200R⊙.
where m is now the main sequence turn off mass (the mass of
stars which are currently being transformed into red giants).
The total red giant lifetime tRG ∼ 7 × 108y, so the fraction
3.2.4 Capture analogous to main sequence capture
Main sequence capture occurs by diffusion onto an empty
loss cone at radii below the critical radius, and by scatter-
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
θ2
RG = θlc
2 Rmax
R⊙
,
5
(32)
where (as before) θlc is the angular size of the main sequence
loss cone. Assuming that all stars which become red giants
inside this loss cone will be captured, the rate of red giant
captures from a given radius (per star) is
d NRG
dN
= θ2
RG
d Nnew
dN
≈
460θlc
2
t0
≈ 4.6 × 10−8θlc
2y−1,
(33)
where t0 is the age of the galaxy, and we have chosen Rmax =
200R⊙.
3.2.6 Total capture rate
The rate of red giant capture per star is the sum of equations
(31) and (33):
d NRG
dN
=
R
R⊙
dt θlc
td
2
+
d Nnew
dN (cid:20)Z Rt
Z Rmax
Rt
Rmin
dt
ln(2/θlc)tr
+
Rmax
R⊙
θlc
2(cid:21) ,
(34)
which must be integrated over the system to obtain the total
capture rate
NRG = 4πZ ∞
0
d NRG
dN
ρr2dr.
(35)
For the Nuker galaxies, since rcrit is generally large for main-
sequence stars it is larger still for red giants, and Rt = Rmin.
We also find that growth onto the loss cone (equation 33)
dominates the red giant capture rate.
3.3 Calibration
FR considered a system in which the stars inside the sphere
of influence have relaxed to the 'zero-flow' distribution of
Bahcall & Wolf (1976) with p = 7/4, and outside ra the stars
are approximately isothermal with a homogeneous core of
radius r0 and density ρ0. One case they concentrated on was
that in which rcrit < ra < r0. CK also calculated the capture
rate (as well as the full anisotropic distribution function) in
a system with rcrit < ra and a relaxed cusp with p ≈ 1.8.
N in
They give a scaling relation (their equation 66) for
these circumstances which is substantially the same as FR
equation (16a), except for the normalisation. We use this
normalisation to calibrate our model capture rates.
4 THE NUKER GALAXIES
To calculate a capture rate, we need to know the density
profile ρ(r) and the black hole mass M . Given the observed
surface brightness I(r), we can derive ρ(r) by an Abel inver-
sion, assuming a constant mass-to-light ratio Υ. We do not
assume anything about the density profile in regions which
are not resolved by the observations, merely extrapolating
the profiles inwards. This will not affect the capture rates
provided min[rcrit, rmax] is not much smaller than the reso-
lution limit of the observations, which turns out to be the
case. Most of the data we use are quoted as a Nuker law fit to
I(r) (Byun et al . 1996). Thus I(r) is in the form of two power
Figure 2. Characteristic evolution time, R/ R as a function radius
for red giants (solid line). The dashed line shows an simple fit to
the model.
ing onto a full loss cone above the critical radius (equa-
tions 9, 10). In general, the larger the radius of the star,
the larger the critical radius, and at large radii, the critical
radius scales as the stellar radius (equations 11, 16). For red
giants we must define three regions because the red giant ra-
dius and therefore the critical radius are functions of time.
Below a radius rcrit,1 determined by solving equation (11)
with q = 3qms, the loss cone is depleted on a dynamical time
for all red giants, so the diffusion rate is appropriate. Beyond
the second radius rcrit,2 determined by solving equation (11)
with q = 200qms, the loss cone is full, so the full loss cone
rate is appropriate. Between these two radii, the full rate
is appropriate until the star reaches a size at which it can
scatter entirely into or out of the loss cone in one dynami-
cal time, i.e. the stellar radius Rt for which equation (11) is
satisfied. The rate of captures of red giants from these three
contributions can therefore be written:
d NRG
dN
=
d Nnew
dN (cid:20)Z Rt
Rmin
2
R
R⊙
dt θlc
td
+Z Rmax
Rt
dt
ln(2/θlc)tr(cid:21) .
(31)
The limits in the integrals are understood to mean the time
at which R is equal to the value specified. The critical red-
giant radius Rt is equal to Rmin for r < rcrit,1 and to Rmax
for r > rcrit,2.
3.2.5 Captures from growth onto the loss cone
There will be some maximum radius for red giants Rmax,
fixed either by stellar evolution (at ∼ 200R⊙), or by other
means e.g, the rate at which their envelopes are stripped
by collisions with other stars. As we discuss below, the ap-
propriate maximum radius is almost always set by stellar
evolution. This maximum radius defines a loss cone with
angular size
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
6
Table 1. Observational data and summary of results. Column (2) taken from Byun et al . 1996; Columns (3-4) from Magorrian et al .
1998. Column (1), name of galaxy; (2), rb in pc; (3), Nuker law indices (α, β, γ); (4) log10 M black hole mass; (5) ra/rb; (6) ra in
arcsec; (7) main-sequence capture time ( N /y); (8) red giant capture time ( NRG/y); (9) projected stellar luminosity in central arcsec
ℓ(1) = L(r < 1")/LEdd.
NGC
1399
1600
2832
3115
3377
3379
608
168
4467
4472
4486
4486
4552
4564
4621
4636
4649
4874
4889
6166
7332
7768
221
224
MW
rb
269.15
1258.93
398.11
117.49
4.37
83.18
27.54
446.68
239.88
177.83
562.34
13.49
47.86
38.90
218.78
239.88
263.03
1202.26
758.58
1202.26
75.86
199.53
1.47
288.40
0.38
(α, β, γ)
(1.50, 1.68, 0.07)
(1.98, 1.50, 0.08)
(1.84, 1.40, 0.02)
(1.47, 1.43, 0.78)
(1.92, 1.33, 0.29)
(1.59, 1.43, 0.18)
(1.05, 1.33, 0.00)
(0.95, 1.50, 0.14)
(7.52, 2.13, 0.98)
(2.08, 1.17, 0.04)
(2.82, 1.39, 0.25)
(2.78, 1.33, 0.14)
(1.48, 1.30, 0.00)
(0.25, 1.90, 0.05)
(0.19, 1.71, 0.50)
(1.64, 1.33, 0.13)
(2.00, 1.30, 0.15)
(2.33, 1.37, 0.13)
(2.61, 1.35, 0.05)
(3.32, 0.99, 0.08)
(4.25, 1.34, 0.90)
(1.92, 1.21, 0.00)
(2.00, 1.28, 0.53)
(1.12, 1.52, 0.33)
(1.80, 0.80, 0.00)
log M
9.72
10.06
10.06
8.55
7.79
8.60
8.39
9.08
7.44
9.42
9.55
8.96
8.67
8.40
8.45
8.36
9.59
10.32
10.43
10.45
6.84
9.96
6.36
7.79
6.42
ra
rb
0.38
0.09
0.51
0.07
0.97
0.22
0.54
0.19
0.04
0.17
0.14
9.95
0.24
0.37
0.03
0.07
0.29
0.39
0.64
0.68
0.01
1.25
0.50
0.04
1.23
ra(") − log N − log NRG − log ℓ(1)
1.19
0.48
0.46
0.19
0.09
0.38
0.15
0.48
0.13
0.41
1.06
1.81
0.15
0.20
0.10
0.24
1.04
1.04
1.07
1.50
0.01
0.50
0.19
2.80
11.38
6.13
6.60
6.41
4.80
4.80
5.76
5.54
6.47
5.36
6.05
6.32
5.90
5.57
5.24
4.74
6.24
6.19
6.93
6.72
7.16
4.10
6.34
4.53
5.95
4.32
4.23
3.92
4.09
4.46
5.10
5.00
5.10
5.00
5.76
4.20
4.39
5.02
4.75
4.90
4.52
5.35
4.32
4.13
3.96
4.18
4.81
4.17
5.27
5.76
5.07
4.92
5.02
4.36
3.47
2.99
4.02
3.51
4.11
3.13
4.93
5.21
4.38
3.80
3.34
3.09
4.07
5.11
5.12
4.89
5.42
1.31
4.25
3.05
5.41
6.80
)
2
c
e
s
c
r
a
/
g
a
m
(
V
µ
15
20
0.01
1.0
r (arcsec)
100.0
Figure 3. Deconvolved V-band surface density of M31 (tri-
angles) with Nuker law fits to the inner (r < 1.75") and outer
(r > 3.6") profiles (dotted lines).
laws, separated by a break radius rb, and ρ(r) has a simi-
lar form, with a break radius close to rb. The data we have
used are reproduced in Table 1, which also lists some inter-
esting derived quantities. M32 is listed as NGC 221, M31 is
NGC 224, and the Milky Way is 'MW'. The Milky Way data
is from the model by Genzel et al . 1996. The Nuker param-
eters for M32 are taken from Lauer et al . (1992). For M31
we fit the photometry ourselves using data from Lauer et al .
(1993) (r < 10") and Kent (1987) (r > 10") and matching
at r = 10". The resulting surface density profile is shown in
Figure 3, which also shows the Nuker law fits listed in Table
1. We fit the nuclear component (r < 1.75") separately, and
name it NGC 224N in Table 1. This component is not self
gravitating and is probably rotationally supported (Lauer
et al . 1993, Tremaine 1995), and the capture rate is there-
fore almost certainly much less than in our isotropic model.
We calculate the latter in order to satisfy ouselves that the
capture rate is not dominated by the nuclear component,
which is confirmed by the results. The scale of the flattened
component in M31 is much smaller than the physical scale
which dominates stellar capture rates. Similarly, in the other
galaxies the capture rate is dominated by a physical radius
which is well resolved, so unresolved substructure should not
affect our results.
Throughout this section we assume where necessary
that the mass of the black hole is proportional to the mass
of the galaxy (or spheroid) with constant of proportionality
≈ .006 (Magorrian et al . 1998), and the mass to light ratio
obeys the fundamental plane relation Υ ∝ L0.2. Thus the
black hole mass is
M ≈ 6 × 108(cid:16) L
L∗(cid:17)
1.2
.
4.1 Capture rates
(36)
Figure 4 shows the capture rates as a function of the black
hole mass. In our models all the Nuker galaxies have rcrit ≫
rb and all but two have rb >∼ ra, and hence rmax ≈ ra. This
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
10-2
10-4
)
1
−
y
(
N
10-6
10+7
10+8
10+9
10+10
M (M⊙)
N (y−1) for main sequence stars
Figure 4. The capture rates
(squares) and red giants (triangles) as a function of black hole
mass. The circles are the main sequence capture rates in the case
where loss cones are all full. The dotted line is the best fit to
the main sequence rates (with loss cone), and has slope −0.61.
The dashed line is the upper limit set by N m∗ = M/t0 with
t0 = 1010y.
accounts for why both the main sequence rate and the red
giant rate are correlated with M . In the red giant case, the
capture rate is proportional to the mass within ra which is
of order the black hole mass. In the main sequence case the
fuelling rate is proportional to this mass divided by tr(ra).
The fundamental plane leads to a correlation between tr(ra)
and M , and hence to the correlation in Figure 4. We can
Nmax equation (18), and the result is well
also calculate
N given in Table 1 (over a range
correlated with the value of
in N of 3 orders of magnitude Nmax is within a factor of two
N ). We can also use equation (18) to estimate what
or so of
the effect on N would be of errors in the determination of
the black hole mass. The result for main sequence stars and
empty loss cones is that errors in M tend to move N roughly
along the N −M relation in Figure 4 (dotted line).
4.1.1 Comparison with supernova rate
Main sequence capture is faster for smaller black holes, and
this leads us to expect that the total rate of capture in the
local universe will be dominated by the galaxies with small-
est black holes. To see this more clearly, let us compare the
rate of star capture with the supernova rate in the local
universe. We assume a galaxy population with luminosity
function of the form
dN
dL
∝
1
L (cid:16) L
L∗(cid:17)α
exp(cid:16) −L
L∗ (cid:17) ,
(37)
(Schechter 1976) where empirically α is typically small and
negative. We adopt α = −0.07, and L∗ = 1.8 × 1010L⊙
(Estafthiou et al . 1988). The supernova rate in a galaxy is
roughly proportional to its luminosity L. In the local uni-
verse it is approximately
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
NSN ≈ 10−2 L
L∗
y−1,
7
(38)
and the rate of capture of stars from Figure 4 and equation
(36) is
N ≈ 2 × 10−6(cid:16) L
L∗(cid:17)−s
y−1.
(39)
Together with the fundamental plane relation between Υ
and L, Figure 4 gives s ≈ 0.61 × 1.2 = 0.73. The ratio of the
star capture rate to the supernova rate is
f∗ ≈ Z N (L)
Z NSN(L)
dN
dL
dN
dL
dL
,
dL
whence
f∗ ≈ 2 × 10−4 α + 1
s − α (cid:16) L∗
Lmin(cid:17)s−α
(40)
(41)
where Lmin is the smallest luminosity for which we think the
majority of galaxies have a massive black hole.
The value of Lmin is not determined by observations,
since most of the observations are limited to larger systems.
M32 (L = 3.7 × 108L⊙) is the smallest galaxy in the Nuker
sample with a confirmed black hole, but it is also an un-
usually dense system at this luminosity. Dwarf galaxies in
general are usually thought not to contain black holes. For
the sake of argument consider Lmin = 108L⊙, from which
equation 41 predicts that for every supernova there would be
∼ 0.01 tidal disruptions. If we take Lmin = 107L⊙ that num-
ber goes up to ∼ 0.1. With searches now detecting many tens
of supernovae, these numbers suggest that the same surveys
or similar ones at shorter wavelength should uncover tidal
disruptions at an interesting rate.
The rate of red giant captures
NRG is approximately
proportional to M . From Figure 4 we can read off the con-
stant of proportionality and use equation (36) to write
L∗(cid:17)t
NRG ≈ 10−5(cid:16) L
y−1.
(42)
with t ≈ 1.2. However, some of the corresponding tidal dis-
ruption events will have a very long timescale, and hence
may not be detected as flares. Suppose we can only observe
flares in cases where the return time for tidal debris tmin is
less than 10y (see Ulmer 1997 for details). Then the max-
imum radius of a red giant whose disruption would be ob-
servable is
Rvis ≈ 100R⊙(cid:18) tmin
10y(cid:19)2/3(cid:18) M
106M⊙(cid:19)−1/3
.
(43)
Since red giant disruption is dominated by stars growing
onto the loss cone , the rate of disruptions up to Rvis is
proportional to Rvis, and thus
Nvis ≈ 6 × 10−7(cid:18) tmin
10y(cid:19)2/3
2t/3
y−1.
(cid:16) L
L∗(cid:17)
Integrating over the luminosity function we obtain
f∗vis ≈ 6 × 10−5(cid:18) tmin
10y(cid:19)2/3
α + 1
2t + 3α
,
(44)
(45)
where f∗vis is the ratio of the rate of visible red giant dis-
ruptions to the supernova rate, so the red giant capture rate
8
is probably much less than the main sequence capture rate
in the local Universe. Note however that it is not possible
to rule out intermittent behaviour in the accretion of tidal
debris (e.g. Lee, Kang, & Ryu 1996) and so equation (45)
represents a lower limit to the rate of detectable red giant
disruptions.
4.1.2 Upper limits to capture rate
An upper limit to the capture rate is often given as the full-
loss cone rate in the region r < rcrit. If non-axisymmetric
processes can keep the loss-cone full the capture rate can
be approximately derived from equation (10). The rate of
capture of stars from the full loss cone would be given by
N> but integrated all the way down to r = ra (assuming
that the black hole itself imposes enough symmetry to keep
the loss cone starved at r < ra). The dominant radial scale
is thus that at which ρ/σ is largest. Since σ ∼ r−1/2 for
small r, we see that if ρ ∼ r−p then N is dominated by
the contribution from ra if p > 1/2. This is the case in our
models in all but three cases (and a more realistic model of
the bound stars inside ra would probably also have p > 1/2).
The full loss cone rates for the Nuker galaxies are of order
10−3y−1, roughly independent of M (circles in Figure 4).
Thus the capture rate would be roughly
L∗(cid:17)s
N m∗ = M/t0 ≈ 4 × 10−4(cid:16) L
M⊙y−1,
(46)
with s ≈ 0.2, which is larger than the supernova rate for L
less than a few percent of L∗. Substituting equation (46) in
equation (40), we find
f∗ ≈ 4 × 10−2 α + 1
s − α (cid:16) Lmax
L∗ (cid:17)s−α
(47)
where Lmax is the largest luminosity galaxy which can tidally
disrupt a star (as opposed to swallowing it whole). Tak-
ing the maximum black hole mass for tidal disruption to
be 108M⊙ we obtain f∗ ∼ 0.1 . This means that tidal dis-
ruptions would be about ten times less likely than super-
novae if the loss cone was full. The tidal disruptions in
this case would also be coming mainly from galaxies with
L ∼ Lmax. The red giant capture rate would be simply
fRGRmax/R⊙ ≈ 10 times the main sequence rate (with
Lmax ≈ L∗).
Equations (47) and (41) together provide a constraint
on the mass supply for the black holes. If only a small num-
ber of tidal disruptions were observed, and they came from
the faintest galaxies, then the dominant fuelling would be
via empty loss cones.
Another upper limit to the capture rate is given by as-
suming that the galaxies we see today are not in a special
state, so their capture rates must be smaller than M/t0 (the
dashed line in Figure 4). In fact this is a natural value for
the capture rate in the case that the black holes grew by
eating stars from the cluster around them. For the majority
of the Nuker galaxies M/t0 is actually larger than the full
loss cone rate. Thus there has not been enough time for the
black holes to grow by eating stars from the nuclear star
cluster. A more efficient mechanism must be responsible for
their having grown to the present masses. Merrit & Quinlan
(1997) have argued that the universal x ≈ 0.006 is a result
of the black hole eating stars until the intrinsic triaxiality of
10.0
)
"
(
a
r
1.0
0.1
10+6
10+7
10+8
M (M⊙)
10+9
10+10
Figure 5. The radius of the sphere of influence ra in arcsec as
a function of black hole mass.
the star cluster is reduced by the presence of the black hole
(Gerhard & Binney 1985). At least for the larger galaxies,
triaxiality could never supply enough stars to provide all the
mass in the black holes.
4.2 Keplerian regime
The criticism that ra is often comparable with the resolution
limit of the observations in question (Rix 1993) is no longer
valid. This has been pointed out by Kormendy (1994), and
we support his view with our carefully calculated values of ra
in the sense that the measured values do not cluster around
any particular value (Table 1 and Figure 5). On the other
hand, some of the smallest black holes have ra comparable
to or smaller than 0.1".
The black hole masses measured by Magorrian et al .
(1998) required detailed modelling of the kinematics of the
surrounding galactic nuclei. In principle one should see at
small radius the Keplerian regime in which σ2 ∼ M/r. This
would be a very clear signature of a black hole, giving di-
rectly the mass to within a factor of order unity. The Kep-
lerian regime should occur within some radius of order ra.
We find that the logarithmic slope of σ2 reaches −0.95 at a
radius rK which is generally much less than ra, with median
value ≈ 0.2ra (this is of course consistent with the fact so
much modelling had to be done to measure the black hole
masses). This situation is also seen in the centre of the Milky
Way, where the Keplerian regime is now resolved (Genzel
et al . 1996) at rK ∼ ra/10.
One might expect that the galaxies with the largest val-
ues of rK would have the most secure black hole detections,
but this appears not to be the case. The three galaxies where
we find that rK > 0.5" are NGC 4486B, NGC 7768 and M31,
and these do not appear to have significantly better black
hole masses than any of the others.
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
9
Lee, H., Kang, H., & Ryu, D. 1996, ApJ, 464, 131
Joss, P. C., Rappaport, S., and Lewis, W. 1987, ApJ, 319, 180
Kormendy, J. 1994, in The Nuclei of Normal Galaxies ed. Gen-
zel, R. & Harris, A.I., (NATO ASI Ser. C., 445) (Dordrecht:
Kluwer), p. 379.
Kormendy, J., & Richstone, D., 1995, ARAA, 33, 581
Lasota, J.-P., Abramowicz, M. A., Chen, X., Krolik, J. Narayan,
R., Yi, I. 1996, ApJ, 462, L142
Lightman, A.P., & Shapiro, S.L. 1977, ApJ, 211, 244
Magorrian, J., et al ., 1998, AJ, 115, 2285
Rauch, K.P., & Ingalls, B. 1997, preprint (astro-ph/9710288).
Rauch, K.P., & Tremaine, S. 1996, NewA, 1, 149
Rees, M.J. 1998, Nature, 333, 523
Reynolds, C.S, Di Matteo, T., Fabian, A. C., Hwang, U., &
Canizares, C. R. 1996, MNRAS, 283, L111
Reynolds, C.S, & Fabian, A.C., 1997, MNRAS, 290, 1
Rix, H.-W., 1993, in IAU Symp. 153, Galactic Bulges, ed. De-
jonghe & Habing, p. 423, Dordecht: Kluwer.
Schechter, P. 1976, ApJ, 203, 297
Ulmer, A. 1997, ApJ, submitted, (astro-ph/9706247)
Young, P.J., Shields, G.A., & Wheeler, J.C. 1977, ApJ, 212, 367
This paper has been produced using the Royal Astronomical
Society/Blackwell Science LATEX style file.
5 DISCUSSION
Note that the scaling relations of FR and CK giving N
in terms M , ρ0 and r0 apply only in very specific circum-
stances. The result of CK only applies to the case rcrit < ra.
If rcrit < ra and p ≈ 1.8 for r < ra then CK equation
(66) can be used to calculate N , replacing ρ0 → ρ(ra) and
r0 → ra, with the result roughly independent of what hap-
pens at r > ra. FR are careful to distinguish between the
cases rcrit >< ra, and in these cases they use p = 7/4 or
p = 0 accordingly. It is straightforward to generalise their
expressions using equations (19) and (18).
Rauch & Tremaine (1997) have discussed the enhance-
ment of tidal disruption rates by a process they call 'resonant
relaxation'. This effect can increase tidal disruption rates of
stars which are bound to the black hole (i.e. at r < ra) by
factors of order unity. Detailed calculations by Rauch & In-
galls (1997) also show that the effect is also quencehed by
relativistic precession for black hole masses M >∼ 108M⊙. In
the models we constructed the fuelling rate was never dom-
inated by strongly bound stars, so in common with Rauch
& Ingalls (1997), we find that resonant relaxation has little
effect on tidal disruption rates.
There have been some observations of active galaxies
which develop a broad-lined feature in their spectrum over a
period of the order of months (Storchi-Bergmann et al . 1995,
Eracleous et al . 1995). Eracleous et al . (1995) constructed a
phenomonological model of such an event which involved
an elliptical accretion disc, and suggested that such a disk
could arise from the tidal disruption of a star (see also Syer
& Clarke 1992). The size of the disc they required to fit the
observations was too large to correspond to a main sequence
disruption. It could, however, arise from the disruption of a
red giant. Our discussion in Section 4.1 concluded that red
giant disruptions with such short timescales should be very
rare, so it is perhaps not surprising that such events are not
encountered frequently.
ACKNOWLEDGMENTS
We thank Martin Rees and Achim Weiss for helpful discus-
sions and comments.
REFERENCES
Bahcall, J.N., & Wolf, R.A. 1976, ApJ, 209, 214
Binney, J., and Tremaine, S. 1987, Galactic Dynamics, Princeton:
Princeton University Press.
Bressan, A., Fagotto, F., Bertelli, G., & Chiosi, C. 1992, AA
Supp., 100, 647
Byun, Y.I., et al ., 1996, AJ, 111, 1889
Cannizzo, J.K., Lee, H.M., & Goodman, J. 1990, ApJ, 351, 38
Cohn, H., & Kulsrud, R.M., 1978, ApJ, 226, 1087
Evans, C.R., & Kochanek, C.S. 1989, ApJ, 346, L13
Fabian, A.C., Lee, J.C., Brandt, W.N., Iwasawa, K., & Reynolds,
C.S., 1998, preprint (astro-ph/9803075).
Frank, J., & Rees, M.J. 1976, MNRAS, 176, 633
Genzel, R., Hollenbach, D.J., Townes, C.H., Kroker, H., &
Tacconi-Garman, L.E. 1996, ApJ, 440, 619
Gerhard, O.E. & Binney J. 1985, MNRAS, 216, 467
Haehnelt, M.G., & Rees, M.J. 1993, MNRAS, 263, 168
Hill, J.G. 1975, Nature, 254, 295
c(cid:13) 1998 RAS, MNRAS 000, 1 -- 9
|
astro-ph/9410098 | 1 | 9410 | 1994-10-31T23:16:00 | Polarisation as a Tool for Gravitational Microlensing Surveys | [
"astro-ph"
] | Much interest has been generated recently by the ongoing MACHO, EROS and OGLE projects to identify gravitationally lensed stars from the Large Magellanic Cloud and Galactic bulge, and the positive identification of several events (Alcock et al, (1993), Aubourg et al, (1993) and Udalski et al, (1993)). The rate at which such events are found should provide considerable information about the distribution of the low mass compact objects, brown dwarfs etc. responsible for the lensing, and their relative importance as a component of the cosmological `dark matter'. We show measurement of the polarisation of starlight during these events can yield considerably more information about the lensing objects than was previously considered possible. Furthermore, the consideration of extended sources is shown to have a significant effect on the interpretation of the profiles and statistics of the events. | astro-ph | astro-ph |
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
missing; you have not inserted them
ASTRONOMY
AND
ASTROPHYSICS
25.1.2018
Polarisation as a Tool for Gravitational Microlensing
Surveys
John F.L. Simmons1, Jon P. Willis1, and Andrew M. Newsam1
Department of Physics and Astronomy, University of Glasgow
Received date; accepted date
Abstract. Much interest has been generated recently by the
ongoing MACHO, EROS and OGLE projects to identify grav-
itationally lensed stars from the Large Magellanic Cloud and
Galactic bulge, and the positive identification of several events
(Alcock et al, (1993), Aubourg et al, (1993) and Udalski et
al, (1993)). The rate at which such events are found should
provide considerable information about the distribution of the
low mass compact objects, brown dwarfs etc. responsible for
the lensing, and their relative importance as a component of
the cosmological 'dark matter'. We show measurement of the
polarisation of starlight during these events can yield consid-
erably more information about the lensing objects than was
previously considered possible. Furthermore, the consideration
of extended sources is shown to have a significant effect on the
interpretation of the profiles and statistics of the events.
More detailed modelling of the gravitational microlensing
of stars by low mass objects can yield considerable information
about the lens and the lensing geometry. In particular, the mea-
surement of the variable polarisation produced by gravitational
lensing of stars could in principle provide further information
about the mass, distance and velocity of the lensing objects.
This possibility appears to have been overlooked in both the
MACHOs and EROS programmes.
Although the idea that polarisation could be produced by
gravitational lensing was raised and investigated in relation
to supernovae (Schneider and Wagoner, 1987), it seems that
not to have been considered for stars. Polarisation produced
by lensing of stars is not large, but should be measurable for
sufficiently bright stars.
Normally the light received from stars is unpolarised, unless
the star is rotationally distorted, or has an asymmetric enve-
lope or wind. However the light that emerges from the limb
of a star can be expected to be polarised (up to 10%). This
effect, predicted by Chandrasekhar (1960) who calculated it
for an electron scattering photosphere, depends on the scatter-
ing mechanism and on the direction of the emergent radiation
from the normal to the stellar photosphere. It has been ob-
served in the Sun. However, for a star, the polarised light flux
observed at the Earth is obtained by integrating the polarised
intensity over the stellar disc. The polarisation of the light at
the north/south points is in an opposite direction to that at
Send offprint requests to: J. Simmons
Fig. 1. Two stages during an event. In (a), the event is about to
begin. There is no noticeable difference in the amplification of the
N/S and E/W regions so the star remains unpolarised (although a
flux increase might already have been seen). In (b), however, the
lens is at closest approach and the significant amplification of South
w.r.t. East and West gives a net polarisation in the direction at S.
the east/west points (see figure 1). If the star is rotationally
symmetric, no net polarisation should be observed owing to
cancellation. In the case of a gravitationally lensed star, the
amplification at different points on the stellar disc will be dif-
ferent, so the net polarisation differs from zero by an amount
depending on the distance of the lensing object projected onto
the source (star's) plane. The direction of this polarisation will
also vary with as angle of the line of centres of source and
lensing object changes.
One can obtain an estimate of the degree of polarisation
for a Schwarzschild lens as follows. Distances are most conve-
niently written in units of the Einstein radius projected onto
the source plane, η0, given by
η0
2 = (cid:16) 1
aL
−
1
aS(cid:17) 2RSaS
2
(1)
aL and aS are the distances to the lens and source re-
spectively, and RS = 2GM /c2 is the Schwarzschild radius of
the lensing object. The Amplification at distance x is given
by A(z) = 1
. For small x,
2 (cid:0)z + 1
z(cid:1) where z = (cid:0)1 + 4
x2(cid:1)1/2
2
A(x) ∼ 1/x. For polarisation greater that 1%, say, this ampli-
fication must vary by 10% over one stellar radius, R (assuming
a limb polarisation of around 10%). This immediately yields
the condition that R > 0.1d, where d is the distance of ap-
proach to the centre of the stellar disc.
Exact expressions for the polarised and unpolarised flux,
FU ,FQ and FI can readily be obtained from the form of the po-
larised and unpolarised intensity (stokes parameters) as func-
tions of the angle of emergence to the normal, θ, at the surface
of the star. If we assume these to take the form
I = i0 + i1µ
U = u0(1 − µ)
(2)
(3)
where µ = cos θ, we obtain on integrating over the stellar disc
the expressions
behaviour of the time profile of the flux for an extended source
is significantly different to that of a point source (Simmons et
al, 1994). If R ∼ d0 then the amplification is smaller in the
wings but higher at closest approach, giving a sharper profile.
This is possibly the explanation of the outlier at maximum am-
plification observed by the MACHOs group for the first event
reported (Alcock et al, 1993). On the other hand if R > 3 the
amplification is almost unobservable regardless of the impact
parameter. Since the Einstein radius, η0 depends on M 1/2, this
would seriously impede the detection of low mass lenses.
Cases (ii) and (iii) will only occur for extended sources.
The polarisation time profile in (iii) will be quite specific, with
two equal maxima, and a central minimum. The relative fre-
quency of the different types of events will indicate the spatial
distribution of lensing objects.
0
2 Z 2π
2 Z 2π
2 Z 2π
0 Z 1
0 Z 1
0 Z 1
0
0
R2
as
R2
as
R2
as
FI =
FU =
FQ =
where
(i0 + i1µ) A(x(µ, φ))µ dµdφ
(4)
u0(1 − µ) A(x(µ, φ))µ cos 2φ dµdφ
(5)
u0(1 − µ) A(x(µ, φ))µ sin 2φ dµdφ
(6)
x2 = d2 + R2(1 − µ2) − 2(1 − µ2)1/2Rd cos φ
(7)
as is the distance to the source and φ the position angle.
The observed degree of polarisation is given by
p =
(F 2
U + F 2
Q)
1
2
FI
(8)
and is a function of R and d. (In fact FQ = 0 in the coor-
dinate system chosen). The latter is easily expressed in terms
of the transit velocity of the lensing object and the distance of
closest approach or impact parameter, d0 and time (all in the
source plane).
Thus simultaneous measurement of the time variable polar-
isation and unpolarised flux yields a lot of information about
the lensing set-up. If one assumes the radius of the star is
known, then one immediately obtains the value of η0. Mod-
elling of the stellar atmosphere should not even be necessary,
as the parameters i0, i1 and u0 could be obtained from the
time profile of p, FI and position angle. If one assumes addi-
tionally, though less reasonably,a transit velocity for the lens-
ing object one can obtain the distance to the lens, and indeed
its Schwarzchild radius (mass). (Of course one does need to
assume a distance to the source).
The statistics of the events are potentially more important.
We may consider three types of event
(i) flux variation
(ii) polarisation variation
(iii) transit events (i.e. where some part of the source is directly
in line with the lens centre and observer) which should
appear as sudden increases in flux.
An event (i) will be recorded when the flux variation is
greater than some chosen value. In the case of a point source,
this will simply be when the impact parameter is less than
some constant times the Einstein radius, η0, i.e. d0 < αη0. The
Fig. 2. Contours of 0.1% polarisation and a flux amplification of
1.34 (corresponding to amplification of a point source at the Ein-
stein radius) as a function of source radius and impact parameter.
For values of R and d0 inside the polarisation "ellipse" variable po-
larisation will be seen. Similarly, points in the bottom left corner
enclosed by the flux contour are flux events. The points with R > d0
are transit events.
The domain of values in the R and d plane is given in
Figure 2 (see also Simmons et al, 1994), the unit of distance
being η0. Thus for lenses of the same mass near the galactic
centre R and d take small values, and for lenses near the LMC
R and d are large. If the stars in the source plane are taken to
be uniformly and randomly distributed, the fraction of events
of any one type for a fixed radius of star is simply proportional
to the length of the interval in d. For example, for R = 1,
more than half of flux events also show variable polarisation
(see figure).
From the rate of events, and the relative frequency of types
(i), (ii) and (iii) it should be possible, given a sufficiently large
sample, determine the spatial distribution of lenses.
Two further points should be stressed. The rise in polari-
sation takes place later than the rise in total flux, by approxi-
mately a factor of 2. Thus an event suspected because of an ob-
served amplification in flux, could be monitored and confirmed
polarimetrically. Secondly, the polarisation effects arise from
the limb dependency of polarisation in an extended source.
3
Even for an unpolarised source the parameters occurring in
expression (2) for the intensity should depend on wavelength
both in continuum and lines, so one might expect that for an
extended source where limb darkening is present the relative
amplitude of the flux variation would depend on frequency. For
a point source, or for a star with no limb darkening, this effect
would not be seen, and the profile would be achromatic.
Polarisation measurements are certainly feasible. Interstel-
lar polarisation might make interpretation more difficult, but
as we are dealing here with variable polarisation, this could be
overcome. This should also provide valuable information about
the lensing, and the importance of MACHOs as a dark matter
component.
1. Acknowledgements
JPW performed this work as part of a Cormack Research
Scholarship. AMN was funded by a PPARC Postgraduate Stu-
dentship.
References
Alcock, C et al, Nature, 365, 621, (1993)
Aubourg, E. et al, Nature, 365, 623, (1993)
Chandrasekhar, S., Radiative Transfer, Dover, (1960)
Schneider, P. and Wagoner, R., ApJ, 314, 154, (1987)
Simmons, J., Newsam, A., Willis, J., In preparation
Udalski, A. et al Ann. N.Y. Acad. Sci. 688, 626, (1993)
This article was processed by the author using Springer-Verlag LaTEX
A&A style file 1990.
E
S
N
R
Source
W
d 0
Direction of Polarisation
N
d
Source
W
R
E
S
d 0
a)
Lens
Lens
b)
|
astro-ph/0511830 | 1 | 0511 | 2005-11-30T15:31:09 | Weighing the young stellar discs around Sgr A* | [
"astro-ph"
] | It is believed that young massive stars orbiting Sgr A* in two stellar discs on scales of 0.1-0.2 parsecs were formed either farther out in the Galaxy and then quickly migrated inward, or in situ in a massive self-gravitating disc. Comparing N-body evolution of stellar orbits with observational constraints, we set upper limits on the masses of the two stellar systems. These masses turn out to be few times lower than the expected total stellar mass estimated from the observed young high-mass stellar population and the standard galactic IMF. If these stars were formed in situ, in a massive self-gravitating disc, our results suggest that the formation of low-mass stars was suppressed by a factor of at least a few, requiring a top-heavy initial mass function (IMF) for stars formed near sgr A*. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 24 August 2018
(MN LATEX style file v2.2)
Weighing the young stellar discs around Sgr A∗
Sergei Nayakshin1,2, Walter Dehnen1, Jorge Cuadra2 & Reinhard Genzel3,4
1Department of Physics & Astronomy, University of Leicester, Leicester, LE1 7RH, UK
2Max-Planck-Institut fur Astrophysik, Karl-Schwarzschild-Strasse 1, 85741 Garching bei Munchen, Germany
3Max-Planck-Institut fur Extraterrestrische Physik (MPE), Garching bei Munchen, Germany
4Department of Physics, University of California, Berkeley, CA 94720, USA
24 August 2018
ABSTRACT
It is believed that young massive stars orbiting Sgr A∗ in two stellar discs on scales
of ∼ 0.1 − 0.2 parsecs were formed either farther out in the Galaxy and then quickly
migrated inward, or in situ in a massive self-gravitating disc. Comparing N-body
evolution of stellar orbits with observational constraints, we set upper limits on the
masses of the two stellar systems. These masses turn out to be few times lower than
the expected total stellar mass estimated from the observed young high-mass stellar
population and the standard galactic IMF. If these stars were formed in situ, in a
massive self-gravitating disc, our results suggest that the formation of low-mass stars
was suppressed by a factor of at least a few, requiring a top-heavy initial mass function
(IMF) for stars formed near Sgr A∗.
Key words: Galaxy: centre -- accretion: accretion discs -- galaxies: active -- stars:
formation
1
INTRODUCTION
Sgr A∗ is a MBH ∼ 3.5 × 106 M⊙ super-massive black hole
(SMBH) in the centre of our Galaxy (e.g., Schodel et al.,
2002; Ghez et al., 2003). Few tens of close young massive
stars dominate the energy output of the central parsec of
the Galaxy (Krabbe et al., 1995; Genzel et al., 2003). The
ages of the young stars are estimated at t = 6 ± 1 mil-
lion years (Paumard et al., 2005). The origin of these stars
is an important mystery for astrophysics of Active Galac-
tic Nuclei (AGN). "Normal" modes of star formation at 0.1
pc distances from a SMBH are forbidden due to the huge
tidal field of the central object. Star formation in a mas-
sive gravitationally unstable accretion disc (Paczy´nski, 1978;
Kolykhalov & Sunyaev, 1980; Collin & Zahn, 1999; Good-
man, 2003) has been suggested (Levin & Beloborodov, 2003;
Milosavljevi´c & Loeb, 2004; Nayakshin & Cuadra, 2005).
Alternatively, the observed close young stars could be rem-
nants of a massive young star cluster whose orbit decayed
due to dynamical friction (e.g., Gerhard, 2001; Kim & Mor-
ris, 2003; Kim et al., 2004; McMillan & Portegies Zwart,
2003; Gurkan & Rasio, 2005).
A successful model for the origin of young stars in
Sgr A∗ will have to explain quantitatively not only the cre-
ation of the stars but also their present day orbits. Almost
all of the observed young stars in Sgr A∗ belong to one of
two stellar rings (Levin & Beloborodov, 2003; Genzel et al.,
2003). Internal N-body disc evolution sets an upper limit on
the total mass of each of the stellar systems of the order of
∼ 3 × 105 M⊙, too high to be constraining for either of
M <
the models (Nayakshin & Cuadra, 2005). In addition to the
internal disc thickening, discs precess in their mutual (non
axi-symmetric) potential and are warped with time (Nayak-
shin, 2005). Stellar discs that end up too strongly warped or
thick will contradict the observations. Nayakshin & Cuadra
(2005) suggested that this sets an upper limit on the total
stellar mass in the range of M <
noted that more detailed tests are needed.
∼ (3 − 10) × 104 M⊙, but
The goal of this paper is to perform such tests numer-
ically. Let us estimate the magnitude of the warping effect.
Nayakshin (2005) calculated the rate at which a massless
accretion disc is warped by a massive ring inclined with re-
spect to the (initially flat) disc at an angle β. For radii R
much smaller or much larger than the ring radius, Rring, the
precession angular frequency ωp(R) can be approximated as
ωp(R)
ΩK (R) ≈ −
3Mring
4MBH
cos β
R3R2
(cid:2)R2 + R2
ring
ring(cid:3)5/2 .
(1)
Here Mring is the ring mass, assumed to be much smaller
than the blackhole mass, MBH, and ΩK (R) is the Kepler cir-
cular frequency for the disc at radius R. The period of circu-
lar motion 3" (1′′ ≈ 0.04 pc at the distance of Sgr A∗) away
from Sgr A∗, where most of the bright young stars are found,
is around 2 thousand years. Therefore, the stars at this lo-
cation make about a thousand revolutions in a few million
years. Let ∆γ = ωpt be the angle on which an annulus of the
disc will precess during this time: ∆γ ∼ (Mring/MBH)×1000.
2
S. Nayakshin, W. Dehnen, J. Cuadra & R. Genzel
Very approximately, the disc warping will be noticeable
when ∆γ ∼ 1. Thus mass ratios (Mring/MBH) in excess of
10−3 or so may lead to warping of the stellar discs in Sgr A∗.
Equation 1 depends on the angle β, and would also
depend on the eccentricity of a stellar orbit if it were not
assumed to be zero in the derivation. Stars are expected to
be born on nearly circular Keplerian orbits (e.g., Nayakshin
& Cuadra, 2005), i.e. with eccentricity e ∼ H/R ≪ 1, for
star formation in a self-gravitating disc. On the other hand,
Levin et al. (2005) have recently demonstrated that, due to
a repetitive interaction with the IMBH, stars peeled off the
IMBH-star cluster gain a substantial eccentricity to their
orbits. Even for the case of the circular IMBH inspiral, stars
acquire a mean eccentricity of e ≈ 0.5. We may thus expect
different mass limits for the two models of the origin of the
young stars near Sgr A∗.
Reader not interested in the technical details of the sim-
ulations and data comparison may simply look up the rel-
evant limits and proceed to the discussion of the results in
§5.
2 NUMERICAL METHOD
To follow stellar orbits, we use the N-body package
'NEMO' (Teuben, 1995) with the orbit integrator 'gyrfal-
cON' (Dehnen, 2002). The code calculates gravitational in-
teraction of all the particles and updates their velocities and
positions. We model the blackhole as a Plummer sphere with
a core radius of 0.01". This radius is much smaller than
peri-centres of stellar orbits we consider. The stars are rep-
resented as particles with softening radius of 0.01′′ or less
for some tests. Depending on the total mass of the stars, we
use between few hundred to few thousand particles for each
of the two stellar systems.
To set up the problem, we shall rely on the gross re-
sults of Genzel et al. (2003), and the more recent analysis of
the data by Paumard et al. (2005). For convenience only, we
shall refer to the clock-wise rotating system in the GC as a
disc, and the counter clock-wise system as a ring. To model
the disc, we start with stars in a flat disc with the radial
extent from Rin = 2′′ to Rout = 5′′. The observed counter
clock-wise system contains fewer stars and it is harder to as-
sign the radial extent for this system (Paumard et al., 2005).
We used two plausible guesses for the ring, therefore: one is
a ring between Rin = 4′′ and Rout = 5′′, and the other is
a ring between Rin = 5′′ and Rout = 7′′. The initial incli-
nation between the two systems is set at β = 113◦ (Genzel
et al., 2003; Paumard et al., 2005). Other parameters of ini-
tial conditions specific to a model will be discussed below.
The simulations were ran until time t ≈ 3 Million years.
To ensure numerical precision, the individual timesteps of
the particles were kept small enough, e.g. in the range of
0.06 to 2 years. The angular momentum and total energy
of the system were conserved to a relative error of 10−3
or better. We did not include the isotropic cluster of late
type stars around Sgr A∗ in these simulations because in
the radial range of interest its mass is small compared to
that of Sgr A∗, and it would not present any torques for the
discs because of the cluster's spherical symmetry.
At the end of a simulation, we fit the two stellar systems
by a plane in velocity space , (vx, vy, vz) . A plane is charac-
terised by vector ~n orthogonal to the plane and normalised
to unity, ~n2 = 1. The best fitting planes are found via the
reduced χ2 method for both of the two stellar systems. The
reduced χ2 for a fit is defined as in Levin & Beloborodov
(2003):
χ2 =
1
Ns − 1
Ns
X
i=1
(vxnx + vyny + vznz)2
(σxnx)2 + (σyny)2 + (σznz)2 ,
(2)
where Ns is the number of stars in the given stellar system,
nx, ny and nz are the projections of ~n, and σx, σy and σz
are the errors in stellar velocities in the three directions. For
simplicity we assume that these errors are isotropic, i.e. each
component is equal to σ/√3 ≈ 40 km/sec, where σ = 70
km/sec, the typical value for the absolute value of velocity
error vector in Paumard et al. (2005).
Finally, to emulate the effects that the observational
errors have on the χ2 fits through statistical scatter of the
data around the mean, we added random Gaussian velocity
kicks to each star's velocity vectors according to the errors
σx, etc, at the end of the simulations.
3 EXAMPLE RUNS FOR CIRCULAR INITIAL
STELLAR ORBITS
As explained in §2, we start with the two stellar systems
oriented as the observed best fitting planes (Paumard et al.,
2005). The discs are populated by star particles with sur-
face density Σ(R) ∝ R−2 for Rin ≤ R ≤ Rout. The initial
velocity dispersion of stars with respect to the local circular
Kepler velocity is isotropic and small, h~v − ~vKi = 0.017vK ,
consistent with the small initial gaseous disc thickness.
A convenient way to illustrate the results is via the
'Aitoff map projection' of the stellar orbits. Each stellar or-
bital plane can be characterised by the inclination angle i
with respect to the observer and the position angle φ which
reflects the position of the lines of the nodes of the orbit on
the plane of the sky (e.g. Schodel et al., 2002; Ghez et al.,
2003; Eisenhauer et al., 2005; Ghez et al., 2005). The normal
vectors to the plane, introduced above, are related to these
angles as nz = cos i, nx = sin i cos φ, and ny = sin i sin φ.
The Aitoff's projection then shows the latitude, defined as
π/2 − i, and longitude = φ.
Figure 1 shows the state of the two stellar systems 3
million years into one of the simulations, e.g. the final state.
The total stellar masses in the disc and the ring systems
(CW and CC systems, respectively) used in this simulation
are noted in the left upper corner of the Figure. The stars
in the clock-wise system (marked as CW on the Figure) are
shown with green asterisks, whereas the counter-clockwise
system is shown with red crosses, marked as CC. The lower
left corner shows the resulting values of the minimum re-
duced χ2 values as a (ring, disc) pair. These are comfortably
smaller than the observed values χ2
cc = 4.0
for the clock-wise and the counter clock-wise stellar systems,
respectively (Paumard et al., 2005). Apparently such a weak
disc warping would not contradict observations.
cw = 2.5 and χ2
Figure (2), shows the same numerical experiment but
with the disc and ring masses both around 104 M⊙. As is
clear from the Figure, both systems suffer considerably from
the gravitational torques imposed on each other, and the
3
Figure 1. The orientation angles of the stellar orbital planes, as
described in the text. The positions of the clockwise and counter-
clockwise systems are shown with symbols CW and CC, respec-
tively. Total disc and ring masses are labelled at the upper left
corner. The values of the minimum reduced χ2 fits to the (ring,
disc) systems are displayed in the lower left corner of the Figure.
Figure 3. Same as Figures 1,2 but for the largest values of the
disc and ring masses considered. The reduced χ2 values for both
the ring and the disc are much larger than the observed values.
Such a stellar distribution cannot be classified as consisting of
two discs at all.
Figure 2. Same as Figure 1 but for much higher disc and ring
masses. Note that this stellar distribution is somewhat inconsis-
tent with the data, with the reduced χ2 values exceeding the
observed values of χ2
cc,cw = (4.0, 2.5).
Figure 4. Same test as shown in Figure 3, but with stars di-
vided into populations based on whether they rotate clock-wise or
counter clock-wise. Note that the resulting χ2 values are smaller
but still much larger than the observed values.
resulting plane-parallel fits to these are poorer than before.
The reduced χ2 values are larger than the observed values
for both the clock-wise and the counter clock-wise systems.
Figure 3 presents the Aitoff map for the most massive
case that we have explored, with the disc and ring masses
both around 3.5× 104 M⊙. The resulting χ2 values are quite
large and are clearly inconsistent with the data. Note that
the degree of mixing occurring in this simulation is extreme,
and some of the particles that originally belonged to the
clockwise disc are now classified as members of the counter-
clockwise system. A stellar system with such poor values of
χ2 may not even remotely be considered as consisting of two
or one stellar discs, obviously.
In this latest example we have seen that stars from one
system may evolve on orbits more consistent with the other
system, and it would then be improper to continue to assign
them to the disc of their birth place. Certainly observation-
ally such practice is impossible as it is not a priory known in
which system the stars originated. Instead, one divides the
stars on the clock-wise and the counter clock-wise ones (Gen-
zel et al., 2003). To make a maximally fair comparison of the
simulations with the data, at the end of the simulation, we
assign particles to either the clock-wise or the counter clock-
wise systems based on the sign of Jz = (xvy − yvx), where
x, y are the coordinates of the star. The resulting change is
illustrated in Figure 4. The Figure demonstrates that when
the stellar systems are warped to the degree that their orbits
diffuse into one another's phase space, even a more careful
division of orbits still fails to produce systems as well defined
as those observed to exist near Sgr A∗.
4
S. Nayakshin, W. Dehnen, J. Cuadra & R. Genzel
Table 1. Simulations results for circular initial orbits with the
larger "ring", with R = 5" − 7". "fit quality" below, + or − sign,
indicates whether the reduced χ2 values are smaller or larger than
the observed values. Clearly this is just a rough measure of the
model's feasibility.
Mdisc Mring
run
χ2
disc
χ2
ring
fit qualitya
3500
6300
12000
35000
12000
12000
38000
38000
1750
1750
1750
1750
3500
10500
10500
35000
CL1
CL2
CL3
CL4
CL5
CL6
CL7
CL8
0.9
1.6
1.4
3.2
1.3
6.0
6.2
9.5
1.9
1.0
5.6
12.7
3.7
6.3
11.5
13.0
+
+
-
-
+
-
-
-
Table 2. Simulations results for circular initial orbits with the
smaller ring, R = 4" − 5".
Mdisc Mring
run
χ2
disc
χ2
ring
fit quality
3500
6300
12000
35000
12000
12000
38000
38000
1750
1750
1750
1750
3500
10500
10500
35000
CS1
CS2
CS3
CS4
CS5
CS6
CS7
CS8
0.9
1.2
1.
3.6
1.2
4.2
8.7
10.0
1.5
1.0
5.7
17.6
4.6
7.0
16.0
10.8
+
+
-
-
-
-
-
-
4 RESULTS
4.1 Circular initial orbits
Table 1 lists results of several runs with the larger stellar ring
(i.e., R = 5"−7"). The first two column show the total stellar
mass of the clock-wise (the disc) and the counter clock-wise
(the ring) systems in Solar masses. The next column shows
an identifier of the run (CL stands for "circular large"). Next
two columns list the best fitting χ2 values. The last column
contains a "+" if the χ2 values are smaller than the observed
ones, and a "-" in the opposite case. Clearly this is not to
be taken literally as in some cases the obtained χ2 are just
slightly larger than the observed ones.
In a similar fashion, the results of the tests with a
smaller stellar ring (the counter clock-wise system, R =
4" − 5") are presented in Table 2. As one can see, the two
tables are in general similar. The combined results of the
small and larger scale ring tests are best described as these
limits:
Mdisc
Mring
<
∼ 1.5 × 104 M⊙
∼ 1. × 104 M⊙ .
<
(3)
(4)
4.2 Infalling star cluster: eccentric initial orbits
Levin et al. (2005) have recently modelled the dynamics of
the stars peeled off from the IMBH inspiralling to smaller
radii in the field of Sgr A∗. The orbits of these stars were
found to gain significant eccentricities due to repetitive in-
Figure 5. The initial distribution of eccentricities and inclination
angles for a test with 500 particles in the clock-wise (disc) system.
The distribution is chosen to mimic that resulting from a circular
inspiral of an IMBH-star cluster (see Fig. 1 in Levin et al., 2005).
teractions with the IMBH while they have not yet distanced
themselves very far from it. The effect is the strongest if
the IMBH is itself on an eccentric orbit, when some of the
stars are flung to orbits with an angular momentum op-
posing their initial one. Some of the stars in fact become
unbound and escape to infinity. The eccentric star cluster
inspiral thus defies the purpose of the model -- to bring
massive young stars in -- as many are flung out on orbits
in which they spend most of the time outside the central
parsec. This would contradict observations (Genzel et al.,
2003). We thus consider only the circular cluster inspiral
model here. The stars then obtain(Levin et al., 2005) rela-
tively mild eccentricities, with a median of around 0.5, and
the resulting stellar disc is rather thin since the stellar ve-
locities in the direction perpendicular to the IMBH inspiral
orbit are quite small.
To generate initial orbits similar to those obtained by
Levin et al. (2005), we place the stars radially in the same
way as before, but we now add to their velocities random
components both in the plane of the disc and perpendicular
to the disc. In the plane of the disc, we added a random ve-
locity component ( in both radial and azimuthal directions),
∆v, distributed between −(1/5)vc to +(1/5)vc, where vc is
the local circular speed. A smaller random component in the
direction perpendicular to the discs was also given to imitate
distribution of orbital inclinations to the system's plane. The
resulting distribution of stellar initial eccentricities and in-
clinations (here defined with respect to their respective disc
planes), shown in Figure 5, is indeed similar to Figure 1 of
Levin et al. (2005).
The results of the tests with the given eccentricity and
inclination distribution (Figure 5) are summarised in Table
3. The resulting χ2 values are significantly larger than those
Table 3. Simulations results for eccentric initial orbits with larger
ring.
Mdisc Mring
run
χ2
disc
χ2
ring
fit quality
3500
6300
12000
3500
6300
12000
35000
3500
3500
12000
6650
3500
1750
1750
1750
3500
3500
3500
3500
10500
14000
10500
10500
7000
EL1
EL2
EL3
EL4
EL5
EL6
EL7
EL8
EL9
EL10
EL11
EL12
1.6
1.5
2.9
2.2
1.9
2.7
15.6
7.9
17.5
8.5
5.1
3.0
3.8
9.5
6.9
1.7
7.2
13.0
3.4
2.8
2.7
12.2
5.5
2.1
+
+
-
+
-
-
-
-
-
-
-
-
for the initially circular orbits at same masses. We estimate
that the disc masses must be less than
Mdisc
<
Mring
<
∼ 5 × 103 M⊙
∼ 5 × 103 M⊙ .
(5)
(6)
We interpret these tighter upper limits as a result of a wider
spectrum of orbits in the eccentric discs. The greater diver-
sity in the orbits leads to a greater difference in the rates at
which the orbital planes precess.
4.3 The maximum current mass of the stellar
systems
Using the same ideas described above, we can ask a slightly
different question. We can start with orbits spread around in
the velocity space sufficiently wide to yield χ2 values consis-
tent with the observed discs, follow these orbits for a shorter
time, say a million year, and then measure the χ2 again. If
χ2 significantly increases during this "short" time, such a
disc should be rejected. The argument here is that during
a time considerably shorter than the age of the discs, the
observed system should not evolve (get warped and mixed)
too much. We have ran a grid of models corresponding to
disc and ring radii from the "larger" ring tests. The approx-
imate upper limits on the masses of the system obtained in
this way are
Mdisc
<
Mring
<
∼ 1 × 104 M⊙
∼ 5 × 103 M⊙ .
(7)
(8)
5 DISCUSSION AND CONCLUSIONS
In this paper we modelled N-body evolution of two stellar
discs guessing their initial geometrical arrangement based on
their present day observed configuration, and then following
stellar orbits for 3 Million years. Given the uncertainty in
the initial guess, these tests cannot be precise, but since the
discs are about twice older than 3 Million years (Paumard
et al., 2005), we feel our upper mass limits are conserva-
tive. In summary, we found that the total mass of each of
the stellar systems orbiting Sgr A∗ cannot be greater than
∼ (5 − 15) × 103 M⊙, with the lower values for eccentric
M <
5
initial stellar orbits, and the higher limits for the circular
orbits. It is somewhat disappointing to us that these lim-
its are consistent with both of the models for star forma-
tion near Sgr A∗. The minimum gaseous mass of the disc at
which it becomes gravitationally unstable and forms stars
is around Mdisc ≃ 104 M⊙ for Sgr A∗ case (Nayakshin &
Cuadra, 2005). Note that this was derived from the basic
Shakura & Sunyaev (1973) model, without including self-
gravity into the hydrostatic balance equation for the disc.
We expect the minimum Mdisc to be another factor of ∼ 2
lower, therefore. The mass of the stars formed through the
disc collapse should be close to the original gas mass, as ar-
gued by Nayakshin & Cuadra (2005), because the viscous
time scale for gas accretion is much longer than the time
scale on which the stars can devour the disc. Thus the min-
imum stellar mass in that model is of order 5 × 103 M⊙, in
line with the upper limits obtained here.
Now, for the cluster infall model, the initial mass of the
cluster should be as large as 105−106 M⊙ to allow it to spiral
in rapidly enough, as well as to form an intermediate mass
black hole that is heavy enough to prevent cluster dissolution
away from the central parsec (Kim et al., 2004; Gurkan &
Rasio, 2005). However the final stellar masses that make it
into the central parsec are always a very small fraction of
the original cluster mass, and they appear to be consistent
with the limits obtained here.
We may also approach these mass limits from another
angle. Estimating the total mass of the observed young mas-
sive stars, and assuming a standard Salpeter (1955) IMF,
one would predict the total stellar mass to be around ∼ (few
to 10) ×104 M⊙, depending on the low mass cutoff in the
power-law. This would be a factor of several larger than the
mass limits for the circular initial orbits. Thus, if the stars
were formed inside a massive self-gravitating disc, their IMF
should be at least somewhat deficient in the low mass stars.
In fact, in a separate paper, Nayakshin & Sunyaev (2005)
show that a similar but even stronger argument can be made
on purely observational grounds.
We thank an anonymous referee for his useful com-
ments. This research was supported in part by the National
Science Foundation under Grant No. PHY99-07949 during
SN's visit to the Kavli Institute for Theoretical Physics in
Santa Barbara.
REFERENCES
Collin S., Zahn J., 1999, A&A, 344, 433
Dehnen W., 2002, Journal of Computational Physics, 179, 27
Eisenhauer F., Genzel R., Alexander T., et al., 2005, ApJ, 628,
246
Genzel R., Schodel R., Ott T., et al., 2003, ApJ, 594, 812
Gerhard O., 2001, ApJ, 546, L39
Ghez A. M., Duchene G., Matthews K., et al., 2003, ApJ, 586,
L127
Ghez A. M., Salim S., Hornstein S. D., et al., 2005, ApJ, 620, 744
Goodman J., 2003, MNRAS, 339, 937
Gurkan M. A., Rasio F. A., 2005, ApJ, 628, 236
Kim S. S., Figer D. F., Morris M., 2004, ApJ, 607, L123
Kim S. S., Morris M., 2003, ApJ, 597, 312
Kolykhalov P. I., Sunyaev R. A., 1980, Soviet Astron. Lett., 6,
357
Krabbe A., Genzel R., Eckart A., et al., 1995, ApJ, 447, L95
Levin Y., Beloborodov A. M., 2003, ApJ, 590, L33
6
S. Nayakshin, W. Dehnen, J. Cuadra & R. Genzel
Levin Y., Wu A. S. P., Thommes E. W., 2005, ArXiv Astrophysics
e-prints
McMillan S. L. W., Portegies Zwart S. F., 2003, ApJ, 596, 314
Milosavljevi´c M., Loeb A., 2004, ApJL, 604, L45
Nayakshin S., 2005, MNRAS, 359, 545
Nayakshin S., Cuadra J., 2005, A&A, 437, 437
Nayakshin S., Sunyaev R., 2005, MNRAS, 364, L23
Paczy´nski B., 1978, Acta Astron., 28, 91
Paumard T., Genzel R., Eisenhauer F., Ott T., Trippe S., 2005,
submitted to ApJ
Salpeter E. E., 1955, ApJ, 121, 161
Schodel R., Ott T., Genzel R., et al., 2002, Nature, 419, 694
Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337
Teuben P., 1995, in ASP Conf. Ser. 77: Astronomical Data Anal-
ysis Software and Systems IV, 398 -- +
|
astro-ph/0310734 | 1 | 0310 | 2003-10-26T09:56:09 | A library of galaxy mergers | [
"astro-ph"
] | We have undertaken a large series of numerical simulations with the goal to built a library of galaxy mergers (GALMER). Since the aim is to have more than a thousand of realisations, each individual run is simplified, with a small number of particules (24 000), but following the chemodynamical evolution, with star formation and feedback. We illustrate in this short report some preliminary results. | astro-ph | astro-ph | SF2A 2003
F. Combes, D. Barret, T. Contini and L. Pagani (eds)
3
0
0
2
t
c
O
6
2
1
v
4
3
7
0
1
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
A LIBRARY OF GALAXY MERGERS
Combes, F., Melchior A.-L. 1
Abstract. We have undertaken a large series of numerical simulations
with the goal to built a library of galaxy mergers (GALMER). Since the
aim is to have more than a thousand of realisations, each individual run
is simplified, with a small number of particules (24 000), but following the
chemodynamical evolution, with star formation and feedback. We illustrate
in this short report some preliminary results.
1 Models and techniques
We use a TREE-SPH code to follow the self-gravity of all components, and the
dissipative nature of the gas (cf Combes & Melchior 2002). The 24 000 particles
are distributed among stars, gas and dark matter halo, with varying numbers,
according to the initial morphological types of the galaxies. For this fist series of
128 runs, we have 4 types of galaxies (i.e. 16 types of initial couples): E0, Sa, Sbc
and Sd, two masses (giants at 2 1011 M⊙ and dwarfs at 5 1010 M⊙), two different
vectors for relative initial velocities, and two opposite senses on these orbits (direct
or retrograde). The relative inclination of the galaxy planes is fixed to 45◦. The
star formation rate for the various runs is time-averaged in Figure 1.
2 Star formation recipe
We have also varied the recipe for the star formation, and tested the corresponding
efficiency. It is basically a "local" Schmidt law, where a fraction of any gas particle
transformed to stars is ∝ ρn−1, where ρ is the volumic gas density, and n the
power of the Schmidt law. An algorithm with hybrid particules is used (cf Mihos &
Hernquist 1994): when their gas fraction drops below 5%, they are then turned into
pure stars, the gas being spread among the neighbours. The energy of supernovae
and stellar winds is partly reinjected in the ISM under the form of kinetic energy,
through expanding velocity of the surrounding gas. To simulate the increase of
star-formation efficiency in violent interactions, we also added a term proportional
1 LERMA, Observatoire de Paris, 61 Av. de l'Observatoire, F-75014, Paris, France
c(cid:13) EDP Sciences 2003
238
SF2A 2003
Fig. 1. Average star formation rate (in units of percentage of initial gas mass consumed
per Myr) over the whole merging simulation (1.2 Gyr) for various runs as a function of
morphological type of the more massive galaxy of the pair. The type of the second galaxy
is indicated by different symbols. The direct orbits lead in average to larger SFR.
to the local gas velocity divergence, to a power β (this divergence is counted only
when divv negative, like in the viscosity term). The comparison between the rate
of star formation in the merger, and that in isolated galaxies, depends strongly
on the adopted SF rate. When all gas is consumed before the first pericenter, the
boost due to the dynamical triggering is very limited. The influence of the divv
term is less that the absolute rate of star formation (see figure 2).
Fig. 2. Star formation ratio between the merger run and the corresponding control
run with the two galaxies isolated. Left, the SF recipe is the Schmidt law, with n=1.5;
Right, the SF recipe is the same n=1.5 law, but with now a term proportionnal to the
velocity divergence, to simulate the action of shocks and ISM agitation.
References
Combes F., Melchior A-J.: 2002, Ap&SS 281, 383
Mihos, J. C., Hernquist, L.: 1994, Apj 437, 611
|
0804.4185 | 1 | 0804 | 2008-04-25T22:39:59 | Infall caustics in dark matter halos? | [
"astro-ph"
] | We show that most particle and subhalo orbits in simulated cosmological cold dark matter halos are surprisingly regular and periodic: The phase space structure of the outer halo regions shares some of the properties of the classical self-similar secondary infall model. Some of the outer branches are clearly visible in the radial velocity - radius plane at certain epochs. However, they are severely broadened in realistic, triaxial halos with non-radial, clumpy, mass accretion. This prevents the formation of high density caustics: Even in the best cases there are only broad, very small (<10 percent) enhancements in the spherical density profile. Larger fluctuations in rho(r) caused by massive satellites are common. Infall caustics are therefore too weak to affect lensing or dark matter annihilation experiments. Their detection is extremely challenging, as it requires a large number of accurate tracer positions and radial velocities in the outer halo. The stellar halo of the Milky Way is probably the only target where this could become feasible in the future. | astro-ph | astro-ph |
Draft version October 24, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
INFALL CAUSTICS IN DARK MATTER HALOS?
Jurg Diemand1,2 and Michael Kuhlen3
Draft version October 24, 2018
ABSTRACT
We show that most particle and subhalo orbits in simulated cosmological cold dark matter halos are
surprisingly regular and periodic: The phase space structure of the outer halo regions shares some of
the properties of the classical self-similar secondary infall model. Some of the outer branches are clearly
visible in the radial velocity - radius plane at certain epochs. However, they are severely broadened
in realistic, triaxial halos with non-radial, clumpy, mass accretion. This prevents the formation of
high density caustics: Even in the best cases there are only broad, very small (< 10%) enhancements
in the spherical density profile. Larger fluctuations in ρ(r) caused by massive satellites are common.
Infall caustics are therefore too weak to affect lensing or dark matter annihilation experiments. Their
detection is extremely challenging, as it requires a large number of accurate tracer positions and radial
velocities in the outer halo. The stellar halo of the Milky Way is probably the only target where this
could become feasible in the future.
Subject headings: dark matter -- galaxies: formation -- Galaxy: halo -- methods: n-body simulations
1. INTRODUCTION
radial
Idealized models of self-similar,
infall onto
an initial spherical overdensity in an expanding and
otherwise homogenous universe can be solved exactly
(Fillmore & Goldreich 1984; Bertschinger 1985) (FGB
hereafter). The two key predictions of these models are:
(i) Nearly isothermal (ρ ∝ r−2) halo density profiles, as
found in earlier models (Gott 1975; Gunn 1977). (ii) Near
their apocenters mass shells pile up and form spherical
caustic surfaces of infinite density. This occurs at con-
stant fractions of the current turnaround radius, result-
ing in slowly outward moving caustics as the turnaround
radius increases with time. Each mass shell moves on a
radial orbit with a nearly fixed (only slowly decreasing)
apocenter of about 0.85 of its own turnaround radius.
After a shell falls in it forms part of the first (outermost)
caustic when the shell is near its first apocenter passage.
Later, near its second apocenter, it forms part of the
second caustic, and so on. Such caustics would increase
the dark matter annihilation rate (Bergstrom et al. 2001;
Mohayaee & Shandarin 2006), and it has been argued
that infall would lead to peaks in the local CDM velocity
distribution and change the predictions for direct dark
matter detection (Sikivie et al. 1997).
Both key predictions of the original FGB-model are at
odds with the properties of dark matter halos found in
cosmological simulations: Halo density profiles are found
to be steeper than ρ ∝ r−2 in the outer parts and shal-
lower in the inner regions (Dubinski & Carlberg 1991;
Navarro et al. 1996). However, by relaxing some of the
simplifying assumptions of the FGB-model it is possible
to obtain a more realistic secondary infall model which
successfully reproduces the halo density profiles found
in cosmological simulations (e.g. Ascasibar et al. 2007
and references therein): allowing for angular momen-
1 University of California, High Street 1156, Santa Cruz, CA
95064; [email protected]
2 Hubble Fellow
3 Institute for Advanced Study, Einstein Drive, Princeton, NJ
0854; [email protected]
tum lowers the inner densities relative to the original
ρ ∝ r−2.25 profile, simply because fewer particles orbit
through the central regions (Ryden 1993). More realis-
tic initial perturbations lead to a steeper outer profile
(Hoffman & Shaham 1985).
Infall does not lead to high density caustic surfaces
in cosmological simulations.
In early simulations most
particle orbits differed widely from the slowly shrink-
ing, nearly periodic orbits found in the FGB-model
(Zaroubi et al. 1996). But these simulations contained
only a few thousand particles, which leads to signif-
icantly distorted orbits due to artificial two body re-
laxation (Diemand et al. 2004b). Despite the numerical
noise, Zaroubi et al. (1996) found that the final particle
energies are strongly correlated with the initial ones, sug-
gesting a rather gentle, orderly halo build-up as in FGB,
and not violent relaxation. Another consequence of this
fairly regular build-up is that material from biased, early
high density regions ends up in the inner parts of ha-
los today (Diemand et al. 2005). However, recent high
resolution simulations still produce rather smooth mass
profiles without high density caustics (e.g. Moore et al.
2001; Helmi et al. 2002; Diemand et al. 2007). Moore
(2001) emphasized that in the hierarchical buildup of
CDM halos the internal velocity dispersion of infalling
satellites σsat is typically only a few times smaller than
σhost, which would broaden caustics significantly, to a
width of about Rhostσsat / σhost. Indeed resolved merg-
ers do bring in practically all the mass and there is no
evidence for smooth accretion (Madau et al. 2008). Here
we show that despite clumpy, non-radial, anisotropic ac-
cretion into a non-spherical potential, the outer parts of
halos do show a few weak, broad branches in vr − r-
space. Their properties agree well with some of the FGB
predictions. In contrast to the model however, realistic
infall "caustics"4 increase the dark matter densities only
slightly (≃ 10% for the strongest ones). Unfortunately
4 For brevity we refer to the turnaround regions of branches in
vr − r-space simply as "caustics", even tough they don't have the
high densities implied by the true sense of the word.
2
Diemand & Kuhlen
/
]
s
m
k
[
y
t
i
c
o
e
v
l
l
i
a
d
a
r
/
]
s
m
k
[
y
t
i
c
o
e
v
l
l
i
a
d
a
r
400
300
200
100
0
-100
-200
-300
-400
400
300
200
100
0
-100
-200
-300
-400
first infall
100
200
300
400
500
600
third orbit
100
200
300
400
500
600
40
400
35
300
30
200
25
20
15
10
5
100
0
-100
-200
-300
-400
400
300
200
100
0
-100
-200
-300
-400
14
12
10
8
6
4
2
0
first orbit
100
200
300
400
500
600
fourth orbit
100
200
300
400
500
600
35
400
30
25
20
15
300
200
100
0
-100
10
-200
5
-300
-400
10
400
9
8
7
6
5
4
3
2
1
0
300
200
100
0
-100
-200
-300
-400
second orbit
100
200
300
400
500
600
fifth orbit
100
200
300
400
500
600
30
25
20
15
10
5
10
9
8
7
6
5
4
3
2
1
0
radius [kpc]
radius [kpc]
radius [kpc]
Fig. 1. -- Phase space distribution of subhalos at z=0 separated by the number of orbits they have completed. This figure is also available
as an mpeg animation running from z=12 to 0 in the electronic edition of ApJ.
this makes them practically undetectable and irrelevant
for dark matter detection.
2. WEAK OUTER CAUSTICS IN VR − R SPACE
In this section we compare the histories of subhalos in
the Via Lactea simulation (Diemand et al. 2007, VL07
hereafter)5 with predictions from the FGB-model. Later
we will include the distribution of all dark matter par-
ticles. The subhalos are selected by peak circular ve-
locity Vmax > 3 km s−1 before accretion, and are there-
fore very similar to dark matter particles in their spa-
tial and velocity distribution (Faltenbacher & Diemand
2006). This sample contains 10,652 halos and subhalos
with masses above about 3 × 106 M⊙. The host halo has
size of Vmax = 181 km s−1, consistent with the dark halo
of the Milky Way. The subhalos which have completed
the same small number of orbits end up in a relatively
narrow branch in vr − r space, despite their range of or-
bital eccentricities (VL07), the prolate host halo poten-
tial (about 0.8:1, see Kuhlen et al. 2007) and occasional
kicks from larger satellites (Sales et al. 2007). Each panel
in Figure 1 would correspond to one infinitesimally thin
branch in the vr − r plane of the FGB-model. Already at
first infall the radial and tangential velocity dispersions
are quite large: σr ≃ 26 km s−1 and σtan,2D ≃ 60 km s−1.
These motions induced by large scale tidal forces and also
by structures forming within the turnaround region pre-
vent the formation of thin, high density caustics.
From the positions of subhalos near their apocenter we
can measure the position and width of the outer six infall
caustics (Table 1). In the FGB-model the radial velocity
of a caustic at radius r is vcaustic(r) = 8/9 r/t, where t
5 Subhalo histories, animated versions of Figures 1 and 2 and
other movies are available at www.ucolick.org/∼diemand/vl
is the time since the Big Bang6. We define those sub-
halos near their kth apocenter passage and with a radial
velocity within 24 km s−1 of vcaustic(r) as the members
of caustic k. For each member i we then determine its
turnaround radius tk,i and the ratio of today's position
to turnaround radius rk,i/tk,i. Note that members of one
caustic fell in from all directions, not as a bound group,
and merely share similar turnaround times and radii.
The scatter in the turnaround radii is relatively large
and comparable to the width of the caustics. This again
hints at a broadening of caustics caused already early
(near the turnaround time) by proper motions. Caustics
4, 5, and 6 are so wide that they completely overlap, and
they disappear in the full vr − r plot of all particles (Fig-
ure 2). Caustic 3 is the most prominent one at z=0 and it
leads to a small over-density of about 10% over a smooth
density profile around 220 kpc. The abundance of parti-
cles with vr − vcaustic < 24 km s−1 increases by a factor
of about 1.6 around 220 kpc. The outermost caustics 1
and 2 are quite wide, but parts of their corresponding
branches from Figure 1 are still visible in Figure 2. Note
that all of the outer three vr − r branches have large
gaps, which are caused by epochs of little or no mass
accretion. The animated versions of Figures 1 and 2
shows how these phase-space features and the gaps in
between form and move. Whether a certain branch is
currently visible in the vr − r plane depends on whether
there was enough mass falling in at some earlier time to
reach a sufficient density in the corresponding region of
phase-space. The clearest caustic at z=0 is the third one;
it was generated by continuous, large infall of material
around z=1.7. At later epochs Via Lactea's mass accre-
tion becomes quite small and sporadic (VL07), leading to
6 For simplicity we assume this formula derived by FGB in an
Ωm = 1 universe to be approximately right also in ΛCDM.
Infall caustics in dark matter halos?
3
Positions and widths of caustics and turnaround regions in the Via Lactea halo.
TABLE 1
k
1
2
3
4
5
6
rk,med
[kpc]
rk,68%
[kpc]
∆rk
rk,med
tk,med
[kpc]
tk,68%
[kpc]
∆tk
tk,med
" rk
tk "med
" rk
tk "68% " rk
tk "FGB
rk,med
r1,med
r1 "FGB
" rk
453
310
220
173
141
121
370−534
242−384
204−237
137−207
110−191
89−170
0.36
0.46
0.15
0.41
0.57
0.67
491
343
261
222
179
157
443−551
297−407
211−316
180−266
131−229
105−201
0.22
0.32
0.40
0.39
0.55
0.61
0.92
0.93
0.84
0.78
0.78
0.81
0.77−1.12
0.57−1.24
0.67−1.10
0.58−1.25
0.52−1.46
0.54−1.46
0.876
0.864
0.856
0.843
0.832
0.834
1
0.68
0.49
0.38
0.31
0.27
1
0.65
0.49
0.40
0.34
0.30
Note. -- Positions and widths of the outer six caustics rk and turnaround radii of the particles in the caustic tk. Median values
and 68% ranges of all caustic members are given, ∆ refers to the full width of the 68% range. The number of subhalo members in
the caustics 1 to 6 are 551, 300, 49, 32, 56 and 15. "FGB" refers to predictions from the secondary infall model.
200
100
0
-100
-200
910
810
/
]
s
m
k
[
y
t
i
c
o
e
v
l
l
i
a
d
a
r
]
c
p
k
/
M
[
2
r
M
kpc
s-1
km-1
7.3
7.2
7.1
7
6.9
6.8
6.7
6.6
6.5
6.4
0
100
200
300
400
radius [kpc]
500
600
700
800
Fig. 2. -- Phase space distribution of the dark matter particles at z=0. The straight line is vcaustic(r). The lower panel shows the ρ(r) r2
of all particles (black) and of a subset with vr − vcaustic < 24 km s−1 (red). This figure is also available as an mpeg animation running
from z=12 to 0 in the electronic edition of ApJ.
weak first and second branches with large gaps. The Via
Lactea halo formed in a series of major mergers before
z=1.7. These mergers could be responsible for a broad-
ening of the fourth and higher caustics to a degree where
they become completely washed out and undetectable in
the z=0 snapshot.
3. COMPARISON WITH THE FGB-MODEL
In Table 1 we make some quantitative comparisons
with the ǫ = 1 version of the FGB-model, which cor-
responds to a point-mass as initial over-density: δ =
δM/M ∝ M −ǫ. In CDM the initial peaks are extended.
The Via Lactea halo has ǫ ≃ 0.2 up to about 1011 M⊙.
For larger masses ǫ increases and it is about one at
1012 M⊙, equivalent to a scale of 150 kpc today. The
outer halo features discussed here correspond to the ǫ ≃ 1
regime7.
The FGB-model was solved for an Einstein-de Sitter
(Ωm = 1) universe, while the Via Lactea simulation as-
sumed the now standard ΛCDM model, but we do not
In the
believe that this difference affects our results:
7 Larger collapse factors in the inner halo (see Fig. 1 in VL07)
might be related to the smaller rk/tk ratios in the ǫ = 0.2 FGB-
model.
r
4
Diemand & Kuhlen
model, the position of a caustic is similar (0.8 to 0.9) to
the turnaround radius of the caustic members. Members
of the first caustic turned around before z=1 (earlier for
higher caustics), when the ΛCDM model was still matter
dominated and similar to an Einstein-de Sitter universe.
After turnaround the enclosed mass dominates over the
enclosed vacuum energy and the dynamics are practically
the same in both cosmologies.
We find that the ratios of the current caustic posi-
tions are in good agreement with the FGB-model (Ta-
ble 1). Also the median of the ratios of caustic radii
to turnaround radii (rk/tk)med agree quite well with the
FGB-model, i.e. the typical orbits go out close to their
turnaround scale. Earlier models (Gunn 1977) assumed
they would decrease by a factor of two during virializa-
tion. This assumption is also made in the definition
of the formal virial radius8, which therefore only en-
closes a fraction of the material orbiting around a galaxy
(Prada et al. 2006,VL07). The (rk/tk)med in the clusters
from (Diemand et al. 2004a) are also close to 0.85, i.e the
orbits of bound material extend well beyond the formal
virial radius both in galaxy and cluster halos9.
The main difference between the FGB-model and cos-
mological halos is the large scatter in space and veloc-
ity around the lines in vr − r space on which all matter
is found in the model. The scatter prevents the forma-
tion of high density caustic surfaces and makes the broad
caustic-like features from cosmological infall practically
undetectable: To find them one needs an accurate, large
sample of vr − r measurements, which are available in
simulations, but unfortunately not in the real Universe.
For members of the outer four caustics the scatter in
the individual rk/tk ratios is consistent with the shape
of the potential in the outer parts of Via Lactea halo,
which is about 1:0.8 prolate at z=0 and about 1:0.7 pro-
late at z=0.5 (Kuhlen et al. 2007). We also find a few
outliers with much larger rk/tk ratios, probably resulting
from dynamical interactions between multiple satellites
(Sales et al. 2007). The scatter is significantly larger for
the 5th and 6th caustics, presumably because this mate-
rial was accreted early (z > 1.7), when a series of major
mergers caused large fluctuations in the host potential.
4. MULTIMODAL RADIAL VELOCITY DISTRIBUTIONS
From Figure 2 it is clear that the radial velocity dis-
tribution of particles in a shell in the outer halo will not
be smooth and unimodal, but rather a superposition of
modes caused by the various branches in vr − r space.
The velocity distribution and relative importance of each
mode changes with distance from the galactic center, and
fewer modes are present at large radii (Figure 3).
The subhalo radial velocity distributions are similar to
those of particles and we can use subhalo histories to
illustrate how the particle distributions divide up into
their components. Between caustics 2 and 3 one finds
a broad, doubly peaked distribution. The negative ve-
locity peak is a superposition of first infall and subha-
los approaching the center on the second half of their
first or second orbit. The subhalos on first infall have
8 The virial radius of the Via Lactea halo is rvir = r104 = 288
kpc. It encloses a mass of Mvir = 1.5 × 1012 M⊙.
9 Unlike galaxies halos, typical cluster halos do accrete a lot of
mass around z = 0 leading to a net infall between their formal
virial radius and the turnaround radius (Cuesta et al. 2007).
between 2. and
3. caustic
228 to 276 kpc
0.005
1. caustic
370 to 534 kpc
0.004
0.003
0.002
0.001
-300
-200
-100
0
0
300
0.005
100
200
3. caustic
204 to 238 kpc
-300
-200
-100
0
100
200
300
inner halo
10 to 20 kpc
]
s
1
-
m
k
[
)
r
v
(
f
]
s
1
-
m
k
[
)
r
v
(
f
0.005
0.004
0.003
0.002
0.001
0.005
0.004
0.003
0.002
0.001
0.004
0.003
0.002
0.001
0
100
200
0
300
-300
-200
-100
rv
[ km/s ]
-100
rv
0
100
200
300
[ km/s ]
0
-300
-200
Fig. 3. -- Radial velocity distributions of particles in different ra-
dial shells (solid lines). Only in the inner halo (lower right panel)
the distribution is uni-modal and close to a Gaussian (dashed line).
The other panel show multi-modal distributions. To illustrate how
to decompose these particle distributions we also plot the median
and 68% ranges of ingoing and outgoing subhalos, which have com-
pleted 0, 1 or 2 two orbits from bottom to top (circles). On the
third caustic (lower left panel) the subhalos which have completed
3 orbits are now near their apocenter. The vertical lines give the
local caustic velocity vcaustic(r) = 8/9 r/t.
the most negative typical velocities, followed by those on
their first and second orbit, in qualitative agreement with
the FGB-model. In contrast to the model, we find these
components have wide radial velocity distributions which
largely overlap and give rise to only one broad infalling
peak. On the third caustic the broad, double peaked dis-
tribution from in- and outgoing material is the same as
above, but now there is an additional peak at low radial
velocity from particles near their third apocenter.
In the inner halo particles which completed the same
number of orbits are distributed so widely that the to-
tal radial velocity distribution becomes one featureless
peak.
It is close to a Gaussian distribution with zero
mean velocity and a dispersion of 150 km s−1. But note
that a smooth, inner f (r, vr) does not preclude the ex-
istence of coherent structures in the full phase-space
density f (~x, ~v), such as bound clumps (subhalos) and
their unbound debris (tidal streams) (Moore et al. 2001;
Helmi et al. 2002).
5. SUMMARY
We have discovered broad, caustic-like features in
f (r, vr) in the outer parts of CDM halos from cosmo-
logical simulations. They are too broad to cause signifi-
cant density enhancements and would therefore be very
challenging to detect. Since a large number of accurate
distances radial velocities are needed, we believe that the
stellar halo of the Milky Way is the only system where
one might possibly detect such features in the foreseeable
future.
The basic properties of infall caustics are indepen-
dent of redshift and halo mass; they are similar in the
Via Lactea galaxy halo and in the cluster halos from
Infall caustics in dark matter halos?
5
Diemand et al. (2004a). The galaxy halos in that work
were resolved with only a few million particles per halo
and clearly show similar features. A new, one billion
particle galaxy halo (Diemand et al. in prep.) also only
shows broad, weak outer caustics. Since the features de-
scribed here do not change qualitatively when the nu-
merical resolution is increased by almost three orders of
magnitude we conclude that they are well resolved in
simulations with a few million particles per halo.
We thank Ed Bertschinger and Roya Mohayaee for
helpful discussions and Peter Goldreich, Roya Mohayaee
and Scott Tremaine for helpful comments on an earlier
version of this letter. We acknowledge support from
NASA through a Hubble Fellowship grant (JD) and and
from the Hansmann Fellowship at the Institute for Ad-
vanced Study (MK).
REFERENCES
Ascasibar, Y., Hoffman, Y., & Gottlober, S. 2007, MNRAS, 376,
393
Bergstrom, L., Edsjo, J., & Gunnarsson, C. 2001, Phys. Rev. D,
63, 083515
Bertschinger, E. 1985, ApJS, 58, 39
Cuesta, A. J., Prada, F., Klypin, A., & Moles, M. 2007, ArXiv
e-prints, 710
Diemand, J., Kuhlen, M., & Madau, P. 2007, ApJ, 667, 859
Diemand, J., Madau, P., & Moore, B. 2005, MNRAS, 364, 367
Diemand, J., Moore, B., & Stadel, J. 2004a, MNRAS, 352, 535
Diemand, J., Moore, B., Stadel, J., & Kazantzidis, S. 2004b,
MNRAS, 348, 977
Dubinski, J. & Carlberg, R. G. 1991, ApJ, 378, 496
Faltenbacher, A. & Diemand, J. 2006, MNRAS, 369, 1698
Fillmore, J. A. & Goldreich, P. 1984, ApJ, 281, 1
Gott, J. R. I. 1975, ApJ, 201, 296
Gunn, J. E. 1977, ApJ, 218, 592
Helmi, A., White, S. D., & Springel, V. 2002, Phys. Rev. D, 66,
063502
Hoffman, Y. & Shaham, J. 1985, ApJ, 297, 16
Kuhlen, M., Diemand, J., & Madau, P. 2007, ApJ, 671, 1135
Madau, P., Diemand, J., & Kuhlen, M. 2008, ArXiv e-prints, 802
Mohayaee, R. & Shandarin, S. F. 2006, MNRAS, 366, 1217
Moore, B. 2001, in Identification of Dark Matter, ed. N. J. C.
Spooner & V. Kudryavtsev, 93 -- +
Moore, B., Calc´aneo-Rold´an, C., Stadel, J., Quinn, T., Lake, G.,
Ghigna, S., & Governato, F. 2001, Phys. Rev. D, 64, 063508
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1996, ApJ, 462,
563
Prada, F., Klypin, A. A., Simonneau, E., Betancort-Rijo, J.,
Patiri, S., Gottlober, S., & Sanchez-Conde, M. A. 2006, ApJ,
645, 1001
Ryden, B. S. 1993, ApJ, 418, 4
Sales, L. V., Navarro, J. F., Abadi, M. G., & Steinmetz, M. 2007,
MNRAS, 379, 1475
Sikivie, P., Tkachev, I. I., & Wang, Y. 1997, Phys. Rev. D, 56,
1863
Zaroubi, S., Naim, A., & Hoffman, Y. 1996, ApJ, 457, 50
|
astro-ph/0405115 | 2 | 0405 | 2005-07-14T11:05:41 | The Hubble constant from gravitational lens CLASS B0218+357 using the Advanced Camera for Surveys | [
"astro-ph"
] | We present deep optical observations of the gravitational lens system CLASS B0218+357, from which we derive an estimate of the Hubble constant. Extensive radio observations have reduced the degeneracies between Hubble's constant and the mass model in this lens to one involving only the position of the radio-quiet lensing galaxy relative to the lensed images. B0218+357 has an image separation of only 334 mas, so optical observations have previously been unable to resolve the lens galaxy from the bright lensed images. Using the new Advanced Camera for Surveys installed on the Hubble Space Telescope, we have measured the separation between the lens galaxy centre and the brightest image. The position found, and hence our estimate of Hubble's constant, depends on our approach to the spiral arms in B0218+357. If the most prominent arms are left unmasked, we find a value for Hubble's constant of 70+/-5 km/s /Mpc (95% confidence). If the spiral arms are masked out, we find a value of 61+/-7 km/s /Mpc (95% confidence). | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 15 (2004)
Printed 30 November 2018
(MN LATEX style file v2.2)
The Hubble Constant from gravitational lens CLASS B0218+357
using the Advanced Camera for Surveys
T. York1, N. Jackson1, I.W.A. Browne1, O. Wucknitz2, J.E. Skelton1†
1University of Manchester, Jodrell Bank Observatory, Macclesfield, Cheshire, SK11 9DL
2Universitat Potsdam, Institut fur Physik, Am Neuen Palais 10, 14469 Potsdam, Germany
†Current address: Institute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill, EH9 3HJ
30 November 2018
ABSTRACT
We present deep optical observations of the gravitational lens system CLASS B0218+357
from which we derive an estimate for the Hubble Constant (H0). Extensive radio observations
using the VLA, MERLIN, the VLBA and VLBI have reduced the degeneracies between H0
and the mass model parameters in this lens to one involving only the position of the radio-
quiet lensing galaxy with respect to the lensed images. B0218+357 has an image separation
of only 334 mas, so optical observations have, up until now, been unable to resolve the lens
galaxy from the bright lensed images. Using the new Advanced Camera for Surveys, installed
on the Hubble Space Telescope in 2002, we have obtained deep optical images of the lens
system and surrounding field. These observations have allowed us to determine the separation
between the lens galaxy centre and the brightest image, and so estimate H0.
We find an optical galaxy position -- and hence an H0 value -- that varies depending
on our approach to the spiral arms in B0218+357. If the most prominent spiral arms are left
unmasked, we find H0 = 70±5 km s−1 Mpc−1(95% confidence). If the spiral arms are masked
out we find H0 = 61±7 km s−1 Mpc−1(95% confidence).
Key words: gravitational lensing, distance scale
1 INTRODUCTION
Objects at cosmological redshifts may be multi-
ply imaged by the action of the gravitational field
of foreground galaxies. The first such example
of gravitational lensing was the system 0957+561
(Walsh, Carswell & Weymann 1979) in which the
core of a background quasar is split into two im-
ages 6′′ apart. Since then approximately 70 cases
of gravitational lensing by galaxies have been
found.1
Refsdal (1964) pointed out that such multiple-
image gravitational lens systems could be used to
measure the Hubble constant, if the background
source was variable, by measuring time delays be-
tween variations of the lensed image and inferring
the difference in path lengths between the corre-
sponding ray paths. The combination of typical
1 A full compilation of known galaxy-mass lens systems is given on the
CASTLeS website at http://cfa-www.harvard.edu/glensdata
deflection angles of ∼1′′ around galaxy-mass lens
systems with typical cosmological distances im-
plies time delays of the order of weeks, which
are in principle readily measurable. Time delays
have been measured for eleven gravitational lenses
to date: CLASS B0218+357 (Biggs et al. 1999;
Cohen et al. 2000), RXJ 0911.4+0551 (Hjorth
et al. 2002), 0957+561 (Kundic et al. 1997; Os-
coz et al. 2001), PG 1115+080 (Schechter et al.
1997) CLASS B1422+231 (Patnaik & Narasimha
2001), SBS 1520+530 (Burud et al. 2002a),
CLASS B1600+434 (Koopmans et al. 2000; Bu-
rud et al. 2002b), CLASS B1608+656 (Fassnacht
et al. 1999; Fassnacht et al. 2002), PKS 1830−211
(Lovell et al. 1998), HE 2149−2745 (Burud et al.
2002b) and HE 1104−1805 (Ofek & Maoz 2003).
In principle, given a suitable variable source, the
accuracy of the time delay obtained can be bet-
ter than 5%. This has already been achieved in
some cases (eg. 0957+561) and there is no doubt
2
that diligent future campaigns will further improve
accuracy and also produce time delays for more
gravitational lens systems.
Gravitational
lenses provide an excellent
prospect of a one-step determination of H0 on cos-
mological scales. The major problem is that, in
addition to the time delay, a mass model for the
lensing galaxy is required in order to determine
the shape of the gravitational potential. The model
is needed to convert the time delays into angu-
lar diameter distances for the lens and source. In
double-image lens systems in which the individ-
ual images are unresolved this is a serious problem
as the number of constraints on the mass model
(lensed image positions and fluxes) allows no de-
grees of freedom after the most basic parameters
characteristic of the system (source position and
flux together with galaxy mass, ellipticity and po-
sition angle) have been fitted. In four-image sys-
tems the extra constraints provide assistance, and
in a few cases, such as the ten-image lens system
CLASS B1933+503 (Sykes et al. 1998) more de-
tailed constraints on the galaxy mass model are ex-
ploited (Cohn et al. 2001).
There are two further systematic and poten-
tially very serious problems. The first is that the
radial mass profile of the lens is almost com-
pletely degenerate with the time delay, and hence
H0 (Gorenstein, Shapiro & Falco 1988; Witt, Mao
& Keeton 2000; Kochanek 2002). Given a time
delay, H0 scales as 2 − β, where β is the profile
index of the potential, φ ∝ rβ. Work by Koop-
mans & Treu (2003) shows that mass profiles may
vary from an isothermal slope by up to 10% for
single galaxies, producing corresponding uncer-
tainties in H0. The problem is particularly seri-
ous for four-image systems, because the images
are all approximately the same distance from the
centre of the lens and thus constrain the radial pro-
file of the lensing potential poorly. On the other
hand, for CLASS B1933+503, with three sources
producing ten images, the radial mass profile is
well constrained (Cohn et al. 2001). Unfortunately
B1933+503 does not show radio variability (Biggs
et al. 2000) and is optically so faint that measuring
a time delay is likely to be very hard.
In some cases Einstein rings may pro-
vide enough constraints, despite the necessity to
model the extended source which produces them
(Kochanek, Keeton & McLeod 2001) although
models constrained by rings may still be degen-
erate in H0 (Saha & Williams 2001).
The mass profile degeneracy is particularly
sharply illustrated by "non-parametric" modelling
of lens galaxies (Williams & Saha 2000; Saha
& Williams 2001). Such models assume only
basic physical constraints on the galaxy mass
profile, such as a monotonic decrease in sur-
face density with radius. They find consistency
with the observed image data for a wide range
of galaxy mass models, which are themselves
consistent with a wide range of H0. Combin-
ing two well-constrained cases of lenses with a
measured time delay, CLASS B1608+656 and
PG 1115+080, Williams & Saha (2000) find H0 =
61 ± 18km s−1 Mpc−1(90% confidence).
There are a number of approaches to the res-
olution of the mass profile degeneracy problem.
One is to assume that galaxies have approximately
isothermal mass distributions. (β ∼ 1). There are
two parts to the lensing argument in favour of
isothermal galaxies: from the lack of odd images
near the centre of observed lens systems Rusin &
Ma (2001) are able to reject the hypothesis that
significant number of lensing galaxies have pro-
files which are much shallower (β > 1.2) than
isothermal, assuming a single power-law model is
appropriate for the mass contained interior to the
lensed images. Similarly it can be shown that mod-
els which are significantly steeper than isother-
mal are unable to reproduce constraints from posi-
tions and fluxes in existing lenses with large num-
bers of constraints (e.g. Munoz, Kochanek & Kee-
ton 2001; Cohn et al. 2001). The most straight-
forward approach, that of assuming an isothermal
lens, has been taken by many authors. In most
cases this yields H0 estimates of between 55 and
70 km s−1 Mpc−1 (e.g. Biggs et al. 1999; Koop-
mans & Fassnacht 1999; Koopmans et al. 2000;
Fassnacht et al. 1999) but studies of some lenses
imply much lower values (e.g. Schechter et al.
1997; Barkana 1997; Kochanek 2003). In fact,
Kochanek (2003) finds a serious discrepancy with
the HST key project value of H0=71km s−1 Mpc−1
(Mould et al. 2000; Freedman et al. 2001) unless,
far from being isothermal, galaxy mass profiles
follow the light distribution.
Falco, Gorenstein & Shapiro (1985) pointed
The Hubble Constant from CLASS B0218+357 using the Advanced Camera for Surveys
3
out the second important problem. Any nearby
cluster produces a contribution to the lensing po-
tential in the form of a convergence which is
highly degenerate with the overall scale of the
lensing system and hence with H0. Unfortunately,
the systems with the most accurately determined
time delays and the best-known galaxy positions
are often those with large angular separation such
as 0957+561, and these are the systems in which
lensing is most likely to be assisted by a clus-
ter. Again progress can be made by appropriate
modelling of the cluster, and many attempts have
been made to do this for 0957+561 (e.g. Kundic
et al. 1997; Bernstein & Fischer 1999; Barkana
et al. 1999) although there remain uncertainties
in the final H0 estimate. As an alternative, the
optical/infra-red images of the host galaxy may
make an important contribution towards the break-
ing of degeneracies (Keeton et al. 2000).
Kochanek & Schechter (2004) summarise the
contribution of lensing to the H0 debate so far
and present options for further progress. One ap-
proach is simply to accumulate more H0 determi-
nations and rely on statistical arguments to iron
out the peculiarities which affect each individual
lens system; this approach is vulnerable only to a
systematically incorrect understanding of galaxy
mass profiles. The alternative approach is to select
a lens system in which additional observational ef-
fort is most capable of decreasing the systematic
errors on the H0 estimate to acceptable levels. In
this paper we argue that CLASS B0218+357 is the
best candidate for this process. We describe new
Hubble Space Telescope (HST) observations using
the Advanced Camera for Surveys (ACS) which
are aimed at removing the last major source of
systematic uncertainty in this system. We then de-
scribe how we use the imaging data to derive a
value for the Hubble constant.
2 CLASS B0218+357 AS A KEY OBJECT IN H0
DETERMINATION
CLASS B0218+357 was discovered during the
early phase of the CLASS survey, the Jodrell
Bank-VLA Astrometric Survey (JVAS; Patnaik et
al. 1992). B0218+357 consists of two images (A
and B) of a background flat-spectrum radio source
separated by 0.′′334, together with an Einstein ring
(Patnaik et al. 1993). The optical spectrum shows
a red continuum source superimposed on a galaxy
spectrum. The redshift of the lensing galaxy has
been measured optically by Browne et al. (1993)
and Stickel & Kuehr (1993), and by Carilli, Ru-
pen & Yanny (1993) at radio wavelengths giving
the most accurate result of 0.6847. Cohen et al.
(2003) have measured the source redshift of 0.944.
It quickly became apparent that the lensing
was performed by a spiral galaxy. Spiral lenses
generally produce smaller image separations than
elliptical lenses because of their lower mass. The
spiral nature of B0218+357 was deduced directly
from early high-resolution optical images from
the Nordic Optical Telescope (Grundahl & Hjorth
1995), and was consistent with evidence from
molecular line studies which revealed absorption
of the radio emission from the background radio
source by species in the lensing galaxy includ-
ing CO, HCO+, HCN (Wiklind & Combes 1995),
formaldehyde (Menten & Reid 1996) and water
(Combes & Wiklind 1997). Moreover, in the opti-
cal, the A image, which is further from the galaxy,
is fainter than the B image (Grundahl & Hjorth
1995) , despite being a factor ∼3 brighter in the ra-
dio. This suggests that the line of sight to the A im-
age intercepts a great deal of dust, possibly asso-
ciated with a giant molecular cloud in the galaxy.
The lensing galaxy appears close to face-on, a con-
clusion deduced from its symmetrical appearance
in optical images. This conclusion is consistent
with the small velocity line-width of the absorp-
tion lines (e.g. Wiklind & Combes 1995).
Further radio imaging resolved both the A and
B images into core-jet structures (Patnaik, Porcas
& Browne 1995; Kemball et al. 2001; Biggs et al.
2003), as well as revealing more details of the Ein-
stein ring (Biggs et al. 2001). The combined con-
straints from the core-jet structure and the ring to-
gether strongly constrain mass models. A and B
lie at different distances from the galaxy, and with
the Einstein ring constraints permits determination
of both the angular structure of the lensing mass
(Wucknitz et al. 2004) and (most importantly) the
mass -- radius relation for the lens.
Biggs et al. (1999) have determined a time
delay of 10.5±0.4 days (95% confidence) for
B0218+357 using radio monitoring observations
made with the VLA at both 8.4 GHz and 15 GHz.
4
Consistent results were obtained from the varia-
tions in the total intensity, the percentage polar-
ization and the polarization position angle. Biggs
et al. used the time delay and existing lens model
to deduce a value for the Hubble constant of
69+13
−19 km s−1 Mpc−1(95% confidence). It should
be noted, however, that the error bars on the as-
sumed position for the lensing galaxy with respect
to the lensed images were over-optimistic and
hence their quoted error on H0 is too small. Cohen
et al. (2000) also observed B0218+357 with the
VLA, and measured a value for the time delay of
−1.6 days. This corresponds to an H0 value of
10.1+1.5
71+17
−21 km s−1 Mpc−1(95% confidence), the larger
error bars in Cohen et al.'s measurement being due
to their use of a more general model for the source
variability, although they used the same model for
the lensing effect as Biggs et al. The error bars do
not take into account any systematic error associ-
ated with the uncertain galaxy position.
Leh´ar et al. (2000) used the then existing con-
straints to model the CLASS B0218+357 sys-
tem. They found that, even for isothermal mod-
els, the implied value of H0 was degenerate with
the position of the centre of the lensing galaxy,
with a change of about 0.7 km s−1 Mpc−1(about
1 per cent) in the value of H0 for every 1 mas shift
in the central galaxy position. Their uncertainty on
the position derived from HST infra-red observa-
tions using the NICMOS camera is approximately
±30 mas.
Recently, using a modified version of the
LENSCLEAN algorithm (Kochanek & Narayan
1992; Ellithorpe et al. 1996; Wucknitz 2004),
Wucknitz et al. (2004) have been able to constrain
the lens position from radio data of the Einstein
ring. With the time delay from Biggs et al. (1999)
of (10.5 ± 0.4) days, they obtain for isothermal
models a value of H0 = 78±6 km s−1 Mpc−1(95%
confidence). They use VLBI results from other au-
thors (Patnaik et al. 1995, Kemball et al. 2001) as
well as their own data (Biggs et al. 2003) to con-
strain the radial profile from the image substruc-
ture and obtain β ≈ 1.04, very close to isothermal
(β = 1). Our aim in this paper is to use new optical
observations to determine the lensing galaxy posi-
tion directly and to compare this with the indirect
determination of Wucknitz et al. (2004).
We briefly summarise the reasons why,
given the observations presented here, CLASS
B0218+357 offers the prospect of the most unbi-
ased and accurate estimate of H0 to date.
(i) The observational constraints are arguably
the best available for any lens system with a mea-
sured time-delay.
(ii) The radio source is relatively bright (a few
hundred mJy at GHz frequencies) and variable
at radio frequencies, so time delay monitoring is
relatively straightforward and gives a small error
(Biggs et al. 1999) which can be improved with
future observations.
(iii) The system is at relatively low redshift.
This means that the derived values for H0 will not
depend on the matter density parameter and cos-
mological constant by more than a few percent.
(iv) The lens is an isolated single galaxy and
there are no field galaxies nearby to contribute to
the lensing potential (Leh´ar et al., 2000).
Although most lenses have at least one of these
desirable properties, CLASS B0218+357 is the
only one known so far which has all of them. It
thus becomes a key object for H0 determination.
It is only the lack of an accurate galaxy position
that led in the past to it being excluded from con-
sideration by many authors (e.g. Schechter 2001;
Kochanek 2003).
3 THE ACS OBSERVATIONS
Resolving the lens galaxy and lensed images is an
aim that benefits from high resolution combined
with high dynamic range, and so we asked for and
were awarded time on the Advanced Camera for
Surveys (ACS; Clampin et al. 2000) on the Hubble
Space Telescope.
Although observations in the blue end of the
spectrum would have maximised angular resolu-
tion, the likely morphological type of the lensing
galaxy (Sa/Sab) meant that asymmetry due to star
formation in the spiral arms could have caused
problems in the deconvolution process which re-
lies on symmetry in the lensing galaxy (see Sec-
tion 5). Hence observations at wavelengths longer
than 4000 A in the rest frame of the galaxy were
desirable. At the redshift of the lens this dictated
the use of I band, i.e. the F814W filter on the HST.
The ACS has two optical/near-IR "channels",
The Hubble Constant from CLASS B0218+357 using the Advanced Camera for Surveys
5
ing visit (16) was designed to permit the pro-
gramme to be salvaged in the unlikely case that the
observing pattern chosen for visits 10-15 proved
to be both inappropriate and uncorrectable. This
visit suffered from increased observing overheads
relative to the other visits and provided an integra-
tion time of 1 hour, 22 minutes on B0218+357 as
a result. In order to deconvolve the images we re-
quired an ACS/WFC PSF, so two short visits (1
and 2 in Table 1) were dedicated to observing two
Landolt standard stars (Landolt 1992). Follow-
ing McLure et al. (1999), observations were taken
with several different exposure times to allow the
construction of a composite PSF that would have
good signal-to-noise in both core and wings whilst
avoiding saturation of the core. Standards were se-
lected to be faint enough not to saturate the WFC
chip on short integration times, and to have the
same V − I colour, to within 0.2 magnitudes, as
the lensed images in the B0218+357 system. The
exposure times on the standard stars ranged from
0.5 seconds to 100 seconds each, the longest ex-
posures each being split into two 50 second expo-
sures to simplify cosmic ray rejection.
4 REDUCTION OF THE ACS DATA
The uncalibrated data produced by automatic pro-
cessing of raw HST telemetry files by the OPUS
pipeline at the Space Telescope Science Institute
(STScI) were retrieved along with flat fields, su-
perdarks and other calibration files.
The data were processed through the ACS cali-
bration pipeline, CALACS (Pavlovsky et al. 2002),
which runs under NOAO's IRAF software. The
CALACS pipeline de-biased, dark-subtracted and
flat-fielded the data, producing a series of cali-
brated exposures. The pipeline also combined the
CR-SPLIT exposures in visits 1, 2 and 16 to elim-
inate cosmic rays. The calibrated exposures were
in general of acceptable quality for use in the next
stage of reduction, except for visit 15 in which
there was some contamination of the images by
stray light, probably from a WFPC2 calibration
lamp (R. Gilliland and M. Sirianni, private com-
munication). It is possible that this defect can be
corrected in the future using the techniques which
were used by Williams et al. (1996) to remove
stray light from some HDF exposures, but we
Figure 1. 4+4 dither pattern used for the observations of 0218+357. The
pattern provides steps to the level of 1
4 -pixel.
the Wide Field Channel (WFC) and the High
Resolution Channel (HRC). The HRC's pixel re-
sponse exhibits a diffuse halo at longer wave-
lengths due to scattering within the CCD. As a re-
sult, roughly 10% of the flux from a point source
will be scattered into this halo at 8000 A, possi-
bly making it more difficult to resolve the lens-
ing galaxy from the lensed images. We selected
the WFC for use in our observing programme
since it does not suffer from this effect. Unlike
the HRC, the WFC moderately under-samples the
HST point-spread function (PSF) at 8000 A. To
counter this effect, we selected two distinct four-
point dither patterns, alternating between them
over the course of the observations. The dither pat-
terns used are shown in Figure 1 (see also Mutch-
ler & Cox 2001 for more information on HST
dither patterns).
The WFC has a field of view of 202′′×202′′,
and a plate scale of approximately 50 mas pixel−1.
Since we used a gain of unity, saturation in the
images occurs near the 16-bit analogue to digital
conversion limit of 65,000 e− pixel−1 rather than
at the WFC's full well point of 85,000 e− pixel−1.
We determined the exposure time required on
B0218+357 through simulation.
The full programme of B0218+357 observa-
tions was carried out over the period from Au-
gust 2002 to March 2003. Details of the observ-
ing dates and exposure times are shown in Table 1.
The total available telescope time was split into 7
visits on B0218+357, 6 of which provided 2 hours
integration time on the science target. The remain-
6
Visit no.
Target
Observation date
Exposure time
Dither pattern
File name root
10
11
12
13
15
14
16
1
2
CLASS B0218+357
CLASS B0218+357
CLASS B0218+357
CLASS B0218+357
CLASS B0218+357
CLASS B0218+357
CLASS B0218+357
2003 February 28
2003 March 01
2003 January 17-18
2003 March 06
2003 March 11
2002 October 26-27
2002 September 11
92 245
2002 October 17-18
PG0231+051B
2002 August 25
20×360 sec
20×360 sec
20×360 sec
20×360 sec
20×360 sec
20×360 sec
12×360 sec
8×75 sec
8×0.5 sec
8×8 sec
1×360 sec
8×0.5 sec
8×8 sec
8×50 sec
4+4
4+4
4+4
4+4
4+4
4+4
4+4
4C
4+4
4+4
-
4+4
4+4
4C
j8d410
j8d411
j8d412
j8d413
j8d415
j8d414
j8d416
j8d401
j8d402
Table 1. Log of HST observations. All observations were taken with the ACS using the F814W filter, corresponding to I band. A 4+4 dither pattern refers to
an eight-point dither consisting of two nested parallelograms, whereas 4C refers to a four-point dither parallelogram with the exposure at each point split into
two for explicit cosmic-ray rejection.
have not attempted to deconvolve the contami-
nated visit.
The calibrated exposures were fed to the next
stage of reduction, based around the dither pack-
age (Fruchter & Hook 2002), and the STSDAS
packages pydrizzle (Hack 2002) and multidrizzle
(Koekemoer et al. 2002). These tools clean cos-
mic rays, remove the ACS geometric distortion
and "drizzle" the data on to a common output im-
age (Fruchter & Hook 2002). The drizzle method
projects the input images on to a finer grid of out-
put pixels. Flux from each input pixel is distributed
to output pixels according to the degree of overlap
between the input pixel and each output pixel. To
successfully combine dithered images into a single
output image, knowledge of the pointing offsets
between exposures is needed. The expected offsets
are determined by the dither pattern used, but the
true offsets might vary from those expected due
to thermal effects (Mack et al. 2003) within sin-
gle visits. To determine the true pointing offsets
between dithered exposures, we cross-correlated
the images. Since the WFC at I-band moderately
undersamples the HST's PSF, and since most of
the features detected in the images were extended
rather than stellar, we opted for pixel-by-pixel
cross-correlation rather than comparisons of posi-
tions of stars between different pointings, to max-
imise our use of the available information.
Images in each visit were drizzled on to a
common distortion-corrected frame and then pairs
of these images were cross-correlated. The two-
dimensional cross-correlations have a Gaussian
shape near their centres. The shift between pairs
of images is measured by fitting a Gaussian func-
tion to the peak in the cross-correlation, and the
estimated random error in the shift is derived from
the position error given by the Gaussian fit. For
our data, the random errors in the measured shifts
ranged from 0.8 to 2.5 mas. The RMS scatter be-
tween corresponding pointings within a visit was
typically less than 10 mas, or 20% of a single
WFC pixel. We fed these shifts to the multidriz-
zle script, which carried out the drizzling of visits
to common, undistorted output frames. As part of
the drizzling process, we opted to decrease the out-
put pixel size from 50 mas square (the natural size
of the undistorted output pixels) to 25 mas square.
To avoid blurring and "holes" in the output im-
age, the input pixels were shrunk to 70% of their
nominal size before being drizzled on to the output
frames (Fruchter & Hook 2002). We used a gaus-
sian drizzling kernel to slightly improve resolu-
tion and reduce blurring. At the end of this process
each visit, except for visit 15, provided us with a
output image with improved sampling compared
to the individual input exposures. The deconvolu-
tion of these images is described in Section 5.
Since deconvolution depends greatly on the
accuracy of the PSF model, we have produced a
number of different PSFs. Unfortunately the Lan-
dolt standard star (Landolt 92 245) observed in
visit 1 was resolved into a 0.5′′ double by the
ACS/WFC, so we have concentrated on extracting
a PSF from visit 2 (Landolt PG0231+051B). The
calibrated exposures were combined using mul-
The Hubble Constant from CLASS B0218+357 using the Advanced Camera for Surveys
7
tidrizzle. Saturated pixels were masked. The re-
sulting PSF suffered from serious artifacts consist-
ing of extended wings approximately 80 mas up
and down the chip from the central peak, possibly
due to imperfect removal and combination of pix-
els which were affected by bleeding of saturated
columns. We believe these extended wings to be
artifacts since they rotate with the telescope rather
than the sky, and are not present in stars in other
visits.
In lieu of ideal standard star PSFs, we have
created per-visit PSFs by averaging field stars to-
gether. These field star PSFs do not suffer from the
artifacts present in the visit 2 PSF. Manual exam-
ination, pixel-by-pixel, of the difference between
the PSFs and the central regions close to image
B, indicate that the RMS error in the PSFs is be-
tween 5 and 15%. Such an error increases linearly
with the counts, rather than with their square root;
lacking a perfectly-fitting PSF, we have allowed
for this error when performing the data analysis
described in Section 5.
5 ANALYSIS OF THE POSITION DATA AND RESULTS
5.1 General remarks
Figure 2 shows the ACS image of the CLASS
B0218+357 system from the combined dataset,
consisting of all science visits on B0218+357 ex-
cluding visit 15. To produce this combined im-
age, separate visits were related through restricted
linear transformations (rotation and translation)
based on the positions of unsaturated stars com-
mon to all images, and then co-added. In locating
the lensing galaxy, however, we did not use this
combined image, preferring to work with the sepa-
rate visits. The two compact images (A and B) can
be distinguished as can the lensing galaxy which
lies close to B. The spiral arms of the galaxy are
clearly seen confirming the earlier deductions that
the lensing galaxy is a spiral (Wiklind & Combes
1995; Carilli et al. 1993). The spiral arms appear
to be smooth and regular and there is no sign of
significant clumping associated with large-scale
star formation. The galaxy appears almost exactly
face-on. We deduce this by assuming a galaxy po-
sition close to B and comparing counts between
pixels at 90 degree angles from each other about
Figure 2. Combined ACS image of B0218+357. The lensed images are both
visible; the brighter image, B, is close to the centre of the lensing galaxy.
The spiral arms of the galaxy are clearly visible. The plot above the image
shows a one dimensional slice passing through images A and B. The best-fit
positions of A and B on this slice are marked, along with A', the position
of A expected from the radio image separation (334 mas). The separation
between A and B in the optical image is 317±4 mas (2 σ).
the assumed centre. Examination of the residuals
reveals no sign of ellipticity.
The core of the lensing galaxy is strongly
blended with B (Figure 2) and is relatively weak.
Extrapolation of an exponential disk fit to the outer
isophotes of a slice through the central region
shows that the peak surface brightness of image
B exceeds that of the galaxy by a factor of about
30-50. Thus the determination of an accurate po-
sition for the galaxy is a challenging task. Before
discussing the process in more detail below, we
outline the various steps we go through to obtain a
galaxy position. They are:
(i) For each visit we measure the positions of
the A and B images.
(ii) We subtract PSFs from these positions.
(iii) Using PSF-subtracted data we look for the
8
galaxy position about which the residuals left af-
ter PSF subtraction appear most symmetric. We do
not subtract a galaxy model from the images. This
approach finds the centre of the most symmetric
galaxy consistent with the data.
(iv) We compute the mean and variance of the
galaxy positions found from the individual visits.
5.2 Analysis procedure
are
reduced ACS/WFC images
Our
over
8450x8500 pixels in size, and cover 202′′x202′′on
the sky. We cut out a region of 128x128 pixels
(3.2′′x3.2′′) centred on the lens and analyse this
to make fitting computationally practical and to
isolate the lens from other objects on the sky.
Since the images, particularly image B, have
much higher surface brightnesses than the galaxy,
their positions can be located relatively accurately
by subtracting parametric or empirical PSFs from
the data and minimising the residuals. Fits were
carried out on circular regions centred on the
brightest pixel of each compact image. The re-
gions chosen were 11 pixels in diameter. To avoid
any bias arising from the choice of PSF, we used
both parametric models (Airy and Gaussian func-
tions) and the field star PSFs in the fits. The field
star PSFs were consistently better fits to both A
and B than the parametric models, Gaussians be-
ing insufficiently peaked and Airy functions hav-
ing diffraction rings that were too prominent. A
linear sloping background was modelled along
with the PSF in order to take account of the flux
due to the galaxy. Typically the various methods
agreed on positions to within a tenth of a pixel
(2.5 mas).
The separation of A and B determined by op-
tical PSF fitting is consistently less than the radio
separation of 334 mas. We find that the mean im-
age separation in the optical is 317±2 mas (1 σ)
when the field star PSFs are used, and 315±4 mas
when Gaussian PSFs are used. The correspond-
ing result for Airy function PSFs is 311±10 mas.
These values are mean separations taken over
the six processed visits on B0218+357. We have
checked the plate scale of drizzled images against
stars listed in the US Naval Observatory's B1.0
catalogue, and find that the nominal drizzled plate
scale of 25 mas pixel−1 is correct to better than 1%
for all visit images.
An anomalous optical image separation has
been suggested before for B0218+357, starting
with ground-based optical imaging by Grundahl &
Hjorth (1995) and again by Hjorth (1997). Jack-
son, Xanthopoulos & Browne (2000) used NIC-
MOS imaging to find an image separation of
318±5 mas, in agreement with our result from
field star PSFs.
We hypothesise that that this low separation
may be a result of the high, and possibly spatially
variable, extinction in the region of A. We suggest
that some of the image A optical emission arises
from the host galaxy rather than from the AGN
which dominates the B image emission. Thus the
centroid of A may not be coincident with the AGN
image. Image A may be obscured completely and
the emission seen could be due to a large region of
star formation associated with an obscuring giant
molecular cloud. In view of this possibility and the
fact that B is consistently much brighter than A in
the optical, we measure galaxy positions as offsets
from our measured position for B. Thus, although
A's position may be distorted in the optical it does
not directly influence our measurements of H0.
Having determined positions and fluxes for A
and B, we created an image containing two PSFs
as a model for the flux from the lensed images
alone (the "model image"). A model image was
made separately for each visit and subtracted from
each observed image, leaving a residual image
which contained only the galaxy plus subtraction
errors. Figure 3 shows a typical image with A and
B subtracted. The model allowed us to keep track
of how much PSF flux was removed from each
pixel in producing the residual images.
We opted to use the criterion of maximum
symmetry in the residuals as a goodness-of-fit pa-
rameter rather than attempting to fit a parametric
model, such as an exponential disk profile, to the
light distribution of the galaxy. The symmetry fit
statistic is expressed as
χ2 = X
P
s(r) − s(r
σ(r)2 + σ(r
′)2
′)2 ,
(1)
in which s(r) is the count-rate at image pixel
The Hubble Constant from CLASS B0218+357 using the Advanced Camera for Surveys
9
age. We have checked this noise estimate against
the background noise in our images and against
simulations of the drizzling process.
The set of pixels (P) included in the calculation
of this χ2 figure can bias the fit if it is ill-chosen.
When r
′ falls outside the boundaries of the image,
the pixel r is considered to contribute nothing to
the χ2 statistic and the pixel is not included in the
set P. Such pixels therefore do not contribute to the
number of constraints available and as a result do
not increase the number of degrees of freedom in
the fit. Alternative treatments can introduce bias;
for instance, if these pixels are assigned large χ2
values the fitting program is biased towards plac-
ing the galaxy in the geometrical centre of the im-
age. If the same pixels are considered to contribute
zero towards the χ2 statistic but are still counted
as part of the set P, they increase the number of
degrees of freedom in the fit and bias the fit to
positions away from the image's geometrical cen-
tre. To avoid these possibilities we do not count
degrees of freedom from pixels whose reflection
about the galaxy centre ends up outside the image
boundaries.
The symmetry criterion is non-parametric and
has the advantage of minimizing the assumptions
that are imposed on the data; the use of a particu-
lar distribution as a function of radius in any case
contains an implicit assumption of symmetry. Us-
ing the symmetry criterion on its own is in princi-
ple robust whether or not the galaxy has a central
bulge, and should also be unaffected if the galaxy
contains a bar. The symmetry criterion will also
hold for galaxies having moderate inclinations to
the line of sight. For a circularly symmetric galaxy,
the main effect of a small deviation away from a
face-on orientation will be to render the observed
image slightly elliptical. The basic symmetry cri-
terion is that points and their reflections about
the true centre of the galaxy's image should have
the same flux (to within the measurement errors).
Therefore whether the galaxy's image has circu-
lar or slightly elliptical isophotes is unimportant
because in both cases the same isophote passes
through both point and reflection. This argument
breaks down for spirals with significant inclina-
tions as absorption is likely to become important
and destroy any symmetry present in the image.
The centre of maximum symmetry could be
Figure 3. A visit image with A and B subtracted after fitting fluxes and posi-
tions. The residuals near the centre of each subtracted image show maxima
of approximately 20% of the unsubtracted light.
position r = (x, y) in counts per second, σ(r) is
the estimated noise at r, P is the set of pixels in-
cluded in the fit (see below) and r
′ is the reflection
of r around the galaxy position g, given by simple
geometric considerations as
′ = 2g − r.
r
(2)
The random error for each pixel in the image,
σ(r) is estimated from the CCD equation (Merline
& Howell 1995) summed with a contribution from
the assumed random error in the PSF. The estimate
of the noise is given by
σ(r)2 =
q4
τ 2
q−2τ s(r) + S + R2
N
+ µ2sp(r)2, (3)
where τ is the integration time at a single
dithered pointing, µ is the assumed fractional er-
ror in the PSF, sp(r) is the count-rate from A and
B alone (stored in the model image), q is the driz-
zling scale factor (the ratio between output and in-
put pixel size, 0.5 for these images) N is the num-
ber of dithered pointings combined in the drizzle
process, S is the sky noise and R is the ACS/WFC
read noise. Both sky noise and read noise are ex-
pressed in units of electrons, and the integration
time is in seconds. The resulting noise figure has
units of counts per second, as does the drizzled im-
10
displaced by spiral arms if they are not themselves
symmetric about the galaxy centre. To analyse the
possible effect on H0 of spiral arms we have made
two sets of fits. In the first set we used all the data
and did not apply any masking. In the second set
we masked off the most obvious spiral arms by
using an annular mask centred on image B with
an inner radius of 0.375′′and an outer radius of
0.875′′. Regions within the inner radius or beyond
the outer radius of the mask were left free to con-
tribute to the symmetry fit. We implemented the
symmetry fits so that masked pixels did not con-
tribute to either the number of degrees of freedom
or to the χ2 value.
The PSF error (µ in equation 3) can cause a
systematic change in galaxy position when varied
between 0.05 and 0.15 (5 to 15%). An increase
of µ from 0.05 to 0.15 can increase H0 by up to
10 km s−1 Mpc−1. For values above 0.15 the sys-
tematic change in galaxy position is small com-
pared with the random error. We estimate the PSF
error separately for each visit, by taking the range
of the highest and lowest residuals and dividing
that range by the peak count-rate of image B. We
find values between 0.07 (visit 11) and 0.19 (visit
10) for µ. For the other visits, the estimated value
for µ is found to be 0.12. In the remainder of the
paper we do not allow µ to vary freely but fix it to
these estimated values.
5.3 Extraction of the galaxy position
In applying the symmetry criterion to 0218 we
calculated the symmetry χ2 statistic (i.e. that of
equation 1) for a grid of galaxy positions extend-
ing 20 mas east to 100 mas west of B, and from
80 mas south to 50 mas north of B. The spacing
between adjacent grid points is 5 mas. We present
these grids in Figure 4 for visits 10, 11, 12, 13, 14
and 16. The grids are shown both with and without
masking of spiral arms. Table 2 lists the resulting
galaxy positions.
In Figure 5 we show the per-pixel χ2 contri-
butions between the two cases of no masking and
masked spiral arms for the optimum galaxy po-
sition, together with the contributions when no
masking is used and zero PSF error is applied.
With zero PSF error it is clear that the residu-
als from the subtraction of A and B dominate the
Visit
Centre
Centre
(No masking)
∆δ
∆α
+50
+6
−4
+60
+9
+59
−2
+54
+0
+59
+61
−6
(Spiral arms masked)
∆α
+70
+69
+84
+72
+76
+79
∆δ
+12
−18
+8
−5
−16
−14
10
11
12
13
14
16
Mean +57±4 +1±6 +75±6
−6±13
Table 2. Derived optical centre of the galaxy, expressed as offsets in mas
from the measured optical position of B. RA offsets are given with west as
positive.
Figure 5. These images show the contribution of each pixel to the symme-
try χ2 for visit 11, using the best-fit optical galaxy positions. The top-left
image shows the per-pixel χ2 contributions when no masking is used and
no PSF error is used. The image on the top-right shows the contributions
when no masking is applied and a 7% PSF error is used. The bottom-left
image shows the contributions when spiral arms are masked, with no PSF
error. The bottom-right image shows the χ2 contributions when the promi-
nent spiral arms are masked and a 7% PSF error is used. All images are 128
pixels (3.2′′) in both width and height.
χ2 measure. The effect of a non-zero PSF error
is to suppress the A/B subtraction residuals and
cause the spiral arms to dominate the χ2 mea-
sure, unless they are masked. Because it is unclear
which position (masked or unmasked) best repre-
sents the mass centre of the lensing galaxy, we re-
port both masked and unmasked galaxy positions
(and hence estimates for H0) on equal terms.
Deriving errors on position from the individ-
ual visits is difficult because the symmetry χ2 in-
creases very rapidly away from the minimum. An
error measure derived from the shape of the min-
The Hubble Constant from CLASS B0218+357 using the Advanced Camera for Surveys
11
No masking
Spiral arms
No masking
Spiral arms
10
11
12
13
14
16
Figure 4. χ2 grids for the galaxy position. The visit number is shown to the right of each pair of grids. The right-hand plot for each visit shows the effect of
masking out the spiral arms, whilst the left-hand plot shows the χ2 grid when no masking is applied. The position of B is marked, as is a line pointing towards
the (radio) A component. The ellipse represents the position of the galaxy centre found by Wucknitz et al. (2004) using LENSCLEAN modelling of the Einstein
ring, and the dotted lines represent H0 of (90,80,70,60,50)km s−1 Mpc−1 from left to right, assuming an isothermal model. The axes are RA/Dec offsets from
the position of B, expressed in arc-seconds. The RA offset is given with west as positive.
imum for a single visit implies a spuriously high
accuracy for the galaxy position. It is likely that
the number of degrees of freedom in the fit is over-
estimated and that many pixels do not contribute
any useful information to the fit statistic, since the
drizzling process introduces correlations between
neighbouring drizzled pixels. However, the scatter
between positions derived from different visits is
large. We therefore estimate errors on the galaxy
position by taking ellipses that enclose 68% and
95% of the measurements from all visits to define
our 1 σ and 2 σ confidence levels. Figure 6 shows
the 95% confidence ellipses on the galaxy position
for both sets of fits, as well as the position derived
from LENSCLEAN applied to VLA data by Wuck-
nitz et al. (2004).
The positions obtained from the symmetry fits
were combined with the extra constraints avail-
able from the VLBI substructure described in Pat-
naik et al. (1995) and Kemball et al. (2001), which
were used to constrain mass models by Wucknitz
(2004). The optical galaxy position was combined
with the models of Wucknitz (2004) by adding χ2
values for the galaxy position to the χ2 values from
the lens models. However, the χ2 values taken
from the symmetry fitting grid have too many de-
grees of freedom, and so we assume that the opti-
cal position minimum is parabolic and form a new
χ2 statistic based on our 68% and 95% confidence
ellipses. We define the new χ2 statistic to have a
value of 2.31 on our 68% confidence ellipse, and
a value of 5.99 on the 95% ellipse, and sum this
statistic with that from the lens modelling. We em-
phasize that the scatter between visits dominates
the random error budget for our measurement of
H0.
Combining the VLBI and optical constraints
shows that the best-fit galaxy position and the opti-
cal galaxy position are not coincident, as shown in
Table 3. The galaxy position shifts by up to 13 mas
12
between the optical fit and the optical+VLBI fit.
The shapes of the confidence regions are also al-
tered. However, the value of H0 is not very sensi-
tive to this mainly northerly shift.
6 EXTRACTION OF H0
The general relation between the time delay ∆ti,j
between the ith and jth images, the Hubble con-
stant H0 and the lens model, parametrised by the
potential ψ, is given by
c ∆ti,j =
1 + zd
H0
dd ds
dds
(φi − φj)
,
(4)
where zd is the redshift of the lens, dd and
ds are the angular size distances to the lens and
source, respectively, dds is the angular size dis-
tance to the source measured from the lens, and
φi is the scaled time delay at the position of the ith
image (θi),
φi =
1
2
∇ψ(θi)2 − ψ(θi)
.
(5)
The angular size distances are normalized in
these equations, since they do not include factors
of H0. For general isothermal models without ex-
ternal shear the relation becomes particularly sim-
ple and can be written as a function of the image
positions alone, without explicitly using any lens
model parameters (Witt et al. 2000):
φi =
1
2
θi − θ02
(6)
Here θ0 is the position of the centre of the lens. Ex-
ternal shear γ changes φi by a factor between 1±γ
depending on the relative direction, typically re-
sulting in similar factors for the value deduced for
H0. A general analysis for power-law models with
external shear can be found in Wucknitz (2002).
Using the recipe described in previous sections
our lens position translates to a Hubble constant of
H0 = 79±7 km s−1 Mpc−1in the shearless isother-
mal case2 without masking, and to 66 ± 9 with
masking.
2 A concordance cosmological model with Ω = 0.3 and λ = 0.7 and a
homogeneous matter distribution is used for the calculation of all distances
in this paper
Estimates of external shear and convergence
from nearby field galaxies and large scale struc-
ture are of the order 2 per cent (Leh´ar et al. 2000)
and would affect the result only to the same rel-
ative amount, sufficiently below our current error
estimate to allow us to neglect these effects.
The value of the Hubble constant we derive
depends on the slope of the mass distribution of
the lensing galaxy. In Figure 7 we show the per-
mitted values of the Hubble constant for different
models -- isothermal and with a variable β in an el-
liptical potential model -- plotted against measured
galaxy position. The elliptical power-law potential
is parametrised as follows:
ψ(θ) =
θ2−β
E
β
rβ
ǫ (θ)
,
r2
ǫ =
θ2
x
(1 + ǫ)2 +
θ = (θx, θy)
(1 − ǫ)2
θ2
y
,
,
(7)
(8)
(9)
where θE is the Einstein radius of the model, β is
the power-law index of the potential's profile and
ǫ is the ellipticity of the potential. For details of
our modelling procedure the reader is referred to
Wucknitz et al. (2004). It is evident that the pre-
ferred value of the Hubble constant is somewhat
reduced compared to what is obtained by forcing
the mass distribution to be isothermal. We also
show contours of the radial power law β plotted
against galaxy position. The optical lens position
gives β =1.13+0.07
−0.09 (2 σ).
As discussed before, B0218+357 has the ad-
vantage of clear substructure in the two images
which can be mapped with VLBI. The VLBI data
can be used independently to derive the slope
of the mass profile of the lens (Wucknitz et al.,
2004). Biggs et al. (2003) and Wucknitz et al. find
a value of β=1.04±0.02. Combining the VLBI
constraints with the optical lens position gives
a value of β=1.05±0.03 (95% confidence). We
therefore adopt this value for the mass profile's
logarithmic slope and obtain a Hubble constant of
70±5 km s−1 Mpc−1(95% confidence) for the case
with no masking and 61±7 km s−1 Mpc−1(95%
confidence) when the spiral arms are masked. The
ellipticity of the potential is small (about 0.04) in
each case, but the ellipticity of the mass distribu-
tion will be about three times this, 0.12. The lens
The Hubble Constant from CLASS B0218+357 using the Advanced Camera for Surveys
13
Figure 6. The optical galaxy position compared to that determined by Wucknitz et al. (2004) using LENSCLEAN. The error ellipses are 95% confidence
regions. The left-hand plot is for the case with no masking. The right-hand plot represents the case with masking of spiral arms. The position of B is marked,
as is a line pointing towards the A component. The dotted lines are contours of H0 in the strictly isothermal case, and correspond (from right to left) to H0 =
(90,80,70,60,50) km s−1 Mpc−1. The positions determined from optical data are shown as bold ellipses.
Data used
Masking Mass profile
Position (mas)
∆α
∆δ
H0
H0
(isothermal)
(variable β)
β
Ellipticity
(of potential)
Optical
Optical
VLBI+Optical
VLBI+Optical
VLBI+Optical
VLBI+Optical
None
Spiral
None
Spiral
None
Spiral
-
-
Isothermal
Isothermal
Variable
Variable
+57
+75
+60
+74
+60
+74
+0
−5
−13
−19
−12
−18
79±7
66±9
74±5
64±7
-
-
68±6
56+12
−15
-
-
70±5
61±7
1.13+0.07
−0.09
1.16±0.19
-
-
1.05±0.03
1.05±0.04
0.08±0.03
0.05±0.04
0.05±0.02
0.03±0.02
0.04±0.02
0.04±0.02
Table 3. Lens galaxy positions and the corresponding values of H0 and the mass profile slope β. The optical positions are derived from the ACS images only.
The "VLBI+Optical" positions incorporate constraints from the LENSCLEAN-based lens modelling of Wucknitz et al. (2004). The "Mass profile" column
indicates what mass profile was assumed when combining the VLBI and optical constraints. H0 values are given in km s−1 Mpc−1. Position offsets are
referenced to image B, and RA offsets are given taking west as positive. All errors are quoted at 95% confidence.
galaxy could be more inclined than it appears, or
there could be a bar or similar feature present.
7 CONCLUSIONS
We have analysed the deepest optical image yet
taken of B0218+357 to measure the position of the
lens galaxy. We find that simple subtraction of a
parametric galaxy model and two point sources is
insufficient to constrain the galaxy position, and
we confirm earlier suggestions that the image sep-
aration in the optical is lower than that in the radio,
most probably due to significant extinction around
image A. Taking advantage of the symmetric ap-
pearance of the lens, we have defined the centre
as that point about which the residuals (after sub-
traction of A and B) are most symmetric. To ac-
count for artifacts in our empirical PSF model we
have introduced an extra noise term. We have also
masked off the most prominent spiral arms to test
the effect on H0. We find that the lens galaxy po-
sition is 57±4 mas west and 0±6 mas south of
image B when no masking is applied. Combined
with the results of Wucknitz et al. (2004) this leads
to a value for H0 of 70±5 km s−1 Mpc−1(95%
confidence). When the most obvious spiral arms
are masked out, we find an optical galaxy posi-
tion of 75±6 mas west and −5±13 mas south
from image B. This results in a value for H0 of
61±7 km s−1 Mpc−1(95% confidence) when com-
bined with VLBI constraints.
Further work on this lens will involve in-
creased use of LENSCLEAN to further limit the
power law exponent β using VLBI constraints.
Observations have also been made using the VLA
with the Pie Town VLBA antenna, which together
with VLBI will further improve the lens model for
this system.
14
Figure 7. The 95% confidence regions for the cases of no masking (top row) and masked spiral arms (bottom row). The plots in the left-hand column show
the galaxy position confidence regions and contours of H0 calculated for isothermal models. The plots of the central column show the confidence regions
superimposed over contours of H0 calculated whilst allowing the logarithmic mass slope β to vary. This reduces H0 slightly compared to the fixed β = 1
(isothermal) case. The plots in the right-hand column show the same confidence regions superimposed over contours of β. The confidence regions take both
our optical position and the VLBI constraints used by Wucknitz et al. (2004) into account. The 95% confidence error ellipse of Wucknitz et al. is also shown.
Contours of H0 are again at (90,80,70,60,50)km s−1 Mpc−1reading from left to right across an image, and contours of β are (0.8,0.9,1.0,1.1,1.2,1.3) reading
from bottom to top. The optical+VLBI error ellipses are shown in bold relative to the LENSCLEAN ellipse of Wucknitz et al.
ACKNOWLEDGMENTS
The Hubble Space Telescope is operated at the
Space Telescope Science Institute by Associated
Universities for Research in Astronomy Inc. un-
der NASA grant NAS5-26555. We would like
to thank the ACS team, especially Warren Hack,
Max Mutchler, Anton Koekemoer and Nadezhda
Dencheva for help and advice on the data reduc-
tion. We would also like to thank the referee, Chris
Fassnacht, whose comments greatly improved the
substance of the paper.
TY acknowledges a PPARC research stu-
dentship. OW was supported by the BMBF/DLR
Verbundforschung under grant 50 OR 0208. This
research has made use of the NASA/IPAC Ex-
tragalactic Database. IRAF is distributed by the
National Optical Astronomy Observatories, which
are operated by AURA, Inc., under cooperative
agreement with the National Science Foundation.
REFERENCES
Barkana R., 1997, ApJ, 489, 21
Barkana R., Leh´ar J., Falco E.E., Grogin N.A., Keeton C.R.,
Shapiro I.I., 1999, ApJ, 520, 479
Bernstein G., Fischer P., 1999, AJ, 118, 14
Biggs A.D., Browne I.W.A., Helbig P., Koopmans L.V.E., Wilkin-
son P.N., Perley R.A., 1999, MNRAS, 304, 349
Biggs A.D., Xanthopoulos E., Browne I.W.A., Koopmans L.V.E.,
Fassnacht C.D., 2000, MNRAS, 318, 73
Biggs A.D., Browne I.W.A., Muxlow T.W.B., Wilkinson P.N.,
2001, MNRAS, 322, 821
Biggs A.D., Wucknitz O., Porcas R.W., Browne I.W.A., Jackson
N.J., Mao S., Wilkinson P.N., 2003, MNRAS, 338, 599
Browne I.W.A., Patnaik A.R., Walsh D., Wilkinson P.N., 1993,
MNRAS, 263, L32
Burud I., Hjorth J., Jaunsen A.O., Andersen M.I., Korhonen H.,
Clasen J.W., Pelt J., Pijpers F.P., Magain P., Ostensen R., 2000,
ApJ, 544, 117
The Hubble Constant from CLASS B0218+357 using the Advanced Camera for Surveys
15
E.A., Tzioumis A.K., McCulloch P.M., Edwards P.G., 1998, ApJ,
508, L51
Mack J., et al., 2003, ACS Data Handbook, Version 2.0 (Balti-
more:STScI)
McLure R.J., Kukula M.J., Dunlop J.S., Baum S.A., O'Dea C.P.,
Hughes D.H., 1999, MNRAS, 308, 377
Menten K.M., Reid M.J., 1996, ApJL, 465, L99
Merline W., Howell S.B., 1995, Expt. Astron., 6, 163
Mould, J.R., et al., 2000, ApJ, 529, 786
Munoz J.A., Kochanek C.S., Keeton C.R., 2001, ApJ, 558, 657
Mutchler M., Cox C., 2001, Instrument Science Report ACS 2001-
07 (Baltimore: STScI)
Ofek E.O., Maoz D., 2003, ApJ, 594, 101
Oscoz A., Alcalde D., SerraRicart M., Mediavilla E., Abajas C.,
Barrena R., Licandro J., Motta V., Munoz J.A., 2001, ApJ, 552, 81
Patnaik A.R., Browne I.W.A., Wilkinson P.N., Wrobel J.M., 1992,
MNRAS, 254, 655
Patnaik A.R., Browne I.W.A., King L.J., Muxlow T.W.B., Walsh
D., Wilkinson P.N., 1993, MNRAS, 261, 435
Patnaik A.R., Porcas R.W., Browne I.W.A., 1995, MNRAS, 274,
L5
Patnaik A.R., Narasimha D., 2001, MNRAS, 326, 1403
Pavlovsky C., et al., 2002, ACS Instrument Handbook, Version 3.0
(Baltimore: STScI).
Refsdal S., 1964, MNRAS, 128, 307
Rusin D., Ma C., 2001, ApJL, 549, L33
Saha P., Williams L.L.R., 2001, AJ, 122, 585
Schechter P.L., 2001, in Brainerd T.G., Kochanek C.S., eds, ASP
Conf. Ser. Vol. 237, Gravitational Lensing: Recent Progress and
Future Goals. Astron. Soc. Pac., San Francisco, p. 427
Schechter P.L., Bailyn C.D., Barr R., Barvainis R., Becker C.M.,
Bernstein G.M., Blakeslee J.P., Bus S.J., Dressler A., Falco E.E., et
al., 1997, ApJL, 475, L85
Stickel M., Kuehr H., 1993, A&AS, 101, 521
Sykes C.M., et al., 1998, MNRAS, 301, 310
Walsh D., Carswell R.F., Weymann R.J., 1979, Nature, 279, 381
Wiklind T., Combes F., 1995, A&A, 299, 382
Williams R.E., et al., 1996, AJ, 112, 1335
Williams L.L.R., Saha P., 2000, AJ, 119, 439
Witt H.J., Mao S., Keeton C.R., 2000, ApJ, 544, 98
Wucknitz O., 2002, MNRAS, 332, 951
Wucknitz O., 2004, MNRAS, 349, 1
Wucknitz O., Biggs A.D., Browne I.W.A., 2004, MNRAS, 349, 14
Burud I., Hjorth J., Courbin F., Cohen J.G., Magain P., Jaunsen
A.O., Kaas A.A., Faure C., Letawe G., 2002a, A&A, 391, 481
Burud I., et al., 2002b, A&A, 383, 71
Carilli C.L., Rupen M.P., Yanny B., 1993, ApJL, 412, L59
Clampin M., et al., 2000, SPIE, 4013, 344
Cohen, A.S., Hewitt, J.N., Moore, C.B., Haarsma, D.B., 2000, ApJ,
545, 578
Cohen J.G., Lawrence C.R., Blandford R.D., 2003, ApJ, 583, 67
Cohn J.D., Kochanek C.S., McLeod B.A., Keeton C.R., 2001, ApJ,
554, 1216
Combes F., Wiklind T., 1997, ApJ, 486, L79
Ellithorpe J.D., Kochanek C.S., Hewitt J.N., 1996, ApJ, 464, 556
Falco E.E., Gorenstein M.V., Shapiro I.I., 1985, ApJL, 289, L1
Fassnacht C.D., Pearson T.J., Readhead A.C.S., Browne I.W.A.,
Koopmans L.V.E., Myers S.T., Wilkinson P.N., 1999, ApJ, 527, 498
Fassnacht C.D., Xanthopoulos E., Koopmans L.V.E., Rusin D.,
2002, ApJ, 581, 823
Freedman W.L., Madore B.F., Gibson B.K., Ferrarese L., Kelson
D.D., Sakai S., Mould J.R., Kennicutt R.C., Jr., Ford H.C., Graham
J.A., et al., 2001, ApJ, 553, 47
Fruchter A.S., Hook R.N., 2002, PASP, 114, 144
Gorenstein M.V., Shapiro I.I., Falco E.E., 1988, ApJ, 327, 693
Grundahl F., Hjorth J., 1995, MNRAS, 275, L67
Hack W.J., 2002, in Bohlender D., Durand D., Handley T.H., eds,
ASP Conf. Ser. Vol. 281, Astronomical Data Analysis Software and
Systems XI. Astron. Soc. Pac., San Francisco, p.197.
Hjorth J., 1997, Helbig P., Jackson N., eds, Proc. Golden Lenses,
Hubble's Constant and Galaxies at High Redshift Workshop. Jo-
drell Bank Observatory, Cheshire.3
Hjorth J., et al., 2002, ApJ, 572, L11
Jackson N., Xanthopoulos E., Browne I.W.A., 2000, MNRAS, 311,
389
Keeton C.R., et al., 2000, ApJ, 542, 74
Kemball A.J., Patnaik A.R., Porcas R.W., 2001, ApJ, 562, 649
Kochanek C.S., Keeton C.R., McLeod B.A., 2001, ApJ, 547, 50
Kochanek C.S., Narayan R., 1992, ApJ, 401, 461
Kochanek C.S., 2002, ApJ, 578, 25
Kochanek C.S., 2003, ApJ, 583, 49
Kochanek C.S., Schechter P., 2004, in "Measuring and Modelling
the Universe", Carnegie Obs. Centennial Symposium, CUP, ed. W.
Freedman, p. 117
Koekemoer A.M., Fruchter A.S., Hook R.N., Hack W., 2002, in Ar-
riba S., ed., The 2002 HST Calibration Workshop : Hubble after the
Installation of the ACS and the NICMOS Cooling System. Space
Telescope Science Institute, Baltimore, MD, p. 339
Koopmans L.V.E., de Bruyn A.G., Xanthopoulos E., Fassnacht
C.D., 2000, A&A, 356, 391
Koopmans L.V.E., Fassnacht C.D., 1999, ApJ, 527, 513
Koopmans L.V.E., Treu T., 2003, ApJ, 583, 606
Krist J., 1995, in Shaw R.A., Payne H.E., Hayes J.J.E., eds, ASP
Conf. Ser. Vol. 77, Astronomical Data Analysis Software and Sys-
tems IV. Astron. Soc. Pac., San Francisco, p. 349
Kundic T., Hogg D.W., Blandford R.D., Cohen J.G., Lubin L.M.,
Larkin J.E., 1997, AJ, 114, 2276
Landolt A.U., 1992, AJ, 104, 340
Leh´ar J., Falco E.E., Kochanek C.S., McLeod B.A., Munoz J.A.,
Impey C., Rix H., Keeton C.R., Peng C.Y., 2000, ApJ, 536, 584
Lovell J.E.J., Jauncey D.L., Reynolds J.E., Wieringa M.H., King
3 Proceedings available from http://www.jb.man.ac.uk/research/gravlens/workshop1/prcdngs.html
|
0709.0381 | 3 | 0709 | 2008-07-01T20:28:14 | An Unexpectedly Swift Rise in the Gamma-ray Burst Rate | [
"astro-ph"
] | The association of long gamma-ray bursts with supernovae naturally suggests that the cosmic GRB rate should trace the star formation history. Finding otherwise would provide important clues concerning these rare, curious phenomena. Using a new estimate of Swift GRB energetics to construct a sample of 36 luminous GRBs with redshifts in the range z=0-4, we find evidence of enhanced evolution in the GRB rate, with ~4 times as many GRBs observed at z~4 than expected from star formation measurements. This direct and empirical demonstration of needed additional evolution is a new result. It is consistent with theoretical expectations from metallicity effects, but other causes remain possible, and we consider them systematically. | astro-ph | astro-ph |
DRAFT VERSION NOVEMBER 7, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
AN UNEXPECTEDLY SWIFT RISE IN THE GAMMA-RAY BURST RATE
MATTHEW D. KISTLER1,2, HASAN Y UKSEL1,2, JOHN F. BEACOM1,2,3, KRZYSZTOF Z. STANEK3,2
Draft version November 7, 2018
ABSTRACT
The association of long gamma-ray bursts with supernovae naturally suggests that the cosmic GRB rate
should trace the star formation history. Finding otherwise would provide important clues concerning these
rare, curious phenomena. Using a new estimate of Swift GRB energetics to construct a sample of 36 luminous
GRBs with redshifts in the range z = 0 − 4, we find evidence of enhanced evolution in the GRB rate, with
∼ 4 times as many GRBs observed at z ≈ 4 than expected from star formation measurements. This direct
and empirical demonstration of needed additional evolution is a new result. It is consistent with theoretical
expectations from metallicity effects, but other causes remain possible, and we consider them systematically.
Subject headings: gamma-rays: bursts -- cosmology: theory
1. INTRODUCTION
in astrophysics.
Long gamma-ray bursts are perhaps the grandest spec-
tacles
Their connection with core-
collapse supernovae (Stanek et al. 2003; Hjorth et al. 2003)
indicates progenitors that are very massive,
short-lived
leading one to expect the cosmic GRB history to
stars,
follow the star
(Totani 1997;
Wijers et al. 1998; Lamb & Reichart 2000; Blain & Natarajan
2000; Porciani & Madau 2001). Any deviation from this ex-
pectation would provide new information about why a star
should die as a GRB, complementing microphysical investi-
gations (see Meszaros (2006) for a review).
formation history (SFH)
2006;
In just
Daigne et al.
Bromm & Loeb
the past few years, Swift (Gehrels et al. 2004)
and a worldwide network of observers have detected
gamma-ray bursts from higher
redshifts than was pre-
viously possible, sparking renewed interest
in the GRB
redshift distribution (Berger et al. 2005; Natarajan et al.
2005;
2006;
Jakobsson et al. 2006; Le & Dermer 2007; Yuksel & Kistler
2007; Salvaterra & Chincarini 2007; Liang et al. 2007;
Guetta & Piran 2007; Chary et al. 2007). The Swift bursts
with known redshifts now provide a sufficiently large sample
upon which reasonable cuts can be made. Additionally, im-
proved star formation measurements, compiled in the analysis
of Hopkins & Beacom (2006), now provide a well-defined
baseline for comparison. Using a simple, model-independent
test, we find that the GRB rate history appears to evolve more
strongly than the SFH until at least z ≈ 4. We emphasize that
our result establishes the existence of an evolutionary trend,
but that, at present, we cannot identify the underlying reason
(i.e., the distinction between correlation and causality). We
discuss possible causes, including observational effects, the
suggestion that the lower metallicity of the high-z universe
could allow more bursts, and other possibilities that may
result in additional GRB progenitors.
2. THE EXPECTED GRB RATE
1 Dept. of Physics, The Ohio State University, 191 W. Woodruff Ave.,
Columbus, OH 43210
2 Center for Cosmology and Astro-Particle Physics, The Ohio State Uni-
versity, 191 W. Woodruff Ave., Columbus, OH 43210
3 Dept. of Astronomy, The Ohio State University, 140 W. 18th Ave.,
Columbus, OH 43210
An ever-increasing amount of data has brought about a
clearer picture of the history of cosmic star formation. As
shown in Fig. 1, after a sharp rise up to z ≈ 1, the SFH
is nearly flat until z ≈ 4 with relatively small uncertainties.
These measurements are well-fit by a simple piecewise power
law parametrization (Hopkins & Beacom 2006),
ρSFH(z) ∝
: z < 0.97
(1 + z)3.44
(1 + z)−0.26 : 0.97 < z < 4.48
(1 + z)−7.8 : 4.48 < z ,
scaled to ρSF(0) = 0.0197 M⊙ yr−1 Mpc−3 (a rate per co-
moving volume). We parametrize the intrinsic (comoving)
GRB rate relative to the SFH as nGRB(z) = E(z) × ρSF(z).
The fraction of bursts that can be seen at a given z, 0 < F (z) <
1, depends on the ability to detect the initial burst of gamma
rays and to obtain a redshift from the optical afterglow. We
cast the distribution of observable GRBs as
d N
dz
= F (z)
E(z) ρSF(z)
hfbeami
dV /dz
1 + z
.
(1)
This being an observed rate, cosmological time dilation
requires the (1 + z)−1.
The comoving volume ele-
ment (in terms of the comoving distance, dc), dV /dz =
8
8
13
13
1
4
4
t [Gyr]
t [Gyr]
2
2
1
1
]
3
-
c
p
M
1
-
r
y
O.
M
[
R
F
S
ρ.
]
3
c
p
G
[
z
d
/
V
d
z
+
1
10-1
10-2
200
100
0
0
1
2
3
z
4
5
6
FIG. 1. -- Top panel: The history of star formation. Shown are the collection
of data (circles) and fit of Hopkins & Beacom (2006). Bottom panel: The
"volumetric factor" in Eq. (1), which distorts the observed GRB distribution
when viewed in z.
KISTLER, Y UKSEL, BEACOM & STANEK
1
1
2
2
3
3
2
0
0
1053
1052
1051
]
1
-
s
g
r
e
[
o
s
i
L
1050
1049
Q [(c/H0)3]
Q [(c/H0)3]
4
4
5
5
6
6
7
7
8
8
9
9
Absorption
Emission
Photometric
Q
d
/
N
d
0
15
10
5
0
0
Q [(c/H0)3]
4
5
6
7
8
9
1
2
3
All GRBs
Liso > 1051 erg/s
1
2
3
z
4
5
6 7
FIG. 3. -- The differential GRB distribution versus Q. Outlined bins contain
the set of all 63 GRBs (median z = 2.3), while shaded bins contain the 44
bursts with Liso > 1051 erg s−1 (with a median of z = 2.9). The rise seen
suggests that GRB rate evolves more strongly than star formation.
l in Fig. 2).
burst detection (Gehrels et al. 2004; Band 2006). While this
sensitivity is "very difficult if not impossible" (Band 2006) to
parametrize exactly, an effective luminosity threshold appears
to present in the data (roughly estimated as ∝ d2
A representative sample of bursts in a given redshift range
can be selected, while avoiding detailed assumptions concern-
ing this threshold, by simply choosing a lower cutoff in Liso.
Considering only bursts that could have been seen from any-
where within the redshift range examined allows for the Swift
contribution to the F (z) term to be treated as constant in
z. This technique effectively integrates the GRB luminos-
ity function, dN/dLiso, above the chosen Liso cut without
assuming its functional form. This reduces the problem to
number counts.
4. TESTING FOR EVOLUTION
Because ρSF(z) is now reasonably well-measured from
z ≈ 0 − 4 (and nearly flat for 1 < z < 4), we consider GRBs
in this range for comparison. Using only bursts with Liso >
1051 erg s−1 creates a set of GRBs with a minimal expected
loss of GRBs up to z . 4. Fig. 3 displays the differential dis-
tribution of these GRBs versus Q. The bursts above this cut
(shaded bins) can be compared to the set of all bursts (out-
lined). As can clearly be seen, removing the population of
lower luminosity bursts that are only observable at low z re-
veals a distinct rise in the observed number of "bright" bursts.
The drop at z & 4 is likely due to the Swift threshold and an
overall drop in star formation.
To compare the GRB data with the SFH expectation, we
make use of Eq. (1), parameterizing the "effective evolution",
for simplicity, as F (z) × E(z) ∝ (1 + z)α. Since the lumi-
nosity cut removes the influence of the Swift threshold, any
z-dependence of F would be due to other potential observa-
tional effects. We compare the predicted and observed cumu-
lative GRB distributions in Fig. 4. This compares the relative
trends, independent of the overall normalization and the value
of hfbeami. A Kolmogorov-Smirnov test reveals that the SFH
fit alone is incompatible at around the 95% level.
Positive evolution results in clear improvement, strength-
ening indications based upon different analyses and smaller
data sets (e.g., Daigne et al. (2006); Le & Dermer (2007);
Yuksel & Kistler (2007)). We find that the K-S statistic is
minimized for α = 1.5. To simulate dispersion in the data,
we replace the T90 values used to calculate Liso with those
0
0
1
1
2
2
3
3
z
z
4
4
5
5
6 7
6 7
FIG. 2. -- The luminosities (Liso) of 63 Swift gamma-ray bursts, determined
from the data of Butler et al. (2007), vs. z (two z < 0.1 GRBs are below
1048 erg s−1). Points are plotted linearly in Q(z), as given in Eq. (2), to
account for the volumetric factor. Open symbols may have underestimated
Liso. The shaded region approximates the effective Swift detection threshold.
Redshifts were measured as denoted, with medians of z = 0.8 for emission
and z = 2.9 for absorption.
c (z)/p(1 + z)3 Ωm + ΩΛ, peaks at z ∼ 2.5.
4π (c/H0) d2
Dividing dV /dz by the 1 + z term yields a volumetric fac-
tor that peaks at z ∼ 1.4 (see Fig. 1)4. For a constant
F (z), relatively fewer bursts should be observed at z ∼ 4
than z ∼ 1. GRBs that are unobservable due to beaming
are accounted for through hfbeami (see Bloom et al. (2003);
Guetta et al. (2005); Panaitescu (2007); Kocevski & Butler
(2007); Nava et al. (2007)).
3. THE SWIFT OBSERVATIONS
Swift has enabled observers to extend the reach of GRB ob-
servations greatly compared to pre-Swift times, resulting in
a rich data set. We consider bursts from the Swift archive5
up to 2007 May 15 with reliable redshifts and durations ex-
ceeding T90 > 2 sec. We estimate each GRB luminosity as
Liso = Eiso/[T90/(1 + z)], where Eiso is the isotropic equiv-
alent (beaming-uncorrected) 1 − 104 keV rest-frame energy
release and T90 is the interval observed to contain 90% of the
prompt GRB emission. To form a uniform Liso set, as dis-
played in Fig. 2, we use the Eiso and T90 values given by
Butler et al. (2007). Note that Liso for several bursts (open
symbols) may be underestimated due to T90 values that over-
estimate the GRB duration.
To make the visual density of GRBs in Fig. 2 more mean-
ingful, we define a linear x-coordinate that "corrects" for the
effects of the volumetric factor as
Q(z) = Z z
0
dz ′ dV /dz ′
1 + z ′
,
(2)
which we will quote in terms of (c/H0)3 ≈ 79 Gpc3. Remov-
ing the influence of the volumetric factor allows for a bet-
ter "by-eye" view, since d N /dQ = (d N /dz)/(dQ/dz) =
F (Q) nGRB(Q)/ hfbeami, so that a flat GRB rate would ap-
pear as a constant density of GRBs with similar Liso per linear
interval in Q.
The Swift
is quite complex, working in the
15 − 150 keV band and using time-dependent background
subtraction and variable time windows in order to maximize
trigger
4 Using ΩΛ = 0.7, Ωm = 0.3, and H0 = 70 km/s/Mpc; changing these
requires correcting the SFH (see Hopkins (2004))
5 http://swift.gsfc.nasa.gov/docs/swift/archive/grb table
A SWIFT RISE IN THE GRB RATE
3
0
1
1
2
Q [(c/H0)3]
4
3
5
6
)
4
<
(
N
/
)
z
<
(
N
0.5
0
0
1
2
z
3
4
FIG. 4. -- The cumulative distribution of the 36 Swift GRBs with Liso >
1051 erg s−1 in z = 0 − 4 (steps), compared to the expectations from an
effective evolution of (1 + z)1.5 with respect to ρSF(z) (dark solid); and the
SFH alone (dashed), which is inconsistent with the data at ∼ 95%.
reported by Swift (Sakamoto et al. 2007) and find similar con-
clusions. In neither case is any significant correlation of lumi-
nosity with z found in this range, disfavoring just an evolving
luminosity function (Efron & Petrosian 1992). Although we
have attempted to minimize the loss of GRBs due to the Swift
threshold, we are more likely to be underestimating burst
counts at the upper end of the z-range examined, which could
imply even stronger evolution.
5. IS THE EVOLUTION INTRINSIC?
This evolution could result from several causes. The K-S
test disfavors an interpretation as a statistical anomaly. While
ρSF(z) could just be mismeasured, the relatively-small uncer-
tainties over the range in question suggests that this is not the
likely origin. We will discuss in more detail whether it is due
to some selection effect (i.e., a changing F (z)), or a real phys-
ical effect related to changes in the number of progenitors (an
evolving E(z)).
It has been suggested that a higher percentage of GRBs
may go undetected at low redshifts (Bloom 2003; Fiore et al.
2007). We perform several simple diagnostics to determine
whether the sample of Swift GRBs without redshifts contains
an overwhelming number of bright, low-z bursts. To simplify
analysis, we consider a sample of bursts that meet the "de-
tectability" criteria of Jakobsson et al. (2006) (in particular,
low Galactic extinction and quick Swift localization), a set of
50 GRBs with a confirmed z and 47 without. With this set,
using the procedure of Sect. 4, the SFH expectation alone is
still incompatible at around 95%, with α = 1.5 minimizing the
K-S statistic.
On average, more distant bursts are expected to have lower
observed gamma-ray fluxes, F, estimated by dividing each
15 − 350 keV fluence and T90 from Butler et al. (2007). In-
deed, GRBs at z > 2 have hFz>2i ∼ 1.4 × 10−7 erg cm−2
s−1, compared to hFz<2i ∼ 2.6 × 10−7 erg cm−2 s−1 at
z < 2. For GRBs without a redshift, the average flux is just
hFno zi ∼ 0.7 × 10−7 erg cm−2 s−1. A two-sample K-S
test between the z < 2 and z-less sets reveals that they are in-
compatible at ∼70%. Limiting the z < 2 set only to GRBs
with Liso > 1051 erg s−1, this increases to ∼99%. While not
conclusive, these results could be interpreted as most z-less
bursts being either at high z, in which case our evolution may
be underestimated, or at low z with lower intrinsic luminosi-
ties, which may not survive our Liso cut. Additionally, we ex-
amine the fractions of GRBs detected by Swift's UV-Optical
Telescope in the Swift archive. Of bursts with a known z, 34
were seen by UVOT with hzi = 2.2, while the 16 not seen
have hzi = 3.0. Bursts lacking redshifts seem more consis-
tent with the high-z set, with only 8/47 seen by UVOT.
We also consider whether the fraction of observable bursts,
F (z) = fz fSwif t,
is somehow increasing with z. While
fSwif t is difficult to quantify, our selection criteria disfavor
incompleteness of our sample at low z. We focus upon the
probability of determining a redshift for a given GRB, which
we subdivide as fz = fE/A fAG fhuman. Perhaps the most ob-
vious influence on this term is the fact that at low z, most red-
shifts are determined by observing emission lines (from the
host galaxy); while at higher redshifts, nearly every redshift is
found through absorption lines in the afterglow spectrum (see
Fig. 2 and the Swift archive). While this emission/absorption
bias, fE/A, is not easy to quantify, it should not cause such
an evolutionary effect, since most redshifts in our sample are
found through absorption.
The observability of an optical afterglow, fAG, might be ex-
pected to be steeply falling with redshift; however, for a spec-
trum ∝ t−α ν−β (Sari et al. 1998), cosmological redshifting
may allow the earliest (brightest) portion of the afterglow to
be more visible (Ciardi & Loeb 2000). If the flatness of the
obscuration-corrected SFH at moderate z arose from a steeply
increasing uncorrected rate and a deceasing dust correction,
the detectable afterglow fraction might increase with redshift;
this does not appear to be the case (Schiminovich et al. 2005).
A paucity of IR detections of hosts of "dark" GRBs also ar-
gues against a significant obscured fraction (Le Floc'h et al.
2006). The decision of which telescopes are made available
to observe GRBs, along with other such non-intrinsic proper-
ties, can be folded into fhuman. None of these terms appear
to be able to increase overall observability with z, disfavoring
an origin of the trend in F (z).
6. POTENTIAL SOURCES OF EVOLUTION
We now investigate whether an evolving E(z) can explain
the observed evolution. While it is now generally accepted
that long GRBs arise from massive stars, the special con-
ditions that are required for such an event are still in ques-
tion. In the collapsar model (MacFadyen & Woosley 1999),
the collapse of a rapidly-rotating, massive stellar core to a
black hole powers a jet that is seen as a GRB. Since ev-
ery observed supernova coincident with a GRB has been of
Type Ic (Woosley & Bloom 2006), the progenitor star should
also have lost its outer envelope (without losing precious
angular momentum). Rather strong observational evidence
now indicates that GRB host galaxies tend to be faint and
metal-poor (e.g., Fruchter et al. (2006); Stanek et al. (2006);
Le Floc'h et al. (2003); Fynbo et al. (2003)), increasing inter-
est in models that use single, low-metallicity stars as a path-
way to a collapsar (Yoon & Langer 2005; Woosley & Heger
2006). Decreasing cosmic metallicity may cause the GRB/SN
ratio to rise with z; e.g., the prediction of Yoon et al. (2006)
can be estimated as (1+z)1.4 (Yuksel & Kistler 2007). While
a preference for low-metallicity environments may be the sim-
plest explanation, absorption line metallicity studies remain
inconclusive (Savaglio 2006; Prochaska et al. 2007). One
may therefore wonder whether it is possible to concoct an evo-
lutionary scenario without direct progenitor metallicity de-
pendence.
are instead produced in binary systems
(2004);
(1999); Podsiadlowski et al.
Fryer et al.
If GRBs
(e.g.,
4
KISTLER, Y UKSEL, BEACOM & STANEK
(2006)),
to occur
some other mechanism might
Dale & Davies
Since most GRBs
lead to the appearance of evolution.
appear
in star clusters (Fruchter et al. 2006),
where the fraction of massive stars in binaries may be
high (Kobulnicky & Fryer 2007), such channels could be
important. For example, the merger of two & 15 M⊙ stars
with M1/M2 & 0.95 (i.e., "twins", which may be common
(Pinsonneault & Stanek 2006)) in close orbits (r . few AU)
can lead to a more-massive, rapidly-rotating core lacking an
envelope (Fryer & Heger 2005).
In a dense cluster, close
binaries tend to end up closer due to gravitational scatterings
with interloping stars (Heggie's Law; Heggie (1975)), with
a scattering timescale of . 10 Myr for a stellar density of
ρ ∼ 106 M⊙ pc−3 (Hut et al. 1992). An increased rate
of "interloper-catalyzed" binary mergers (ICBMs) could
result from a larger fraction of star formation occurring in
such environments at higher z, and seen as an enhancement
in the GRB rate.
Such a speculative scenario presents
a dynamical source of evolution (instead of altering the
microphysics) and, as these rates are ∝(cid:10)ρ2(cid:11), possibly allow
for examination of the "density contrast" of star formation.
It is interesting that GRBs were discovered in searches for
gamma rays from explosions related to ICBMs of a different
sort (Klebesadel et al. 1973).
An evolving IMF, becoming increasingly top-heavy at
larger z, would increase the relative number of massive stars
produced. Since star formation measurements are primarily
based on radiation from such stars, this alone may not lead to
apparent evolution, unless the very massive end (& 25M⊙)
changed significantly. Any evidence of evolution in the IMF
(e.g., Wilkins et al. (MNRAS, submitted)) provides motiva-
tion for considering such a change. However, this need only
occur in those galaxies that host gamma-ray bursts. A sim-
ilar effect could result from a larger high-z population of
small galaxies resembling low-z GRB host galaxies, seen as
evolution in the galaxy luminosity function (e.g., Ryan et al.
(2007)). An additional possibility is that, if the massive stellar
binary fraction decreases with declining metallicity, as might
be evidenced in the LMC (Mazeh et al. 2006), the number of
potential single-star progenitors may increase.
In conclusion, the set of Swift gamma-ray bursts now al-
lows for model-independent tests of the connection between
the GRB and star formation rates. We present quantitative
evidence that the GRB rate does not simply track star for-
mation over a broad range in redshift; some mechanism, of
a presently-unknown nature, is leading to an enhancement in
the observed rate of high-z gamma-ray bursts. The effects
of stellar metallicity appear to be a compelling explanation,
but cannot be proven yet. Additional observations will be the
only way to discern the root cause of this effect, and allow for
proper understanding of the underlying astrophysics.
We thank Andrew Hopkins for helpful discussions and Nat
Butler, Charles Dermer, and the referee, Virginia Trimble, for
useful comments. We acknowledge use of the Swift public
data archive. MDK is supported by the Department of Energy
grant DE-FG02-91ER40690; HY and JFB by the National
Science Foundation under CAREER Grant PHY-0547102;
and all by Ohio State University and CCAPP.
REFERENCES
Band, D. L. 2006, ApJ, 644, 378
Berger, E. et al. 2005, ApJ, 634, 501
Blain, A. W. & Natarajan, P. 2000, MNRAS, 312, L35
Bloom, J. S. 2003, AJ, 125, 2865
Bloom, J. S., Frail, D. A. & Kulkarni, S. R. 2003, ApJ, 594, 674
Bromm, V. & Loeb, A. 2006, ApJ, 642, 382
Butler, N. R. et al. 2007, ApJ, in press (arXiv:0706.1275)
Chary, R., Berger, E. & Cowie, L. 2007, ApJ, in press (arXiv:0708.2440)
Ciardi, B. & Loeb, A. 2000, ApJ, 540, 687
Daigne, F., Rossi, E. M. & Mochkovitch, R. 2006, MNRAS, 372, 1034
Dale, J. E. & Davies, M. B. 2006, MNRAS, 366, 1424
Efron, B. & Petrosian, V. 1992, ApJ, 399, 345
Fiore, F. et al. 2007, A&A, 470, 515
Fruchter, A. S. et al. 2006, Nature, 441, 463
Fryer, C. L. & Heger, A. 2005, ApJ, 623, 302
Fryer, C. L., Woosley, S. E. & Hartmann, D. H. 1999, ApJ, 526, 152
Fynbo, J. P. U. et al. 2003, A&A, 406, L63
Gehrels, N. et al. 2004, ApJ, 611, 1005
Guetta, D., Piran, T. & Waxman, E. 2005, ApJ, 619, 412
Guetta, D. & Piran, T. 2007, JCAP, 7, 3
Heggie, D. C. 1975, MNRAS, 173, 729
Hjorth, J. et al. 2003, Nature, 423, 847
Hopkins, A. M. 2004, ApJ, 615, 209
Hopkins, A. M. & Beacom, J. F. 2006, ApJ, 651, 142
Hut, P. et al. 1992, PASP, 104, 981
Jakobsson, P. et al. 2006, A&A, 447, 897
Klebesadel, R., Strong, I. & Olson, R. 1973, ApJ, 182, L85
Kobulnicky, H. A. & Fryer, C. L. 2007, ApJ, in press (astro-ph/0605069)
Kocevski, D. & Butler, N. 2007, ApJ, submitted (arXiv:0707.4478)
Lamb, D. Q. & Reichart, D. E. 2000, ApJ, 536, 1
Le, T. & Dermer, C. D. 2007, ApJ, 661, 394
Le Floc'h, E. et al., 2003, A&A, 400, 499
Le Floc'h, E. et al., 2006, ApJ, 642, 636
Liang, E. et al., 2007, ApJ, 662, 1111
MacFadyen, A. & Woosley, S. E. 1999, ApJ, 524, 262
Mazeh, T., Tamuz, O., & North, P. 2006, MNRAS, 367, 1531
Meszaros, P. 2006, Rept. Prog. Phys. , 69, 2259
Natarajan, P. et al. 2005, MNRAS, 364, L8
Nava, L. et al. 2007, MNRAS, 377, 1464
Panaitescu, A. 2007, MNRAS, 380, 374
Pinsonneault, M. H. & Stanek, K. Z. 2006, ApJ, 639, L67
Podsiadlowski, Ph. et al. 2004, ApJ, 607, L17
Porciani, C. & Madau, P. 2001, ApJ, 548, 522
Prochaska, J. X. et al. 2007, ApJ, 666, 267
Ryan, R. E., Jr. et al. 2007, ApJ, 668, 839
Sakamoto, T. et al. 2007, ApJS, in press (arXiv:0707.4626)
Salvaterra, R. & Chincarini, G. 2007, ApJ, 656, L49
Sari, R., Piran, T. & Narayan, R. 1998, ApJ, 497, L17
Savaglio, S. 2006, New J. Phys. , 8, 195
Schiminovich, D. et al. 2005, ApJ, 619, L47
Stanek, K. Z. et al. 2003, ApJ, 591, L17
Stanek, K. Z. et al. 2006, Acta Astron., 56, 333
Totani, T. 1997, ApJ, 486, L71
Wijers, R. A. M. et al. 1998, MNRAS, 294, L13
Wilkins, S., Trentham, N. & Hopkins, A. 2007, MNRAS, submitted
Woosley, S. & Bloom, J. 2006, ARA&A, 44, 507
Woosley, S. & Heger, A. 2006, ApJ, 637, 914
Yoon, S. C. & Langer, N. 2005, A&A, 443, 643
Yoon, S. C., Langer, N. & Norman, C. 2006, A&A, 460, 199
Yuksel, H. & Kistler, M. D. 2007, Phys. Rev. D, 75, 083004
|
astro-ph/9601114 | 1 | 9601 | 1996-01-22T17:03:45 | Genus statistics for structure formation with topological defects | [
"astro-ph"
] | We study the efficiency of genus statistics in differentiating between different models of structure formation. Simple models which reproduce the salient features of the structure seeded by topological defects are examined. We consider accretion onto static point masses, modeling slow-moving cosmic string loops or other primordial point-like sources. Filamentary structures and wakes are considered as models of the structures seeded by slow and fast moving string, respectively. Comparison is made with predictions of genus statistics for Gaussian fluctuations and with genus curves obtained by the CfA survey. A generic class of density models with wakes and filaments is found to provide results comparable or better than Gaussian models for this suite of tests. | astro-ph | astro-ph | Genus statistics for structure formation
with topological defects
P. P. Avelino
Department of Applied Mathematics
and Theoretical Physics
University of Cambridge
Silver Street, Cambridge CB EW *
Abstract
6
9
9
1
n
a
J
2
2
4
1
1
1
0
6
9
/
h
p
-
o
r
t
s
a
The main ob jective of this paper is to study the e(cid:14)ciency of genus statistics in di(cid:11)er-
entiating between di(cid:11)erent models of structure formation. Some of the models studied
are over-simpli(cid:12)ed but aim at reproducing some of the features of the structure seeded
by topological defects. We consider accretion onto static point masses that could ap-
proximate accretion onto slow-moving cosmic string loops or other primordial point-like
sources. Filamentary structures and wakes are also considered as an approximation to
the structures seeded by slow-and-fast-moving long wiggly strings. Comparisons are
also made with predictions of genus statistics for Gaussian (cid:13)uctuations and with genus
curves obtained from the CfA survey. A generic class of density models with wakes and
(cid:12)laments was seen to provide results comparable or better than Gaussian models for
this suite of tests.
Sub ject headings: cosmology: large-scale structure of the universe: cosmic strings
Introduction
There are several models that aim at explaining the large scale structure we observe in
the universe today. The two main paradigms for the origin of the seed perturbations are,
on one hand, the in(cid:13)ationary scenario that produces Gaussian (cid:13)uctuations and on the
* Email: ppa @ amtp.cam.ac.uk
other hand, topological defects produced at phase transitions in the early universe that
produce a non-Gaussian spectrum of density perturbations, specially on small scales.
While, in general, in(cid:13)ationary scenarios require great (cid:12)ne tuning of parameters in or-
der for the required perturbation amplitude to be produced by the present time, models
where the structure is seeded by topological defects depend mainly on a single parameter
which usually is the energy scale of the phase transition. It turns out that for cosmic
strings, global monopoles and textures the energy scale is that of GUT phase transi-
tions. In a way these models seem to explain the origin of the seed perturbations more
naturally than in(cid:13)ationary models and so they deserve close attention. At present most
work is done on in(cid:13)ationary models and, because they produce Gaussian (cid:13)uctuations,
all statistics are completely described by the power spectrum; other statistics that could
help to distinguish between in(cid:13)ationary and non-in(cid:13)ationary models are often neglected.
Because both models produce a nearly scale-invariant spectrum of (cid:13)uctuations and be-
cause the uncertainties in the measurements are still large, the power spectrum is not
the best statistic to distinguish between them. So, a higher order statistic is required.
In this article we investigate the e(cid:14)ciency of genus statistics in distinguishing between
di(cid:11)erent models of structure formation such as in(cid:13)ationary and cosmic string models
and compare with observational genus curves obtained from the CfA survey. Some work
on genus curves of isodensity contours for toy models of structure formation seeded by
topological defects was previously done by Brandenberger, Kaplan & Ramsey ( ).
By requiring that all the structures have the same size and mass they appear not to have
properly taken into account the scaling solution. The size of the defect seeds increases
with time (proportionally to the horizon) and the amplitude of the density perturbations
induced by larger defects is smaller because they have less time to grow by gravitational
instability. This e(cid:11)ect is properly taken into account in the present paper and so we
consider our work an improvement over that of Brandenberger et al. ( ). Robinson
& Albrecht ( ) also performed a similar study with cosmic string wakes. Their cosmic
string toy model consisted of a realization of a string power spectrum, where the phases
of the Fourier modes were choosen at random, plus a single cosmic string wake. They
concluded that the genus statistic is not a good discriminator between their model and a
model without the wake included. As we will show in this paper the genus statistic is a
good discriminator between our cosmic string toy models and random phase models with
the same power spectrum at least in the linear regime. So, we conclude that their model
underestimates the presence of sheet-like features in the density (cid:12)eld and, therefore, in
this regard it should not be considered a good cosmic string toy model.
This article is organized as follows. We start in section by introducing the
Zel'dovich aproximation, solving it for accretion onto static point masses, (cid:12)laments and
wakes assuming the dark matter to be cold. In the end of the section a modi(cid:12)cation
to the Zel'dovich aproximation that accounts for the neutrino free streaming length is
introduced when considering a hot dark matter model. In section the genus statistic
is described. Analytic results for Gaussian perturbations are given and topological mea-
sures of departures from Gaussianity are introduced. In section we describe the way
we produce the (cid:13)uctuations for the models considered and relate that to the results of
section . In section the results are presented. The dependence of the genus curves on
the type and number of topological defects present is demonstrated and error bars due
to sample variance are introduced. The density probability distribution for several of
the models studied in this article are given and the parametrization of the genus curve
discussed. The (cid:12)nal section is a discussion of the results.
In this article we shall assume that (cid:10) = and h = H
=( Km=sec=Mpc).
The Zel'dovich approximation
. Cold dark matter
We employ the Zel'dovich approximation (Zel'dovich ) to examine each model we
use. We will assume the universe to be (cid:13)at, with no cosmological constant and the dark
matter to be cold. In the next section we will see how the Zel'dovich approximation can
be modi(cid:12)ed in order to describe hot dark matter as well. The equations for the evolution
of the scale factor are the Friedman and Raychaudhuri equation that in the matter are
(cid:18)
(cid:19)
_a
(cid:25)G(cid:26)
b
=
;
()
a
(cid:127)a
(cid:25)G(cid:26)
b
= (cid:0)
;
()
a
where (cid:26)
is the background density. The position of the CDM particles is given by
b
r = a(t)[q +
(q; t)];
()
~
where q is the unperturbed comoving position of the particle and
is the comoving
~
displacement vector of the particle. In the presence of a perturbing string seed, the cold
dark matter particle will obey
d
r
dt
= F
+ F
;
()
seed
matter
where the force F
due to the surrounding matter is given by
matter
F
= (cid:0)r(cid:8);
()
matter
with the gravitational potential (cid:8) satisfying the Poisson equation
r
(cid:8) = (cid:25)G(cid:26)
= (cid:25)G(cid:26)
( + (cid:14) ):
()
matter
b
The fact that mass is conserved is described by the continuity equation that can be
written in comoving coordinates to (cid:12)rst order in
as
~
@ ((cid:26)a
)
_
~
+ r
(cid:1) ((cid:26)a
) = ;
()
@ t
q
which can be integrated in the limit of small perturbations to give
@ (cid:26)
_
~
(cid:14) =
(cid:25) (cid:0)r
(cid:1)
;
()
(cid:26)
q
again to (cid:12)rst order in
. The Poisson equation can then be integrated to give
~
(cid:25)G(cid:26)
b
~
r
(cid:8) =
(r (cid:0) a
);
( )
r
k
where
and
are de(cid:12)ned by
k
?
~
~
~
~
~
=
+
;
( )
k
?
r
(cid:2)
= ;
()
r
k
~
~
r
(cid:1)
= :
()
r
?
In the linear regime with (j
j << jqj) we can use (), () and ( ), to write the Zel'dovich
~
approximation.
(cid:18)
(cid:19)
@
_a
@
(cid:127)a
~
a
+
+
= F
;
()
@ t
a
@ t
a
k
seedk
(cid:18)
(cid:19)
@
_a
@
~
a
+
= F
:
()
@ t
a
@ t
?
seed?
If the perturbations are irrotational to begin with, and if the source term is irrotational
~
~
~
= so that
=
. Then we can write the Zel'dovich approximation as
?
k
(cid:18)
(cid:19)
@
_a
@
(cid:127)a
~
+
+
=
F
= S
:
()
seed
seed
@ t
a
@ t
a
a
If the seed is a point mass S
is given by
seed
S
= (cid:0)
;
()
pmass
GM (q (cid:0) q
)
s
jq (cid:0) q
j
a
s
where q
is the comoving position of the point mass. If the seed is a line of mass then
s
S
is given by
seed
S
= (cid:0)
;
()
lmass
GM
(q (cid:0) q
)
L
s
jq (cid:0) q
j
a
s
where q
is the comoving position of the point in the line of mass nearer to q and M
s
L
is the mass per unit length of the line of mass. The Zel'dovich approximation can be
solved using the Green's function method. For the case we are considering the Green's
function is
!
(cid:18)
(cid:19)
=
t
t
t
G(t; t
) =
(cid:0)
t > t
;
()
t
t
G(t; t
) =
t < t
:
( )
Using it we (cid:12)nd the solution of the Zel'dovich equation for the static point mass and line
of mass
!
(cid:18)
(cid:19)
=
~
GM xt
t
t
i
i
(x; t) =
(cid:0)
(cid:0)
;
( )
pmass
jxj
t
t
i
!
(cid:18)
(cid:19)
(cid:18)
(cid:19)
(cid:18)
(cid:19)
=
=
GM
xt
t
t
t
t
L
i
i
~
(x; t) = (cid:0)
ln
(cid:0)
+
:
()
(cid:12)lament
jxj
t
t
t
t
i
i
i
The formation of wakes can be modeled given an initial velocity of the form
_
~
(x; t
) = u
(cid:15)(q);
()
i
i
where
(cid:15)(q) = (cid:0);
x (cid:1) q > ;
()
(cid:15)(q) = ;
x (cid:1) q < :
()
This is given to the particles in cosmic string models because of the conical spacetime
with u
= (cid:25)G(cid:22)v
, where v
is the string velocity. The solution of the Zel'dovich equation
i
s
s
in this case is
!
(cid:18)
(cid:19)
=
~
t
t
i
(x; t) =
u
t
(cid:15)(q)
(cid:0)
:
()
i
i
t
t
i
. Hot dark matter
If the dark matter is hot, then small-scale perturbations with wavelength smaller than
the neutrino (or other hot dark matter candidate) free streaming length (cid:21)
will be
F S
erased.
In order to properly describe the formation of structure in the context of a
hot dark matter model one would need to solve the Boltzmann equation for the dark
matter particles. However, it was shown by Perivolaropolous, Brandenberger & Stebbins
( ) in studying the clustering of neutrinos in cosmic string induced wakes, that most
of the results can be described correctly using a naive modi(cid:12)cation of the Zel'dovich
approximation. This modi(cid:12)cation is based on the fact that on average the dark matter
particles will only start to collapse when the comoving free-streaming length has fallen
below jqj. So we modify the Zel'dovich approximation in the context of a hot dark
matter model by setting j
j = on scales jqj < (cid:21)
and evolving scales jqj > (cid:21)
as
F S
F S
~
for the cold dark matter case. We consider a hot dark matter model with two neutrino
species with su(cid:14)cient mass to make (cid:10)
= . In particular we consider these two species
(cid:23)
to have the same mass m
= eV.
(cid:23)
The comoving distance hot particles can move since t (cid:21) t
is
eq
(cid:21)
= v (t)tz (t);
()
F S
where v (t) is the thermal velocity of the hot particles and z (t) is the redshift. The
free streaming length grows in the radiation era attaining a maximum at t
and then
eq
decreases proportionally to t
in the matter era. The maximal free streaming length
(cid:0)=
is
where
(cid:21)
= v
t
z
;
()
F S
eq
eq
eq
v
= T
=m
(cid:25) : ;
()
eq
(cid:23)
(cid:23)
eq
so that the maximum free streaming length is given by (cid:21)
(cid:25) h
Mpc.
F S
eq
(cid:0)
Genus statistics
. De(cid:12)nition of genus
To measure the topology of isodensity contours we will use the Gauss-Bonnet theorem
which relates the integrated Gaussian curvature (a local property) of a surface with the
genus (a global property) of that surface. The Gaussian curvature of a two-dimensional
surface at a particular point is
K =
;
( )
a
a
where a
and a
are the two principal radii of curvature at that point. A surface has a
positive or negative Gaussian curvature respectively if the two radii of curvature point in
the same or in opposite directions. For example, a sphere has a positive radii of curvature
given by K = =r
where r is the radius of the sphere. A cylinder has K = because
one of the radii of curvature is in(cid:12)nite. Saddle points have negative curvature because
a
and a
have opposite signs. The Gauss-Bonnet theorem relates the integral of the
Gaussian curvature over the surface with the genus in the following way
Z
K dA = (cid:25) ( (cid:0) g );
( )
where g is the genus of the surface and dA a surface element. The genus measures the
number of closed curves that may be drawn on a surface without separating it. It can
also be de(cid:12)ned as
g = number of compact regions (cid:0) number of holes + :
()
A curved surface may be approximated by a network of polygonal faces. When we use
such a network to compute the genus we (cid:12)nd that
X
D
= (cid:25) ( (cid:0) g );
()
i
where D
= (cid:25) (cid:0)
V
is the angle de(cid:12)cit at a vertex in radians. In this case the curvature
i
i
P
is e(cid:11)ectively compressed into delta functions at the vertex. We used () to construct
a numerical alghoritm in C in order to determine the genus of an isodensity surface
applying the method sugested by Gott, Mellot & Dickinson ( ). The isodensity
surface is constructed by binning the density onto a cubic lattice and identifying pixels
with density above and below a certain threshold ((cid:14)
). Our program to compute the
c
genus was tested against analytic results for Gaussian perturbations and also against
well known topological con(cid:12)gurations for which we knew the genus beforehand (e.g. a
network of isolated cubes).
. Window function
The data is smoothed with a Gaussian window function
(cid:0)
r
(cid:21)
e
w(r) =
e
;
()
=
(cid:25)
(cid:21)
e
where (cid:21)
is the smoothing scale. This de(cid:12)nition implies that the smoothing length is
e
greater by a factor of
than the usual width of a Gaussian. The smoothing scales
p
are always greater than the average interparticle spacing but not too large as to erase
all the relevant features in the density map. The smoothing scales considered in this
article were , , , , and h
Mpc in order to make a direct comparison with
(cid:0)
the results of Vogeley et al. ( ) for the CfA survey.
. Genus for Gaussian random (cid:12)elds
A comparison with in(cid:13)ationary models predictions for the genus curve is essential if one
has to decide which model describes better the kind of large scale structures we observe
in the universe today. The genus curves of random (cid:12)elds are well studied and some
analytic formulae have been derived (Bardeen et al. ; Hamilton, Gott & Weinberg
). The genus per unit volume is given by
g
((cid:23) ) = N ( (cid:0) (cid:23)
)e
;
()
s
(cid:0)(cid:23)
=
where (cid:23) is the number of standard deviations above or below the mean density contour.
The amplitude N depends on the power spectrum of density (cid:13)uctuations P (k ) as
(cid:18)
(cid:19)
=
hk
i
N =
;
()
(cid:25)
where
R
k
P (k )d
k
R
hk
i =
:
()
k
d
k
If we smooth the structure on a scale (cid:21)
the power spectrum becomes
e
(cid:0)k
(cid:21)
=
e
P
(k ) = P (k )e
;
()
and if the power spectrum is of the form P (k ) / k
then N is given by
n
(cid:18)
(cid:19)
=
+ n
N =
:
()
((cid:25) )
(cid:21)
e
To compute the genus curve we must multiply () by the volume of the grid. In (cid:12)g. we
plotted a random phase genus curve obtained for a P (k ) / k power spectrum obtained
respectivelly analytically, using () and (), and numerically using our program to
compute the genus. The two curves are almost identical as we should expect if the
program to compute the genus is working properly. Small di(cid:11)erences between the two
curves can be atributed mainly to the choice of periodic boundary conditions. Smaller
di(cid:11)erences are also due to the sample variance (the volume of the box is not in(cid:12)nite)
and to small numerical imprecisions. The (cid:12)gure is symmetric with respect to the vertical
axis which puts in evidence the topological equivalence of positive and negative linear
density perturbations for random phase models of structure formation. The shape of the
random phase genus curve is independent of the power spectrum. For j(cid:23) j < the genus
is always positive, the surface has more holes than compact regions and so the surface is
"spongelike". For j(cid:23) j > the genus is negative and the surface has a lot of independent
compact regions. For non-random phase distributions the genus curve will be, in general,
asymmetric because the topological symmetry between high and low density regions is
not expected in most cases.
. Genus metastatistics
To quantify departures from the random phase curve, Vogeley et al. ( ) used genus
related statistics like the amplitude, the width, and the shift of the genus curve. The
amplitude was de(cid:12)ned as the amplitude of the best (cid:12)t random phase curve. That was
done by determining the amplitude of the genus curve that minimizes (cid:31)
. The reason
for the de(cid:12)nition was that they wanted to compare the observational data from the CfA
survey with predictions from Gaussian models; So, it seemed appropriate. Although, we
want to make a comparison of present observations with non-Gaussian models we will
retain the same de(cid:12)nition in order to directly compare our results with those of Vogeley
et al. ( ).
The width of the genus peak W
was de(cid:12)ned as the di(cid:11)erence between the zero
(cid:23)
crossings of the genus curve which for random phases is equal to W
= (cid:23)
(cid:0) (cid:23)
=
(cid:23)
+
(cid:0)
because, as we have seen before the genus for random phases is positive over the range
(cid:0) < (cid:23) < and negative elsewhere. This change of the genus sign is believed to
coincide with the percolation thresholds for random-phase perturbations. In the range
(cid:0) < (cid:23) < both high ((cid:14) > (cid:14)
) and low ((cid:14) < (cid:14)
) density phases percolate, while for
c
c
(cid:23) > only the low density phase percolates and for (cid:23) < (cid:0) only the high density phase
percolates.
The last statistic they used was the shift (cid:1)(cid:23) of the genus curve which was quanti(cid:12)ed
in the following form
R
(cid:0)
(cid:23)G((cid:23) )
d(cid:23)
obs
(cid:1)(cid:23) =
;
( )
R
G((cid:23) )
d(cid:23)
(cid:12)t
(cid:0)
where G((cid:23) )
is the measured genus curve and G((cid:23) )
is the best (cid:12)t random phase cuve.
obs
(cid:12)t
A positive value of (cid:1)(cid:23) indicates a density distribution that is more `meatball-like' (iso-
lated cluster models) than random phase, while a negative value of (cid:1)(cid:23) is characteristic
of a `bubble-like' topology (`swiss chesse' topology).
. Parametrization of the genus curve
For a Gaussian density (cid:12)eld the volume fraction in the high density region is given by
Z
(cid:0)t
=
f ((cid:23) > (cid:23)
) =
e
dt:
( )
p
(cid:25)
(cid:23)
Vogeley et al. ( ) did not compute the mean and standard deviation (cid:23) of the density
distribution and express the genus curve as a function of (cid:23) . Instead they determined
the genus curve as a function of the volume fraction in the high density region and
used ( ) to parametrize f as a function of (cid:23) . Although for Gaussian perturbations
these two ways of calculating the genus curves would give similar results, for some of
the models we investigate in this article the genus curves are quite di(cid:11)erent due to their
non-Gaussianity.
. Smoothing of the genus curve
The genus curves were smoothed using a very simple procedure known as three-point
boxcar smoothing (see for example Vogeley et al. ) that was shown to give better
estimates of the true genus curve for Gaussian random phase models.
It consists in
determining the genus as
G
((cid:23)
) =
(G((cid:23)
) + G((cid:23)
) + G((cid:23)
));
()
i
i(cid:0)
i
i+
where (cid:23)
= (cid:23)
+ :.
i+
i
Models and observations
. Toy models
The toy models we study in this paper are simpli(cid:12)ed cosmic string models.
It is a
well-known fact that the gravitational e(cid:11)ect of a slow-moving small loop can be well
approximated by that of a static point mass, thus generating spherical accretion (see for
example Vilenkin & Shellard ). Also a slow-moving wiggly string can be approx-
imated by a line of mass, thus generating (cid:12)lamentary structures, while a fast moving
string generates a wake. It is also known that cosmic string networks after the friction
dominated era rapidly attain a scaling solution where the average properties of the net-
work (such as the average number of defects and the average correlation length) remain
the same at all times when scaled to the horizon size (Bennett & Bouchet ; Allen
& Shellard , Albrecht & Turok ). This means that although (cid:13)uctuations laid
down at later times are smaller in amplitude because they have less time to grow by
gravitational instability, they will have a larger wavelength (in proportion to the horizon
size). These facts are essential ingredients of the toy models we consider in this paper.
Here we consider several toy models of cosmic string structure formation without
taking into consideration the detailed properties of cosmic string networks known from
numerical simulations. These models are respectively a network of fast-moving strings,
a network of slow-moving wiggly strings and a network of slow-moving small loops.
Both the (cid:12)lament and wake models should give a rough approximation to the structures
seeded by cosmic strings. Although the network of slow moving small loops cannot be
considered a realistic cosmic string toy model because loops of cosmic string produced
by a cosmic string network are born with relativistic velocities and are in much higher
number than what is considered in this article, there are scenarios which this toy model
does approximate, notably those with loop nucleation during in(cid:13)ation. The kind of
shapes we investigate also appear in other defect-seeded structure formation models like
those seeded by global monopoles or global textures. Although we want primarily to
test if the genus statistic is a good discriminator between di(cid:11)erent models of structure
formation (specially between di(cid:11)erent non-Gaussian models) we also want to see if some
of the features of these simpli(cid:12)ed models match current observations. To properly test
cosmic string models of structure formation one would need to go beyond this simpli(cid:12)ed
models and perform large-scale network simulations (Avelino & Shellard ).
In (cid:12)g we plotted the isodensity contours for some of the CDM toy models con-
sidered. We can see mainly (cid:12)lamentary, wake-like, and spherical structures respectively
in the (cid:12)lament, wake and sphere models. Other kinds of shapes can be obtained
due to superposition of density perturbations generated by several defects. In the sphere
model there are more small spheres than big ones because denser ob jects are generated
later in this model when there are less defects inside the box. In the (cid:12)lament and wake
models we can see more smaller wakes and (cid:12)laments than larger ones. This is due to the
fact that smaller ob jects are seeded earlier when there are more defects inside the box.
These give rise to larger density perturbations and so to thicker ob jects in the density
contour plot.
. The CfA survey
Vogeley et al. ( ) studied the topology of large scale structure in the CfA Redshift
Survey. This survey includes (cid:25) galaxies with limit magnitude m
(cid:20) : and
b
(cid:0)
allowed for the computation of the topology on smoothing scales from to h
Mpc.
They used genus statistics to test several variants of the cold dark matter cosmogony
(CDM). All of them failed to match the observations to a high con(cid:12)dence level (> %)
(Vogeley et al. ), even when evolution of the perturbations into the non-linear regime
through the use of N-body codes was taken into account. This is a good motivation for
our work based on non-Gaussian models of structure formation. Our results for the
cosmic string toy models of structure formation are such that direct comparisons with
the CfA genus curves are possible.
. Generation of (cid:13)uctuations
We apply the genus statistics to smoothing scales between h
Mpc and h
Mpc
(cid:0)
(cid:0)
that are in the linear or mildly non-linear regime by the present time. In this paper we
will only consider linear theory, in the form of the Zel'dovich approximation, to evolve
the perturbations in the matter era. Matter accretion during the radiation era was not
considered in the present paper. For linear perturbations the genus curve is isomorphic to
the `initial' genus curve which implies that the genus curve as a function of the number
of standard deviations from mean density does not change with time if no additional
defects enter the box. The e(cid:11)ect of non-linear evolution on the genus curves is work in
progress at the present moment (Avelino & Canavezes ). However, we expect this
to be small on scales greater than h
Mpc. We produce the density (cid:13)uctuations in
(cid:0)
a comoving box whose size is chosen so that the genus curves obtained can be directly
compared with those obtained from the CfA survey. We (cid:12)x the scaling solution by (cid:12)xing
the length and number of defects of each type per horizon volume. These were put in the
box with random positions and orientations. The number of defects per horizon volume
p
is N (cid:6)
N and the wavelength of the perturbations induced by them to be initially (at
t
) (cid:21)
and
(cid:21)
(cid:2)
(cid:21)
for the (cid:12)laments and wakes respectively ((cid:21)
= ct
). Structures
eq
eq
eq
eq
eq
eq
seeded at later times will have a larger wavelength due to the scaling solution. These
structures were produced by displacing the particles according to the symmetry point,
axis and plane and with a dependence on the spatial coordinate given by each of the
solutions of the Zel'dovich equation. In appendix we give a more detailed description
of how the (cid:13)uctuations are produced for each toy model. In this article we assume the
biasing parameter does not depend on the scale so that we can directly compare the genus
curves obtained from the CfA survey with those obtained for the toy models considered
in this paper (at least for scales which are in the linear regime) which deal only with the
distribution of the dark matter .
. Compensation
Arbitrary energy momentum perturbations are not possible in a Robertson-Walker space-
time (Trashen ; Trashen, Turok & Brandenberger ; Veeraghavan & Stebbins
). When the strings are formed in the early universe the excess energy and linear
momentum carried by the string is compensated by an equal de(cid:12)cit in the background
radiation. The fact that energy and momentum must be conserved is well expressed by
the following integral constraints
Z
V
(cid:14)(cid:26)dV (cid:25) ;
()
Z
V
(cid:14)(cid:26)xdV (cid:25) ;
()
where V is a volume much greater than the horizon volume. Because of the way in which
we generate the perturbations, only positive density perturbations would be allowed if
compensation was not taken into account. Although it is di(cid:14)cult to mimic the detailed
e(cid:11)ects of compensation in the genus curve we can at least, as a (cid:12)rst approximation, cor-
rect the genus curve for the average shift of the peak due to not including compensation
by doing
(cid:14)
= (cid:14)
(cid:0) (cid:14)
;
()
new
old
c
such that h(cid:14)
i = inside the box.
new
Results
. Toy models and CDM
In (cid:12)g we show a comparison of the density probability distribution for of the models
studied in the context of cold dark matter, averaged over simulations. These were the
(cid:12)lament model, the wake model and the sphere model.
We can see from the pictures that for (cid:21)
= h
Mpc the density distribution is very
e
(cid:0)
non-Gaussian and that it will approach Gaussianity as we go to larger smoothing scales.
We can also see that the largest departures from Gaussianity occur for the (cid:12)lament
model while for wakes the density probability distribution is very nearly Gaussian even
at small scales. If all the models had the same number of ob jects per unit volume we
would expect the sphere model to exhibit the largest departures from Gaussianity, while
the wake model should be the most Gaussian. The reason for this is that while a point-
like overdensity can only be inside or outside a chosen volume, line-like perturbations
and planar perturbations can be partially inside several volumes (a planar perturbation
with size l (cid:2) l being able to touch more volumes than a line-like perturbation of size
l). Consequently, the number of pieces contained in a chosen smoothing volume will be
maximized for the wake model, making this model more Gaussian, and minimized for
sphere model making it less Gaussian. The reason why our particular sphere model is
more Gaussian than the (cid:12)lament model is that there are more ob jects per unit volume
in the sphere toy model than in the (cid:12)lament toy model.
The large departure from Gaussianity, especially on small scales, makes the genus
curve very sensitive to the parametrization so that di(cid:11)erent genus curves are obtained if
the genus is expressed directly as a function of the number of standard deviations from
mean density or if we use the volume fraction parametrization for (cid:23) . A comparison of
two genus curves for the (cid:12)lament model is shown in (cid:12)g in order to illustrate the e(cid:11)ect
of the parametrization by volume fraction. The genus curves are considerably di(cid:11)erent
(the one parametrized by volume fraction being more like random phase), but as we have
seen this e(cid:11)ect should be most exaggerated for the (cid:12)lament model at the smallest scale
considered ((cid:21)
= h
Mpc). For other toy models and smoothing scales the dependence
e
(cid:0)
of the genus curve on the parametrization will not be so considerable. However, for small
smoothing lengths some information about the density probability distribution is lost. In
this paper we shall use the volume fraction parametrization (unless it is said otherwise)
used by Vogeley et al. ( ) when analysing the CfA observational data.
In (cid:12)g and (cid:12)g we show a comparison of the genus curves, for our CDM models of
structure formation, with the predictions from the standard in(cid:13)ationary CDM model, as
well as the observations taken from the CfA survey. Error-bars due to sample variance
were not included for the sake of clarity. The sizes of the boxes for the toy models
studied in this article were chosen such that a direct comparison between their genus
curves and CfA genus curves is possible for each of the smoothing scales considered. It is
readily apparent from the graphs that from the models studied the one that best (cid:12)ts the
observational data, at least from a topological point of view, is the wake toy model.
The amplitude of the genus curves seems to be only marginally larger than the CfA genus
curves. Also, the width of the genus curves seems to mimic observations much better
than the standard CDM model. The wake model also does marginally better than
standard CDM. The amplitude of the genus curves is always smaller than that predicted
by the standard CDM scenario and in better agreement with observations.
The standard CDM model gives too large amplitude of the genus curve, especially
on small scales and it fails to match other features of the observed genus curves. For
example, the standard CDM model gives W
(cid:25) and although sample variance allows
(cid:23)
some (cid:13)uctuations around this value it is not enough to explain genus peak widths as
large as . or . with a very large con(cid:12)dence level (> %) (Vogeley et al. ). It
is possible to have other random phase models with a smaller amplitude of the genus
curve, as in open models or models with a non-zero cosmological constant. What hap-
pens in this case is that the growth of density perturbations in the linear regime slows
down while the non-linear growth of perturbations proceeds in the normal way. So, for
the same normalization of the spectrum of density perturbations non-linear perturba-
tions are more non-linear in these models. Non-linearities introduce correlations between
phases of di(cid:11)erent Fourier modes, decreasing the number of independent structures and
consequently reducing the genus amplitude. However, as found by Vogeley et al. ( )
the problem of matching the other statistics, especially the width of the genus peak,
remains unsolved.
The and (cid:12)lament models do better than spheres and marginally worse than
standard CDM. The genus amplitude is higher than for standard CDM. However, these
models provide a better (cid:12)tting to the width of the observed genus curves than the stan-
dard CDM in(cid:13)ationary scenario. The sphere model is clearly ruled out. The amplitude
of the genus curves is too large and it fails to match the shape of the observational genus
curves.
Also apparent from the graphs is that for most of the smoothing scales considered,
the genus amplitude is an increasing function of the number of defects. This should be
expected because we are increasing the number of structures present inside the box.
In (cid:12)g we plotted the genus curves obtained for the best model (the wake model)
now with the error bars properly included. The line represents the average genus curve
among realizations of the model. The error bars are one sigma error bars over these
realizations. The asymmetry of the genus curves is visible, as we would expect from
almost any non-Gaussian model of structure formation. However, the parametrization
by volume fraction makes the genus curves look less asymmetric, especially on small
scales, making the curves more like random-phase models.
In (cid:12)g we show a statistical comparison of the genus curves for the wake model with
the predictions from the CfA survey. The genus amplitude for this model is marginally
larger than the observed genus amplitude for some the scales considered, although it does
much better than standard CDM at matching the amplitude of the CfA genus curves.
Again, the error bars we see in the plots are one sigma error bars of simulations.
The shift of the genus curves does not seem to be in agreement with observations for
most of the scales considered. The width of the genus peak seems to be consistent with
observations from the CfA survey for all of the scales considered but in this case the
error bars due to sample variance are very large so that a very wide range of genus peak
widths are possible. This toy model (cid:12)ts the observations better than any random phase
model tried by Vogeley et al. ( ).
. Toymodels and HDM
In (cid:12)g we show a comparison of the density probability distribution for of the models
studied as a function of the smoothing length and averaged over simulations. These
were the (cid:12)lament model, the wake model and the sphere model with hot dark
matter. These models are more Gaussian than the corresponding models with CDM and
consequently the dependence of the genus curves on the parametrization is smaller in
this case. For the wake model the density probability distribution is nearly Gaussian
even for (cid:21)
= h
Mpc.
e
(cid:0)
In (cid:12)g and (cid:12)g we show a comparison of the genus curves for models of structure
formation we studied in the context of hot dark matter with the predictions from the
standard CDM model and observations taken from the CfA survey. The amplitude of the
genus curves is smaller with HDM than with CDM, especially on small scales, because
an adittional smoothing was introduced due to the free-streaming of the neutrinos. The
models that best mimic the observational results are again the wake models (particularly
the wake model).
In (cid:12)g we can see the genus curves obtained for the best model (the wake model)
now with the error bars properly included. These seem to be more like random-phase
than the genus curves obtained for the wake model. The line represents the average
genus curve among realizations of the model and the error bars are one sigma error
bars.
Figure shows a statistical comparison of the genus curves for the wake model
with the predictions from the CfA survey. The genus amplitude for this model seems to
be consistent with the observed genus amplitude for most of the scales considered. Again,
the error bars we see in the plots are one sigma error bars of simulations. The shift
of the genus curves does not seem to be in agreement with observations for most of the
scales considered. The width of the genus peak seems to be consistent with observations
from the CfA survey although in this case the error bars due to sample variance are very
large. This toy model also (cid:12)ts the observations better than any random phase model
tried by Vogeley et al. ( ).
Discussion
We can conclude from the results presented in this article that the genus statistic is a
good discriminator between Gaussian and non-Gaussian models of structure formation.
In addition to this, it was shown to be a good statistic to distinguish between di(cid:11)erent
toy models of structure formation, sensitive to the shape, number of structures seeded,
and dark matter type. We also showed that at least for some of the models considered (in
particular the wake models) that there is a better agreement with the observations than
that veri(cid:12)ed for random phase models of structure formation. It is necessary to use string
network simulations in combination with numerical codes that generate and evolve the
density perturbations seeded by such networks in order to properly test the cosmic string
model for structure formation. However, we have shown that there are some features on
the observed genus curves (e.g. genus peak width considerably larger than and genus
amplitude smaller than standard CDM), that cannot be easily matched by random phase
models of structure formation, but which are matched by some toy models considered
in this paper over a range of smoothing lengths. Although this cannot be considered
to be a serious quantitative test of the cosmic string paradigm for structure formation,
it provides a good motivation for arguing that some statistical features of topological
defect models of structure formation are better than random phase models predicted by
most in(cid:13)ationary scenarios.
Acknowledgements
PPA is supported by JNICT. I thank my supervisor E.P.S. Shellard and Robert Caldwell
for helpful comments on this manuscript.
Appendix : Generation of perturbations for the toy models
In the sphere model a number N (cid:6)
N of spheres were put, per horizon per Hubble
p
time, at random positions in the box. For each sphere we chose a position p at random.
Particles were then displaced according to the Zel'dovich approximation. Consider a
sphere laid down at an instant t. A particle at a position p
would move from that time
t till the present time t
a comoving distance given (in linear theory) by
~
x
=
GM t
b
(t);
()
i
s
jxj
where x = p
(cid:0) p and
!
(cid:18)
(cid:19)
=
t
t
b
(t) =
(cid:0)
(cid:0)
:
()
s
t
t
If we account for the growth of the mass of the loops chopped o(cid:11) by the network of
cosmic strings (M / t) we have that the quotient of the perturbations laid down at two
di(cid:11)erent times t
and t
at the same position in space is given by
~
(t
)
t
(cid:2) b
(t
)
s
=
:
()
~
(t
)
t
(cid:2) b
(t
)
s
We can see that in the case of the sphere model (only) perturbations seeded at later
times can have larger amplitude.
In the (cid:12)lament model for each (cid:12)lament a point p and a unit vector v were chosen at
random. Particles were then displaced according to the Zel'dovich approximation. Let
us consider a particle at a position p
and let us de(cid:12)ne the vector y = p
(cid:0) p. Consider
the vector x = y (cid:0) (y (cid:1) v)v and assume that the (cid:12)lament has a size given by S
. If
f
j(y (cid:1) v)vj < :S
then a particle at a position p
would move a comoving distance given
f
by
~
x
= (cid:0)
GM
t
b
(t):
()
L
i
f
jxj
where
!
(cid:18)
(cid:19)
(cid:18)
(cid:19)
(cid:18)
(cid:19)
=
=
t
t
t
t
b
(t) =
ln
(cid:0)
+
:
( )
f
t
t
t
t
If j(y (cid:1) v)vj > :S
the particle would not move at all. The mass per unit length of the
f
cosmic strings is approximately constant over time (M
= const) and so we have that
L
the quotient of the perturbations laid down at two di(cid:11)erent times t
and t
at the same
position in space is given by
~
(t
)
b
(t
)
f
=
:
( )
~
(t
)
b
(t
)
f
Consequently, the perturbations seeded at earlier times will have larger amplitude. The
initial (cid:12)lament comoving size (at t
) was taken to be S
= (cid:21)
(where (cid:21)
= ct
) and
eq
f
eq
eq
eq
its size increases with time proportionally to the horizon so that S
/ t
. So, at later
f
=
times larger structures are formed but those will be less dense.
In the wake model for each (cid:12)lament a point p and a unit vector v were also chosen
at random. Particles were then displaced according to the Zel'dovich approximation.
sin (cid:11)
Let us consider the vector A =
and a vector perpendicular to it B =
cos (cid:11)
.
@
A
@
A
Consider the rotation matrices
M =
cos (cid:18)
sin (cid:18)
;
()
@
A
(cid:0) sin (cid:18)
cos (cid:18)
cos (cid:12)
sin (cid:12)
M =
:
()
@
A
(cid:0) sin (cid:12)
cos (cid:12)
The vector A
= M (cid:1) M (cid:1) A obtained as the result of multiplying the matrices M
and M by A is a new vector given by A
= (sin (cid:12) ; sin (cid:18) cos (cid:12) ; cos (cid:18) cos (cid:12) ). The vector
B
= M (cid:1) M (cid:1) B is perpendicular to A
and is given by
cos (cid:12) sin (cid:11)
@
A
B
=
cos (cid:18) cos (cid:11) (cid:0) sin (cid:18) sin (cid:12) sin (cid:11)
:
()
(cid:0) sin (cid:18) cos (cid:11) (cid:0) cos (cid:18) sin (cid:12) sin (cid:11)
The angles (cid:12) and (cid:18) were chosen sub ject to the constraint v = A
and (cid:11) was choosen at
random in the interval (cid:20) (cid:11) < (cid:25) . We have now two perpendicular vectors A
and B
and we need to (cid:12)nd C
perpendicular to this two such that
C
:A
= ;
()
C
:B
= ;
()
and
jC
j = :
()
To see how a particle at a position p
will move let us consider the vector y = p
(cid:0) p
and the vectors x
= y (cid:1) B
, x
= y (cid:1) C
and x = y (cid:0) (y (cid:1) x
)x
(cid:0) (y (cid:1) x
)x
. Let us
assume the size of the wake to be S
. If jx
j < :S
and jx
j < :S
then a particle
w
w
w
at a position p
would move a comoving distance given by
~
x
= (cid:0)
u
t
b
(t):
()
i
i
w
jxj
where
!
(cid:18)
(cid:19)
=
t
t
i
b
(t) =
(cid:0)
:
()
w
t
t
i
If jx
j > :S
or jx
j > :S
the particle would not move at all. The quotient of the
w
w
perturbations laid down at two di(cid:11)erent times t
and t
at the same position in space is
given by
~
=
(t
)
b
(t
) (cid:2) t
f
=
:
( )
~
(t
)
b
(t
) (cid:2) t
f
=
We can see again that the perturbations seeded at earlier times will have larger amplitude.
The initial wake size was taken to be
(cid:21)
and it grows proportionally to the horizon as
eq
in the (cid:12)lament case.
Appendix : The simulation boxes
As we have said before the size of the simulation boxes was chosen in a way to enable
direct comparison with the results from Vogeley et al. ( ) for the CfA survey. To
ensure that the topology was not dominated by shot noise the smoothing length must
(cid:0)
(cid:0)
(cid:21)
(h
Mpc)
V
(h
Mpc)
N
N
e
survey
res
galaxies
: (cid:2)
: (cid:2)
: (cid:2)
: (cid:2)
: (cid:2)
: (cid:2)
Table : Volume statistics for the CfA survey (note that the total volume and consequently the
number of resolution elements for the toy model simulations is the same as for the CfA survey).
be larger than the average intergalaxy (or interparticle) separation. To determine the
maximum distance appropriate for a given choice of smoothing length Vogeley et al.
( ) found r
such that
max
n(r
) = (cid:21)
:
( )
max
e
(cid:0)
This means that the number of galaxies included is an increasing function of the smooth-
ing length. Table shows the volume of the survey as a function of the smoothing length.
The number of resolution elements, de(cid:12)ned by N
=
, and the number of galax-
res
=
(cid:25)
(cid:21)
e
V
survey
ies included in the topological analysis of the CfA survey are also given as a function
of the smoothing length. The volume of the simulation boxes and so the number of
resolution elements were chosen to be the same as for the CfA survey analysis so that
direct comparison between our results and observations was possible. The smoothing
length was always more than times larger than the average interparticle spacing.
References
. Albrecht, A., & Turok, N. [ ], `Evolution of cosmic string networks,' Phys. Rev.
D , .
. Allen, B., & Shellard, E.P.S. [ ], `Cosmic String EvolutionA Numerical Sim-
ulation,' Phys. Rev. Lett. , .
. Avelino, P.P., & Canavezes, A. [ ], in preparation
. Avelino, P.A., & Shellard, E.P.S. [ ], `Matter accretion by cosmic string loops
and wakes,' Phys. Rev. D, .
. Bardeen, J.M., Bond, J.R., Kaiser, N. & Szalay, A.S. [ ], `The statistics of peaks
of Gaussian random (cid:12)elds,' Ap. J. , .
. Bennett, D.P., & Bouchet, F.R. [ ], `High resolution simulations of Cosmic
String Evolution: Network evolution,' Phys. Rev. D, .
. Brandenberger, R. H., Kaplan, D. M. & Ramsey, S. A. [ ], astro-ph/
. Gott, J.R., Mellot, A. L. & Dickinson, M. [ ], `The Sponge-like Topology of
Large-Scale Structure in the Universe,' Ap. J. , .
. Hamilton, A.J.S., Gott, & Weinberg, D.H. [ ], `The topology of the large-scale
structure of the universe,' Ap. J. , .
. Perivolaropolous, L., Brandenberger, R.H., & Stebbins, A. [ ], `Dissipationless
clustering of neutrinos in cosmic-string-induced wakes,' Phys. Rev. D, .
. Robinson, J., & Albrecht, A. [ ], Paper submited to Mon. Not. R. Astron. Soc.
. Traschen, J. [ ], `Constraints on stress energy perturbations in general relativ-
ity,' Phys. Rev. D, .
. Traschen, J.,Turok, N. & Brandenberger, R.H. [ ], `Microwave anisotropies by
cosmic strings,' Phys. Rev. D, .
. Veeraghavan, S., & Stebbins, A. [ ], `Causal compensated perturbations in cos-
mology,' Ap. J. , .
. Vilenkin, A. & Shellard, E.P.S. [ ], Cosmic Strings and other topological defects
(Cambridge University Press).
. Vogeley, M.S., Park, C., Geller, M.J., Huchra, J.P., & Gott, J.R. [ ], `Topological
Analysis of Cfa Redshift Survey,' Ap. J. , .
. Zel'dovich, Ya.B. [ ], `Gravitational instability: An approximate theory for large
density perturbations,' Astron. Ap. , .
a)
b)
Fig. Comparison of the genus curve obtained for a
simulation of a P (k) / k
power spectrum smoothed on a scale (cid:21)
equal to
times the grid spacing, obtained
e
p
(a) analytically (b) numerically using our program to calculate the genus.
Fig. Isodensity contours with (cid:14) = :(cid:23) for (a) the (cid:12)lament model b) the
wake model c) the sphere model with cold dark matter. The box size is : (cid:2)
(cid:0)
(cid:0)
h
Mpc
, and the smoothing length is (cid:21)
= h
Mpc.
e
a)
b)
c)
Fig. Density probability distribution for several models studied with cold dark
matter as a function of (cid:23) calculated directly from the variance of the density distri-
bution (a) (cid:12)lament model (b) wake model and (c) sphere model.
a)
b)
Fig. Comparison of genus curves for the (cid:12)lament model assuming the dark
matter to be cold. (a) using the volume fraction to prametrize (cid:23) (b) Calculating (cid:23)
directly from the variance of the density distribution.
a)
b)
c)
Fig. Comparison of genus curves for the wake, (cid:12)lament and sphere models
with both standard CDM genus curves and genus curves obtained from the CfA
survey. Error bars are not included for clarity (see (cid:12)g ). It is clear from the picture
that the most favoured model is the wake model.
a)
b)
c)
Fig. Comparison of genus curves for the wakes, (cid:12)laments and spheres
models with both standard CDM genus curves and genus curves obtained from the
CfA survey.
a)
b)
c)
d)
e)
f)
Fig. Genus curves for the wake model. Line represents the average genus curve
among realizations of the model. Error bars are one-sigma.
a)
b)
c)
Fig. Statistical comparison of genus curves obtained from the CfA survey with
genus curves obtained from realizations of the wake model. Error bars on the
wakes model are one-sigma.
a)
b)
c)
Fig. Density probability distribution for several models studied with hot dark
matter as a function of (cid:23) calculated directly from the density distribution (a)
(cid:12)lament model (b) wake model and (c) sphere model.
a)
b)
c)
Fig. Comparison of genus curves for the wake, (cid:12)lament and sphere
models in the context of hot dark matter with both standard CDM genus curves and
genus curves obtained from the CfA survey. Error bars are not included for clarity
(see (cid:12)g ).
a)
b)
c)
Fig. Comparison of genus curves for the wake, (cid:12)lament and sphere
models in the context of hot dark matter with both standard CDM genus curves and
genus curves obtained from the CfA survey. It is clear from the picture that the most
favoured model is the wake model.
a)
b)
c)
d)
e)
f)
Fig. Genus curves for the wake model with HDM. Line represents the average
genus curve among realizations of the model. Error bars are one-sigma.
a)
b)
c)
Fig. Statistical comparison of genus curves obtained from the CfA survey with
genus curves obtained from realizations of the wake model with HDM. Error
bars are one-sigma
|
astro-ph/0401252 | 1 | 0401 | 2004-01-13T21:06:16 | Disk Galaxy Formation in a LambdaCDM Universe | [
"astro-ph"
] | We describe hydrodynamical simulations of galaxy formation in a Lambda cold dark matter (CDM) cosmology performed using a subresolution model for star formation and feedback in a multiphase interstellar medium (ISM). In particular, we demonstrate the formation of a well-resolved disk galaxy. The surface brightness profile of the galaxy is exponential, with a B-band central surface brightness of 21.0 mag arcsec^-2 and a scale-length of R_d = 2.0 h^-1 kpc. We find no evidence for a significant bulge component. The simulated galaxy falls within the I-band Tully-Fisher relation, with an absolute magnitude of I = -21.2 and a peak stellar rotation velocity of V_rot=121.3 km s^-1. While the total specific angular momentum of the stars in the galaxy agrees with observations, the angular momentum in the inner regions appears to be low by a factor of ~2. The star formation rate of the galaxy peaks at ~7 M_sun yr^-1 between redshifts z=2-4, with the mean stellar age decreasing from \~10 Gyrs in the outer regions of the disk to ~7.5 Gyrs in the center, indicating that the disk did not simply form inside-out. The stars exhibit a metallicity gradient from 0.7 Z_sun at the edge of the disk to 1.3 Z_sun in the center. Using a suite of idealized galaxy formation simulations with different models for the ISM, we show that the effective pressure support provided by star formation and feedback in our multiphase model is instrumental in allowing the formation of large, stable disk galaxies. If ISM gas is instead modeled with an isothermal equation of state, or if star formation is suppressed entirely, growing gaseous disks quickly violate the Toomre stability criterion and undergo catastrophic fragmentation. | astro-ph | astro-ph |
DRAFT VERSION OCTOBER 30, 2018
Preprint typeset using LATEX style emulateapj v. 11/26/03
DISK GALAXY FORMATION IN A ΛCDM UNIVERSE
BRANT ROBERTSON1,4, NAOKI YOSHIDA2, VOLKER SPRINGEL3, AND LARS HERNQUIST1
Draft version October 30, 2018
ABSTRACT
We describe hydrodynamical simulations of galaxy formation in a Λ cold dark matter (CDM) cosmology
performed using a subresolution model for star formation and feedback in a multiphase interstellar medium
(ISM). In particular, we demonstrate the formation of a well-resolved disk galaxy. The surface brightness
profile of the galaxy is exponential, with a B-band central surface brightness of 21.0 mag arcsec- 2 and a scale-
length of Rd = 2.0 h- 1kpc. We find no evidence for a significant bulge component. The simulated galaxy falls
within the I-band Tully-Fisher relation, with an absolute magnitude of I = -- 21.2 and a peak stellar rotation
velocity of Vrot = 121.3 km s- 1. While the total specific angular momentum of the stars in the galaxy agrees
with observations, the angular momentum in the inner regions appears to be low by a factor of ∼ 2. The
star formation rate of the galaxy peaks at ∼ 7 M⊙ yr- 1 between redshifts z = 2 - 4, with the mean stellar age
decreasing from ∼ 10 Gyrs in the outer regions of the disk to ∼ 7.5 Gyrs in the center, indicating that the disk
did not simply form inside-out. The stars exhibit a metallicity gradient from 0.7 Z⊙ at the edge of the disk to
1.3 Z⊙ in the center. Using a suite of idealized galaxy formation simulations with different models for the ISM,
we show that the effective pressure support provided by star formation and feedback in our multiphase model
is instrumental in allowing the formation of large, stable disk galaxies. If ISM gas is instead modeled with an
isothermal equation of state, or if star formation is suppressed entirely, growing gaseous disks quickly violate
the Toomre stability criterion and undergo catastrophic fragmentation.
Subject headings: galaxies: evolution -- galaxies: formation -- methods: numerical
1. INTRODUCTION
Numerical simulations have become one of the most impor-
tant theoretical tools for exploring the complicated problem of
galaxy formation. Modern numerical work on galaxy forma-
tion was initiated by Katz & Gunn (1991), Navarro & Benz
(1991), and Katz (1992), who included radiative cooling
by hydrogen and helium, and attempted to account for star
formation and feedback processes. Subsequent work em-
ployed more sophisticated treatments of supernova feed-
back (Navarro & White 1993), models for chemical enrich-
ment (Steinmetz & Müller 1994), and methods to 'zoom
in' on galaxies of interest within cosmological simulations
(Navarro & White 1994)
While these studies achieved limited success in creating ro-
tationally supported disks, the simulated galaxies typically
failed to reproduce their observed counterparts. In general,
the simulated disks were found to be too small and were more
centrally concentrated than actual galaxies (Navarro & White
1993, 1994; Navarro et al. 1995). In addition, the star forma-
tion in the models was overly efficient, converting too large a
fraction of the gas into stars by the present day (Navarro et al.
1995; Steinmetz & Müller 1994, 1995).
Many of the shortcomings of the early modeling can be tied
to the coarse resolution of the simulations and the manner in
which feedback was treated. In a cosmological context, the
multiphase structure of the ISM within galaxies cannot be re-
solved in detail, so star-forming gas is generally described as
a single-phase medium. If feedback energy is deposited into
1 Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cam-
85740 Garching bei München, Germany
4 [email protected]
bridge, MA 02138, USA
181-8588, Japan
2 National Astronomical Observatory Japan, Osawa 2-21-1, Mitaka, Tokyo
3 Max-Planck-Institut
für Astrophysik, Karl-Schwarzschild-Strasse 1,
this gas purely as thermal energy, it will be quickly radiated
away, and any impact of feedback from star formation will
be lost. In addition, if star formation is very efficient in low
mass halos at high redshifts, a large number of dense, compact
galaxies will form and eventually be incorporated into the in-
ner regions of larger galaxies through hierarchical merging.
If these baryonic clumps lose angular momentum to dark ha-
los, the resulting objects will be smaller in extent than if the
baryons had been accreted smoothly. Thus, an early collapse
of baryons opens a channel for gas to lose angular momen-
tum, yielding galaxies that are too compact, lack angular mo-
mentum, and produce stars overly efficiently compared with
observations.
Subsequent work attempted to alleviate these shortcom-
ings by employing stronger forms of feedback and modi-
fications to the cosmological framework.
In some cases,
the problems noted above were exacerbated. For example,
photoheating by a diffuse ultraviolet (UV) background was
found to further reduce the angular momentum content of
the simulated galaxies (Navarro & Steinmetz 1997). Other
efforts included the impact of preheating and gas blow-out
from small halos (Sommer-Larsen et al. 1999) and the forma-
tion of galaxies in a warm dark matter (WDM) cosmology
(Sommer-Larsen & Dolgov 2001). More direct comparisons
to observations were made possible by incorporating spec-
tral synthesis techniques into the modeling (Contardo et al.
1998) and by using the Tully-Fisher relation to constrain the
hierarchical origin of galaxies (Steinmetz & Navarro 1999;
Navarro & Steinmetz 2000). However, the galaxies in these
simulations were still too concentrated.
The most recent studies of disk formation have yielded
somewhat more promising results. Using a model of self-
propagating star formation combined with supernova feed-
back and a UV background, Sommer-Larsen et al. (2002,
2003) produced disk galaxies deficient in angular momentum
by less than an order of magnitude. Governato et al. (2002)
2
Robertson et al.
report the formation of realistic disk galaxies in ΛCDM and
ΛWDM cosmologies using simulations that include standard
prescriptions for cooling, a UV background, and star forma-
tion. Abadi et al. (2003a,b) present detailed analyses of simu-
lated galaxies with kinematic and photometric properties sim-
ilar to observed Sab galaxies.
In addition,
Although the various attempts to simulate disk forma-
tion have provided some impressive successes, the essential
physics behind this process remains unclear. Governato et al.
the low angular momentum of sim-
(2002) claim that
ulated galaxies in previous works owed at
least partly
to inadequate resolution.
they emphasize
the impact that the matter power spectrum can have on
galaxy formation through simulations of WDM universes
in which the bulge and spheroid components of galax-
ies are smaller than in CDM models.
These findings
are supported by Sommer-Larsen et al. (2002, 2003) and
Sommer-Larsen & Dolgov (2001), whose calculations indi-
cate that high resolution, strong stellar feedback, and warm
dark matter can produce realistic disk galaxies. Finally,
Abadi et al. (2003a) suggest that a crucial ingredient for form-
ing proper disks is an implementation of feedback that can
regulate star formation.
Here, we use a 'multiphase model' to describe star-forming
gas, in which the ISM consists of cold clouds in pressure
equilibrium with an ambient hot phase (Springel & Hernquist
2003a). Radiative cooling of gas leads to the growth of
clouds, which in turn host the material that fuels star for-
mation. The supernovae associated with star formation pro-
vide feedback by heating the ambient medium and evapo-
rating cold clouds. The feedback treatment establishes a
self-regulated cycle for star formation, pressurizing the star-
forming gas. Implemented in a subresolution manner, our ap-
proach makes it possible to obtain numerically converged re-
sults for the star formation rate at moderate resolution. This
feature is particularly important for simulations of CDM cos-
mologies, where early generations of galaxies may be difficult
to resolve.
Using this model of star-forming gas, Springel & Hernquist
the his-
(2003b) obtained a converged prediction for
formation that agrees well with
tory of cosmic star
z ≤ 4 (Springel & Hernquist
observations at
2003b; Hernquist & Springel 2003).1
In subsequent work,
Nagamine et al. (2003a,b,d) showed that this model also ac-
counts for the observed abundance and star formation rates of
damped Lyman-alpha absorbers and Lyman-break galaxies at
z ∼ 3.
redshifts
In what follows, we study the consequences of our multi-
phase model for the ISM on the formation of disk galaxies.
First, we employ a high-resolution simulation to identify re-
alistic disk galaxies in a cosmological context. For one such
disk galaxy, we examine its kinematic and photometric prop-
erties in detail, demonstrating good agreement with observa-
tions of local spirals. Second, we use a set of idealized simula-
tions to study disk formation in individual dark matter halos to
isolate the effect of different models for the equation of state
of the ISM. This analysis demonstrates that the feedback in
our multiphase model alters the dynamics by pressurizing the
1 We note an error in figure 12 of Springel & Hernquist (2003b) where the
observational estimates of the star formation rate are plotted too high by a fac-
tor of h- 1 = 1.4. When corrected, the observed points are in better agreement
with the theoretical estimates; see astro-ph/0206395 and Nagamine et al.
(2003c).
star-forming gas and stabilizing forming disks against frag-
mentation. We emphasize that this aspect of our modeling
does not depend on the details of our prescription for star
formation and feedback, but is determined by the effective
equation of state for star-forming gas. Thus, our conclusions
should obtain generally, provided that the actual bulk equa-
tion of state for the ISM has characteristics similar to those of
our description.
In § 2, we present our simulation method. We review our
analysis procedure in § 3, and discuss our findings for the
structural (§ 4) and kinematic properties (§ 5) of one simu-
lated disk galaxy.
In § 6, we describe our idealized simu-
lations and their results. Finally, we conclude and suggest
directions for further research in § 7.
2. COSMOLOGICAL SIMULATION
in
its
Our
(SPH)
"conservative
simulations were performed using the parallel
code
N-body/smoothed particle hydrodynamics
formulation
entropy"
GADGET2
(Springel & Hernquist 2002),
to mitigate problems with
lack of energy and entropy conservation in older treatments
of SPH (e.g. Hernquist 1993; O'Shea et al. 2003). We adopt
a flat ΛCDM cosmology with cosmological parameters
Ωm = 0.3, ΩΛ = 0.7, Ωb = 0.04, and σ8 = 0.9, and set the pri-
mordial power spectrum index to n = 1. Throughout, we select
a value for the Hubble constant of H0 = 100 h km s- 1Mpc- 1
with h = 0.7.
For our cosmological simulation, we populate a periodic
volume of 10 h- 1Mpc on a side with 1443 low-resolution
dark matter (LRDM) particles. At the center of the box, a
5 h- 1Mpc cubic region is selected as a high resolution re-
gion where we replace the LRDM particles with particles
of eight-times lower mass. The initial displacement field is
then calculated following a standard "zooming" procedure
(Tormen 1997; Power et al. 2003), where small scale per-
turbations are added appropriately in the high-resolution re-
gion. Note that the high-resolution zone does not target
a particular object, with the intent of eliminating bias that
could be introduced by selecting halos that may be intrinsi-
cally favorable for disk galaxy formation. We further split
the high-resolution particles into dark matter (HRDM) and
gas. The resulting particle masses of each component are
then mLRDM = 2.79 × 107h- 1M⊙, mHRDM = 3.02 × 106h- 1M⊙,
and mgas = 4.65 × 105h- 1M⊙. We set the gravitational soften-
ing length for the high-resolution particles to 0.65 comoving
h- 1kpc. While our 10 h- 1 Mpc box is too small to be fully
representative of the z = 0 universe, the volume is sufficient
for our purposes as our current work does not concern, for ex-
ample, large-scale correlations of galaxies or the global mass
function. Here, we are interested only in individual, galactic-
sized objects which are mostly unaffected by the simulation
box size.
We include a UV background and radiative cooling and
heating in the manner of Katz et al. (1996) and Davé et al.
(1999), as well as star formation, supernova feedback, and
metal enrichment. We employ the multiphase model devel-
oped by Springel & Hernquist (2003a) to describe the star-
forming gas (see also Yepes et al. 1997; Hultman & Pharasyn
1999). Our approach accounts for some of the key aspects of
the multiphase structure of the ISM (McKee & Ostriker 1977)
without spatially resolving the different phases explicitly. In-
stead, a statistical mixture of the phases is computed analyti-
cally, taking into account the growth of cold clouds embedded
in a supernova-heated ambient phase, the formation of stars
10-8
10-9
10-10
10-11
10-12
10-13
10-14
]
3
-
m
c
g
r
e
[
P
0.01
0.10
1.00
nH [ cm-3 ]
10.00
100.00
FIG. 1. -- Effective equation of state for star-forming gas in our multiphase
model (Springel & Hernquist 2003a). The effective pressure, Peff, is plotted
in cgs units versus the total gas density in hydrogen atoms per cubic centime-
ter. The solid line is the exact equation of state, while the thick dashed line is
a simple fit (see text), which is accurate to about 1%. The vertical dotted line
shows the transition between an isothermal gas (to the left) and the effective
equation of state in our multiphase model (to the right).
out of the cloud material, and the evaporation of clouds in
supernova remnants.
Our strategy is motivated by two basic limitations in mod-
eling star formation on cosmological scales. First, there is no
fundamental theory of this process. Second, large-scale sim-
ulations lack the resolution that would be required to char-
acterize the ISM in detail. Our statistical method decouples
our ignorance of star formation from its impact on galactic
scales, by coarse-graining the physics of star-forming gas. In
principle, this strategy could be applied to any model of star
formation.
For a detailed description of our multiphase model we refer
the reader to Springel & Hernquist (2003a). For definiteness,
we list in Table 1 the parameter values adopted in our sim-
ulations. The parameters include the efficiency of the evap-
oration process A0, the mass fraction of stars that are short-
lived and die as supernovae β, the cold gas cloud temperature
Tcloud, the effective supernova temperature TSN, and the gas
consumption time-scale t ⋆
0 . As Springel & Hernquist (2003a)
argue, all of these aside from one, e.g. t ⋆
0 , can be constrained
0 to match the Ken-
by simple physical arguments. We adjust t ⋆
nicutt Law which describes empirically observed star forma-
tion rates in nearby galaxies (e.g. Kennicutt 1989, 1998).
On the surface, our description of star-forming gas appears
complicated, with the properties of the different phases de-
scribed by differential equations that include the various pro-
cesses listed above. However, in the end, our model can be
reduced to two main ingredients that make our results easier
to interpret. Springel & Hernquist (2003a) showed that the
star-forming gas quickly establishes a self-regulated cycle in
which the mass and energy of the phases can be approximated
by equilibrium solutions. In this limit, the impact of star for-
mation and feedback can be reduced to: 1) the rate at which
gas is converted into stars, and 2) the effective equation of
state for star-forming gas.
Here, as in Springel & Hernquist (2003a), we choose to pa-
Disk Galaxy Formation in a ΛCDM Universe
3
TABLE 1. SPRINGEL & HERNQUIST (2003A) MODEL PARAMETERS
Parameter
Symbol
Value
Supernova Evaporation
Mass Fraction of Stars > 8M⊙
Cold Cloud Temperature
Effective Supernova Temperature
A0
β
Tclouds
TSN
1000
0.1
1000 K
108 K
rameterize the star formation rate by
dρ∗
dt
= (1 - β)
ρc
t∗
,
where
t∗ = t ⋆
0(cid:18) ρ
ρth(cid:19)
(1)
(2)
.
- 1/2
0 is chosen to match the Kennicutt Law,
The free parameter t ⋆
ρth is a threshold density above which gas is subject to ther-
mal instability and star formation, and ρc is the density of
clouds which, in equilibrium, is given by equation (18) of
Springel & Hernquist (2003a). Springel & Hernquist (2003a)
showed that ρth can be fixed by e.g. requiring the equation of
state to be continuous.
In our subresolution model, feedback from supernovae adds
thermal energy to the ISM, pressurizing it and modifying its
equation of state. On the scales relevant to the dynamics of the
gas, this is described by an effective equation of state, which
can be written
Peff = (γ - 1) ρueff,
(3)
where γ is the ratio of specific heats, ρ is the total gas den-
sity, and ueff is the effective specific thermal energy of the
star-forming gas which, in equilibrium, is given by equation
(19) of Springel & Hernquist (2003a). The effective equation
of state for the particular model parameters employed in our
study is illustrated in Figure 1, in cgs units. Also shown in
Figure 1 is a simple fit to the equation of state, which is accu-
rate to about 1% and is given by
log Peff = 0.050 (lognH)3 - 0.246 (lognH)2 +
1.749 log nH - 10.6 , for log nH > - 0.89 .
(4)
In the particular example shown in Figure 1, the gas is as-
sumed to be isothermal for densities ρ ≤ ρth.
In cgs units,
for the parameter choices adopted here, the critical density at
which the gas begins to depart from isothermality is nH,th =
0.128 cm- 3. This choice ensures that the effective pressure
is a continuous function of density. For the equation of state
shown in Figure 1, a different value of nH,th would introduce
an unphysical jump in pressure at the transition density where
the gas is no longer isothermal.
Our
implementation of
feedback differs significantly
from previous works
(e.g. Sommer-Larsen et al. 2003;
Governato et al. 2002). In our subresolution model, feedback
energy is "stored" in the surrounding gas, adding pressure
support to it. As shown in Figure 1, at densities ρ > ρth, the
equation of state becomes stiffer than isothermal due to this
feedback energy. This pressurization regulates star formation
occurring within the ISM and, as we discuss later, modifies
the bulk dynamics of the gas by altering the pressure gradient
term in Euler's equation. It is this property of our descrip-
tion that we believe is responsible for many of the differences
between our numerical results and earlier simulations.
4
Robertson et al.
FIG. 2. -- Entire simulation volume at z = 0 projected onto a 10 h- 1 Mpc square area, showing the end-state of the simulation. The upper left and right panels
show the HRDM and LRDM particles, respectively. The lower left and right panels show the gas and star particles, respectively. Owing to gravitational effects
the low- and high-resolution particles have intermingled at the boundary of the high-resolution region, but the LRDM particles have not penetrated the inner
high-resolution region. Star particles have spawned in regions where the gas density has exceeded the critical density for star formation.
In passing, we note that for the purposes of large-scale dy-
namics, our multiphase model can be reduced to the choice of
star formation rate, e.g. equation (1), and the effective equa-
tion of state of star-forming gas, e.g. equation (5) and Fig-
ure 1. The results presented here could be obtained with any
hydrodynamical code that uses the same expression for the
star formation rate and an equation of state like that in Fig-
ure 1, without specific reference to all the complexities of our
multiphase model.
In Figure 2, we show the evolved simulation volume at
z = 0, in projection. The upper and lower left panels show
the high-resolution dark matter (HRDM) and gas particles,
respectively. The high-resolution region of the simulation has
increased in size from the initial 5 h- 1 Mpc cubic region at
the center of the simulation, owing to the growth of gravita-
tional structure. Some of the low- and high-resolution parti-
cles have mixed in the boundary region of the low-resolution
zone, whose shape has also been distorted. However, the
LRDM particles (upper right panel) have not penetrated into
the inner regions of the high-resolution region where many
luminous galaxies have formed out of the gas, as marked by
the star particles (lower right panel).
3. DISK GALAXY SELECTION
We use a standard friends-of-friends (FOF) group-finding
algorithm to identify dark matter halos, restricting ourselves
to objects comprised of high-resolution particles. We then
identify the most bound star or gas particle within each ob-
ject and produce a new collection of particles for each halo
that encompasses all particles that lie within the radius R200
that encloses 200 times the critical density around these most-
bound baryonic particles. The group catalogue produced from
this procedure serves as our primary galaxy sample.
From our galaxy catalogue, we have visually inspected the
10 largest objects, measured their angular momenta, and se-
lected a large disk galaxy for further detailed analysis in this
paper. This galaxy displays the most disk-like morphology of
all the galaxies in the simulation resolved with a similar num-
ber of particles. While we do find other flattened systems in
the simulation, the majority of them do not display such well
defined disk-like morphology and kinematics as the galaxy
described below.
The galaxy analyzed here, which we term "galaxy C1" for
simplicity, has 61,372 dark matter, 21,506 gas, and 88,138
star particles within R200 at z = 0. Note that the star particle
mass was set to half the original mass of the gas particles,
so the overall baryon fraction within R200 is 14.1%; slightly
larger than the universal baryon fraction of 13.3%. Figure
3 shows the star (left panels) and gas (right panels) particle
distributions within 0.1 R200 at z = 0, projected perpendicular
(upper panels) and parallel (lower panels) to the direction of
the stellar angular momentum. In addition, the lower right
Disk Galaxy Formation in a ΛCDM Universe
5
FIG. 3. -- Disk galaxy at z = 0, displaying the star (left panels) and gas (right panels) particle distributions within 0.1 R200, projected perpendicular (upper panels)
and parallel (lower panels) to the direction of the stellar angular momentum vector. The lower right panel also traces the gas velocity field and demonstrates the
rotational support of the galaxy.
panel traces the gas velocity field in the plane of the disk. The
gas in the galaxy has collapsed into a rotationally supported
disk, surrounded by a thicker stellar disk. The velocity field of
the galaxy illustrates the mainly circular trajectories of the gas
particles in the disk plane. A weak bar is visible in the disk
gas, but is not strong in the stars that dominate the baryonic
mass of the galaxy. We note that the stellar component is
slightly elongated in the direction of the bar, however. The
galaxy has reached a mass of M200 = 2.16 × 1011h- 1 M⊙ at
z = 0, with a stellar mass of M⋆ = 2.05 × 1010h- 1 M⊙ and a
gas mass of Mgas = 9.44 × 109 h- 1 M⊙. The baryonic mass of
the galaxy is then 68% stellar and 32% gaseous.
4. STRUCTURAL PROPERTIES OF GALAXY C1
4.1. Mass surface density profiles
Figure 4 shows the stellar (diamonds) and gas (triangles)
mass surface densities of galaxy C1 projected along the stellar
angular momentum vector onto the disk plane. The stellar
profile includes particles within 30 h- 1 kpc of the disk center,
while the gas profile includes particles only within 2 h- 1 kpc
of the disk plane. The stellar and gas mass surface densities
are roughly exponential out to r = 8 h- 1 kpc. If we fit them
with an exponential of the form
Σ(r) = Σ0 e- r/Rd ,
(5)
we find best fit values of Σ0,⋆ = 452 h M⊙pc- 2, Rd,⋆ =
2.3 h- 1kpc for the stars and Σ0,gas = 15.3 h M⊙pc- 2, Rd,gas =
4.7 h- 1kpc for the gas. In the plot, we include least squares
fits to the stellar and gas mass surface densities. For com-
parison, the stellar disk of the Milky Way has a scale-length
RMW ∼ 3.5 kpc, similar to the physical scale-length of ∼ 3.3
kpc for our simulated galaxy C1. We note that the gas mass
surface density begins to diverge from the exponential fit be-
tween 8 - 10 h- 1 kpc, at the edge of the gaseous disk.
4.2. Photometric properties
From the known masses, metallicities, and ages of the star
particles, we use a population spectral synthesis code to calcu-
late the SDSS ugriz (Fukugita et al. 1996) AB system magni-
tudes and JHK magnitudes for each particle. The population
synthesis code assumes a Kroupa (2001) IMF and limits input
metallicities to a range of 0.005 Z⊙ < Z < 2.5 Z⊙, outside of
which we adopt the minimum or maximum value. We then
use the Fukugita et al. (1996) conversions from SDSS colors
to the Johnson-Morgan-Cousins UBV RI (Johnson & Morgan
1953; Cousins 1978) Vega system to calculate the absolute
magnitude of the star particles in each band. We bin star par-
ticles within 30 h- 1 kpc of the center of the galaxy into an-
nuli and calculate their total luminosities to determine a sur-
face brightness profile. The measured surface brightness of
galaxy C1 in B, V , I, and K-bands is plotted in Figure 5. The
best fit B-band central surface brightness is µ0,B = 21.0 mag
arcsec- 2 with a B-band luminosity scale-length of Rd,B = 2.0
6
Robertson et al.
FIG. 4. -- Mass surface density for the stars (diamonds) within 30 h- 1 kpc of the galaxy center and gas particles (triangles) within 2 h- 1 kpc of the disk plane.
Using an exponential profile Σ(r) = Σ0 exp(- r/Rd) to fit each component separately, we find best fit values of Σ0,⋆ = 452h M⊙pc- 2, Rd,⋆ = 2.3h- 1kpc for the
stars and Σ0,gas = 15.2h M⊙pc- 2, Rgas = 4.7h- 1kpc for the gas. The least squares fit for the stars (solid line) and gas (dashed line) are plotted over the measured
values.
FIG. 5. -- Surface brightness profile of the galaxy in B (stars), V (dashed line), I (dashed-dotted line), and K (dotted line) bands. The solid line is the fit to the
B-band profile, with best fit central surface brightness µ0,B = 21.0 mag arcsec- 2 and B-band disk scale length Rd,B = 2.0h- 1 kpc. The standard central B-band
surface brightness value for disk galaxies is µB,Freeman = 21.65 mag arcsec- 2. The disk brightness is exponential in each band throughout the inner 10 h- 1 kpc of
the galaxy.
h- 1 kpc. Our measured value for the B-band central surface
brightness agrees well with the canonical value of µ0,B = 21.65
mag arcsec- 2 for spiral galaxies (Freeman 1970).
formed early on, with the majority of stars in the galaxy at
z = 0 being older than 9 billion years. A large fraction of these
stars formed in progenitor halos that later merged to form the
galaxy.
Figure 7 illustrates the history of the metal content of the
galaxy. Each panel shows the average metallicity of newly
formed stars (solid line) over the lifetime of the galaxy and
the average metallicity of all stars formed before each epoch
(dashed line). While the metallicity of newly formed stars
varies in a complex manner, reflecting the details of gas ac-
cretion and merging, the average metallicity of the galaxy
evolves smoothly to the current epoch. The large feature at
z = 0.8 corresponds to a major merger in the assembly his-
tory of the galaxy that significantly altered the metallicity be-
tween z = 0.7 and z = 0.5. Note that stars formed before 9
billion years ago have average metallicities of < 0.8 Z⊙. We
hence expect to see an older population of relatively metal-
poor stars, while regions of the galaxy that actively formed
stars even at late times should feature a younger, metal-rich
population. If a disk does not form stars at the same rate ev-
erywhere, we would expect to see stellar age and metallicity
gradients between regions of differing star formation activity.
Figure 5 demonstrates that the surface brightness profile of
galaxy C1 is exponential from the very center of the galaxy
out to r ∼ 10 h- 1 kpc, and shows no evidence for a bulge com-
ponent. To our knowledge, this object is the first published
example of a galaxy formed within a cosmological simulation
that displays an exponential surface brightness profile with no
significant bulge.
For reference, the u, g, v, r, i, z, U, B, R, I, J, H, and
K-band absolute magnitudes of the galaxy, measured by sum-
ming the luminosities of all stars used to calculate the surface
brightness profiles out to 25.0 mag arcsec- 2 in each band, are
listed in Table 2. The galaxy has realistic colors, and its prop-
erties agree with recent 2MASS determinations of B - K vs.
MK color measurements of nearby spiral galaxies (J. Huchra
2003, private communication). We also list the stellar mass-
to-light ratios Υx in each band, for comparison with observed
galaxies. The stellar mass-to-light ratios in the g, r, i, and z
bands agree well with the properties of observed galaxies in
the Sloan survey (Kauffmann et al. 2003).
4.3. Star and metal formation histories
4.4. Stellar age and metallicity gradients
Within our simulation code, we record the formation time
of each star particle, allowing us to infer the detailed star for-
mation history of each object. Similarly, the metallicity of the
star particles can be used to determine the metallicity evolu-
tion of the galaxies. Figure 6 shows the star formation history
of galaxy C1. As a function of redshift (left panel), the star
formation rate rises below redshift z = 10 and peaks at a value
∼ 7 M⊙ yr- 1 at redshift z ∼ 3, after which it declines sharply
to the present time. In the right panel of Figure 6, we also plot
the star formation rate against lookback time. Viewing the star
formation history in this way illustrates that most of the stars
Combining the formation time and metallicity of each star
particle with information about its position allows us to deter-
mine stellar age and metallicity gradients in the galaxy. The
left panel of Figure 8 shows the stellar age gradient in the
disk, measured for all star particles within 15 h- 1 kpc from
the center of the galaxy. The average age of the star particles
decreases from ∼ 10 Gyr in the outer regions of the disk to
∼ 7.5 Gyr in the center. Beyond a radius of ∼ 10 h- 1 kpc, the
average age of the stars is dominated by contributions from
the stellar halo, where stars are significantly older than in the
disk.
Disk Galaxy Formation in a ΛCDM Universe
7
TABLE 2. PHOTOMETRIC PROPERTIES OF GALAXY C1
Band
u
g
r
i
z
U
B
V
R
I
J
H
K
Magnitude
ΥX
-18.1
12.3
-19.7
3.23
-20.4
1.85
-20.7
1.45
-20.9
1.19
-19.1
5.88
-19.3
4.95
-20.1
2.40
-20.6
1.52
-21.2
0.91
-21.1
0.98
-21.3
0.84
-21.0
1.07
FIG. 6. -- Star formation history of the entire simulated galaxy as a function of redshift (left panel). The star formation rate in the galaxy increases from
redshift z = 10 to a peak of ∼ 7 M⊙ yr- 1 at z ∼ 3. The right panel shows the galactic star formation history as a function of lookback time. Most of the stars in
the galaxy formed earlier than 9 Gyr ago.
FIG. 7. -- Metal formation history of the simulated galaxy. In each panel, the solid line shows the average mass-weighted metallicity of newly formed stars
over the lifetime of the galaxy. The dashed line shows the running mass-weighted average metallicity of all stars within the galaxy. The left panel is plotted as a
function of redshift and the right panel is plotted against lookback time. The large feature at z = 0.8 corresponds to a major merger in the assembly history of the
galaxy that significantly altered the metallicity between z = 0.7 and z = 0.5.
8
Robertson et al.
FIG. 8. -- Stellar age gradient in the disk (left panel), measured by averaging the ages of star particles in 0.5 h- 1kpc radial bins. The average age of the stars
decreases from ∼ 10 Gyr in the outer regions of the disk to ∼ 7.5 Gyr in the center. The right panel shows the stellar (solid) and gas (dashed) metallicity gradients
in the disk. The metallicity of the stars increases from ∼ 0.75 Z⊙ at the edge of the disk to ∼ 1.3 Z⊙ in the center, while that of the gas in the disk increases
strongly from primordial values in the outer disk to ∼ 2.5 Z⊙ in the central regions. Stars forming after z = 0.25 (dotted line) track the metallicity gradient of
star-forming gas in the disk.
In the right panel of Figure 8 we show the metallicity gra-
dient in the galaxy, plotting the average metallicity of the star
(solid line) and gas (dashed line) components in 0.5h- 1 kpc
radial bins. The metallicity gradient of the stars tracks the
stellar age gradient closely, remaining at a roughly constant
value of ∼ 0.7 Z⊙ from r = 15 h- 1 kpc to r = 8 h- 1 kpc and
then increasing to a central value of Z= 1.3 Z⊙. Again, the
halo stars beyond 10 h- 1kpc are older and deficient in metals
compared with the younger stars in the disk.
Interestingly,
the gas metallicity gradient is much stronger than the stellar
metallicity gradient, increasing from close to primordial val-
ues at r = 8 h- 1 kpc to ∼ 2.6 Z⊙ near the center of the disk.
The gas in the inner regions has been enriched by stars form-
ing at late times (dotted line, for stars forming after z = 0.25),
whose metallicities have been enhanced by the long history
of star formation in the galaxy. The metallicity gradients of
the star-forming gas and recently formed stars in the disk then
track each other, again highlighting the strong metallicity gra-
dient in the gaseous disk.
The active star formation in the disk produces a collection
of stars that are younger and more metal enriched than in
the outer regions of the galaxy. Hence, the disk is not sim-
ply built up inside-out, but instead partially forms outside-in
(Sommer-Larsen et al. 2003). The more active star formation
at late times in the inner disk makes this region actually quite
young, on average. In the halo of the galaxy, where star for-
mation is dormant, stars are significantly older and more metal
poor than in the disk.
5. KINEMATIC PROPERTIES OF GALAXY C1
5.1. Angular momentum
Perhaps the most important kinematic property of disk
galaxies is their angular momentum, as this quantity deter-
mines the sizes of thin, rotationally supported disks (Mo et al.
1998). While measuring the specific angular momentum j⋆ of
each star particle is straightforward, deciding which particles
to use for measuring the angular momentum content of the
disk is less clear, in particular when comparing with observa-
tions.
We wish to compare our simulated galaxy with the I-band
Tully-Fisher relation measured by Mathewson et al. (1992)
and Courteau (1997) and the corresponding j⋆ - Vrot relation
(as compiled by Navarro 1998). The measured rotation veloc-
ity Vrot used for these relations is determined observationally
at 2.2 times the exponential scale-length of the stellar surface
brightness, Rd. We then take Vrot to be
Vrot =r GM(< RV )
RV
RV = 2.2Rd,B,
,
(6)
(7)
where Rd,B is the B-band surface brightness scale-length de-
termined in § 4.2. In what follows, when we refer to Vrot we
always mean the expected rotation velocity corresponding to
the circular velocity produced by all mass within 2.2 × Rd,B =
0.045R200 for galaxy C1.
Having determined the rotation velocity Vrot, we then con-
sider 3 different approaches for measuring the specific stellar
angular momentum j⋆, providing slightly different interpre-
tations of the angular momentum content of the galaxy. For
all the star particles, we define the total specific stellar angular
momentum j⋆,tot to be a sum over the star particles of the form
N⋆
j⋆,tot =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xi=1
miri × vi(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
, N⋆
Xi=1
mi,
(8)
where N⋆ is the total number of star particles within R200, vi is
the velocity with respect to the center of mass of the galaxy,
ri is the distance from the center of the galaxy, and mi is the
mass of the ith star particle.
We also consider the specific angular momentum j⋆,RV of
the stars within the radius RV at which the rotation velocity is
measured. This is determined by simply restricting the sums
in equation (8) to the stars within a distance RV of the center
of the galaxy.
The final method we employ mimics the observational esti-
mate for the specific angular momentum content of an expo-
nential disk, viz.
j⋆,obs = 2 Rd,B Vrot.
(9)
Disk Galaxy Formation in a ΛCDM Universe
9
Unlike the previous two methods, this technique indirectly es-
timates the angular momentum content by combining a mea-
surement of the size of the disk with a typical circular veloc-
ity. While allowing an inter-comparison of observed galaxies
with similar rotation velocities (Abadi et al. 2003a), this ap-
proach implicitly assumes that the disk has the rotation curve
of an exponential disk (Courteau 1997), which is not true in
general.
We summarize the angular momentum of our simulated
galaxy in Table 3, listing the values for Vrot and the three
different measurements of
j⋆ along with their values in
physical units. With a rotation velocity Vrot = 121.3 km s- 1
and total specific stellar angular momentum j⋆ = 204.2 h- 1
kpc km s- 1, this galaxy agrees favorably with the observed
range of log( j⋆) ∼ 2.4 - 3.2 for galaxies with log(Vrot) ∼
2.1 (Mathewson et al. 1992; Courteau 1997; Navarro 1998).
However, the region of the galaxy within RV has a specific
angular momentum lower by a factor of ∼ 2 with j⋆,RV =
101.0 h- 1 kpc km s- 1. The lower angular momentum of this
part of the galaxy results from the thickness of the stellar disk.
Finally, the agreement between the value j⋆,obs = 485.2h- 1 kpc
km s- 1 and the observed range of j⋆ demonstrates that the
stellar disk has the proper size when compared with observed
galaxies of similar rotation velocity.
Overall, we find that the total angular momentum content of
the galaxy is appropriate for its rotation velocity, when com-
pared to observations, but that the thick disk reduces the spe-
cific stellar angular momentum content in the inner regions.
We stress that the stellar disk of this galaxy should be consid-
ered deficient in angular momentum content and caution that
only measuring the total specific stellar angular momentum
content of a simulated galaxy can lead to an overestimate of
the specific stellar angular momentum in the central regions
of the galaxy.
At present, we cannot determine whether the low angular
momentum in the inner regions of the galaxy is purely a phys-
ical effect, or is at least partly a consequence of our relatively
poor spatial resolution. The gravitational softening length in
the high-resolution region of our simulation is 0.65h- 1 kpc.
For h = 0.7, this means we cannot resolve thin stellar disks of
galaxies like the Milky Way. It is possible that our estimate
of the angular momentum of the inner parts of galaxy C1 is
underestimated because of this problem.
5.2. Rotation curve
In Figure 9, we plot the rotation velocity of gas parti-
cles from within 2 h- 1 kpc of the disk plane (crosses), pro-
jected onto 0.5 h- 1 kpc wide annuli. We find that the rota-
tion velocity of the gas closely traces the total rotation ve-
locity V (R) = [GM(< R)/R]1/2 (solid line), as expected for a
thin disk without a dominant bulge component. If the cen-
TABLE 3. KINEMATIC PROPERTIES OF GALAXY C1
Method
Total
RV
"Observed"
Radius
(h- 1 kpc)
R200
2.2Rd,B
2Rd,B
Vrot
(km s- 1)
121.3
121.3
121.3
j⋆
(h- 1 kpc km s- 1)
204.2
101.0
485.2
log( j⋆)
+ log(h70)
2.465
2.159
2.841
FIG. 9. -- Rotation velocity of the gas particles within 2 h- 1 kpc of the
disk plane (crosses), projected onto 0.5 h- 1 kpc annuli. We find that the
gas rotation velocity closely traces the total rotation velocity V (R) = [GM(<
R)/R]1/2 (solid line), as expected for a thin disk without a dominant bulge
component.
tral regions of the galaxy contained a kinematically important
bulge, the gas rotation curve would either rise steeply or have
a central maximum and decline outwards. We interpret the
slowly rising rotation velocity of the gas as further evidence
that galaxy C1 consists of a rotationally supported exponen-
tial disk without a significant bulge component.
5.3. I-band Tully-Fisher relation
From the photometric modeling described in § 4.2 and
the kinematic measurements in § 5.1, it is possible to com-
pare galaxy C1 with the I-band Tully-Fisher relation as deter-
mined by Giovanelli et al. (1997), Han & Mould (1992), and
Mathewson et al. (1992). For our measured I-band magni-
tude of I = - 21.2 and rotation velocity Vrot = 121.3 km s- 1,
we find our galaxy to be in good agreement with the aver-
age observed I-band magnitude of I ∼ - 20.8 for galaxies with
log(Vrot) ∼ 2.1.
6. ISOLATED HALO SIMULATIONS
The formation of a well-defined disk galaxy without a dom-
inant bulge can be viewed as a significant success towards
obtaining realistic galaxies in cosmological simulations. It is
therefore of particular interest to understand which aspects of
our modeling are mainly responsible for this outcome.
A novel feature of our approach lies in our use of a sub-
resolution model for the multiphase structure of the ISM, to-
gether with an effective equation of state that this model im-
plies for the dense star-forming gas. As can be seen in Figure
1, when the gas is sufficiently dense to be thermally unsta-
ble and subject to star formation, the equation of state departs
from that of an isothermal gas, and becomes stiffer with in-
creasing density. This will affect the bulk hydrodynamics of
the gas, through the pressure gradient term in the Euler equa-
tion. Consequently, dense gaseous disks will experience addi-
tional pressure support from supernova feedback, in a manner
that is different from previous attempts to model galaxy for-
mation.
In order to examine the importance of this effect further,
10
Robertson et al.
FIG. 10. -- Stellar mass distribution, adaptively smoothed and seen edge- and face-on, for the "iso-ism" model at (clockwise from upper left) T = 0.5 Gyr,
T = 1.0 Gyr, T = 1.5 Gyr, and T = 2.0 Gyr. As the disk develops, gas quickly cools to roughly 104 K and forms stars. The 104 K gas is Toomre unstable and
fragments into dense clumps, which in turn efficiently form stars, leading to the clumpy stellar mass distribution in the disk. As the galaxy ages, the clumps lose
angular momentum through dynamical friction, coalesce, and fall into the center of the galaxy.
gas physics in detail.
we use a set of simulations to consider disks forming in iso-
lated halos (e.g. Springel & Hernquist 2003a), where we vary
the description of star-forming gas and associated feedback
effects. These results show that the physics of dense, star-
forming gas has a substantial impact on the dynamics of disks,
indicating that the model of the ISM plays a crucial role in our
successful formation of a disk galaxy in a cosmological con-
text.
6.1. Initial conditions and simulation set
In order to isolate the dependence of disk formation on
the physics of the ISM, we investigate the dynamics of gas
in static spherically symmetric NFW (Navarro et al. 1995)
dark matter halos of mass 1012h- 1 M⊙ and concentration
cNFW = 20. We place 40,000 gas particles that initially trace
the dark matter in these halos and set the initial temperature
profile such that the gas is in hydrostatic equilibrium when
evolved with non-radiative hydrodynamics. We add angular
momentum to the gas, corresponding to a relatively high spin
parameter of λ = JE1/2/(GM5/2
vir ) = 0.1. We then evolve the
gas forward in time including radiative cooling. Under these
conditions, we expect the gas to collapse smoothly into a thin,
rotationally supported disk. We do not intend for this to be
a realistic model of disk formation in a hierarchical universe,
but the idealized nature of the simulations described below
makes it possible to examine the consequences of varying the
This set-up provides initial conditions for a suite of simu-
lations, where we consider three different descriptions of the
ISM. The first model, which we will refer to as "justcool,"
consists of simply allowing the gas to cool with primordial
abundances without allowing for any star formation or feed-
back. The "justcool" scenario serves as an extreme reference
case where star formation and feedback cannot affect the gas
dynamics. Since the dense gas that settles into a disk can cool
efficiently, its temperature will remain close to 104 K, effec-
tively obeying an isothermal equation of state.
Our second description of the ISM, which we refer to as
"iso-ism," includes gas cooling and star formation, according
to equation (1), but no feedback effects. Gas in the disk again
obeys a nearly isothermal equation of state in this case, but
compared to our "justcool" model, dense gas is converted into
collisionless star particles. For ease of comparison, we use the
same star formation rate for gas of a given density as in our
multiphase model.
the ISM is
third description of
the
Springel & Hernquist (2003a) multiphase ISM model, which
includes radiative cooling, star formation, and supernova
feedback in the form of thermal energy input and cloud evap-
oration, as described earlier. We refer to this model as the
"multiphase-ism" model. The primary difference between the
"multiphase-ism" and "iso-ism" scenarios is that in the mul-
Finally,
our
Disk Galaxy Formation in a ΛCDM Universe
11
FIG. 11. -- Stellar mass distribution, adaptively smoothed and seen edge- and face-on, for the "multiphase-ism" model at (clockwise from upper left) T = 0.5
Gyr, T = 1.0 Gyr, T = 1.5 Gyr, and T = 2.0 Gyr. In the "multiphase-ism" run, the gas cools into a thin disk which begins forming stars at a rate regulated by
supernova feedback. While the gas is comprised of clouds cool enough to form stars, the contributions from the hot surrounding medium to the effective equation
of state of the gas provides pressure support to the ISM to prevent Toomre instability. The end result is a large disk galaxy with only a modest bulge component.
tiphase model feedback from supernovae modifies the effec-
tive equation of state of star-forming gas, providing increased
pressure support, as indicated by Figure 1. The stiffer equa-
tion of state helps to regulate star formation so that it occurs
at a rate consistent with observations of isolated disk galax-
ies (Kennicutt 1989, 1998; Martin & Kennicutt 2001) once
the free parameter t ⋆
0 of the model is properly adjusted (see
Springel & Hernquist 2003a,b).
For each ISM model, we ran a number of simulations, vary-
ing the numerical integration parameters to test the robustness
of the results.
In particular, we varied the number of SPH
neighbors and the value of the time-step taken by the code.
None of these choices had a significant effect on the outcome.
Therefore, our results appear to have converged with respect
to time integration. When the number of SPH neighbors is
varied, instabilities in the disks tend to occur at slightly dif-
ferent times and locations, but the qualitative behavior of the
simulations remains unchanged.
6.2. Comparison of disks formed in isolated halos
The time evolution of simulations with our three ISM mod-
els reveals interesting differences. In Figure 10, we show the
stellar mass distribution, adaptively smoothed and seen edge-
and face-on, for the "iso-ism" model at (clockwise from upper
left) T = 0.5 Gyr, T = 1.0 Gyr, T = 1.5 Gyr, and T = 2.0 Gyr.
Gas in the inner parts of the halo quickly cools to roughly 104
K, settles into a thin, rotationally supported disk, and forms
stars. The 104 K gas in the disk is highly compressible, obey-
ing a nearly isothermal equation of state. Consequently, the
self-gravitating gaseous disk becomes gravitationally unsta-
ble as soon as it reaches a moderate surface density. The
Toomre (1964) instability quickly causes a fragmentation of
the smooth gas disk into clumps, which collapse further under
their own self-gravity and form stars on a short time-scale. As
the galaxy ages, the stellar clumps coalesce through dynam-
ical friction into ever larger lumps, destroying the disk. By
T = 2 Gyr, the stars are mainly confined in two large blobs,
with numerous smaller clumps orbiting the center.
The evolution of the galaxy in the "iso-ism" model contrasts
sharply with that in the corresponding "multiphase-ism" run.
We illustrate this in in Figure 11. Again, the gas cools into a
thin, star-forming disk. However, supernova feedback regu-
lates the rate of star formation and the dynamics of the dense
star-forming gas. Owing to feedback, the ISM is pressur-
ized, stabilizing the gas against the Toomre instability. Con-
sequently, a smooth distribution of gas is maintained in the
plane of the disk and stars form steadily, increasing the size
of the disk from 6 h- 1 kpc at T = 0.5 Gyr to > 12 h- 1 kpc at
T = 2.0 Gyr. Since the gas remains stable and does not frag-
ment into clumps, the disk largely avoids angular momentum
loss and the simulation yields a large disk galaxy with only a
modest bulge.
12
Robertson et al.
FIG. 12. -- Comparison of the gas mass distribution in the "justcool" (left), "iso-ism" (center), and "multiphase-ism" (right) simulations at T = 1.0 Gyr. Both
the "justcool" and "iso-ism" models produce gas disks fragmented by catastrophic Toomre instability. By contrast, the "multiphase-ism" model yields a stable
disk supported by additional pressure supplied by the Springel & Hernquist (2003a) model for the ISM.
In Figure 12, we compare the gas mass distributions in
the "justcool" (left), "iso-ism" (center), and "multiphase-ism"
(right) simulations at T = 1.0 Gyr. Both the "justcool" and
"iso-ism" models produce clumpy gas disks that fragmented
by the Toomre instability. In fact, the time evolution of the
"justcool" and "iso-ism" models is very similar, except that
in the former case the instabilities occur earlier, showing that
star formation can have a stabilizing effect by reducing the gas
surface density. Note, however, that only the "multiphase-
ism" model produces a gaseous disk that exhibits long-term
stability.
In summary, our simulations of the formation of disks in
isolated halos clearly show that the modeling of the ISM has
a strong effect on disk stability. In particular, an isothermal
equation of state for star-forming gas produces gaseous disks
that are violently unstable. Under these conditions it would be
difficult if not impossible to form large, extended disk galax-
ies unless the gas surface density is always kept extremely
low, for example by adopting an unrealistically short star for-
mation time-scale that would conflict with the observed Ken-
nicutt law.
In contrast, the multiphase model for the ISM yields disk
galaxies that are more stable. The stiffer equation of state pro-
duces disks that are more resilient against Toomre instability
than in models with an effectively isothermal ISM. Fragmen-
tation is avoided, enabling the steady formation of large stellar
disks even when the disk is gas-rich and has a high gaseous
surface density. Therefore, we suggest that the behavior of
the equation of state in our multiphase description of the ISM
is the primary reason why our cosmological simulation can
produce disk galaxies with negligible bulge components. This
conclusion should be insensitive to the detailed features of our
multiphase model of the ISM, provided that the actual equa-
tion of state for star-forming gas behaves similarly to the one
proposed by Springel & Hernquist (2003a).
7. CONCLUSIONS
We have performed a high-resolution cosmological simula-
tion that included gas dynamics, radiative heating and cool-
ing, and star formation and supernova feedback, modeled in
the framework of a subresolution description of a multiphase
ISM. Inspecting the ten largest halos, we identified a well-
resolved disk galaxy at redshift z = 0. The disk galaxy has
realistic photometric properties, including a reasonable cen-
tral surface brightness and exhibits good agreement with the
I-band Tully-Fisher relation. Most important, this object is
the first published example of a disk galaxy from a simulation
that features a simple exponential surface brightness profile
without having a large bulge component. Our work demon-
strates that extended disk galaxies can form in a ΛCDM cos-
mology within full hydrodynamic simulations, an important
step forward in understanding the origin of disk galaxies.
While the total angular momentum of the stellar compo-
nent of the disk galaxy is appropriate for observed galaxies
with similar rotation velocities, the inner regions of the galaxy
have low specific stellar angular momentum. It is not clear
whether this is a problem with the physics, e.g. excessive an-
gular momentum loss by the gas, or is at least partly an artifact
of our inability to completely resolve the thin stellar disk. The
disk galaxy in our simulation does not have an obvious prob-
lem with its radial size, but statistical studies with samples of
simulated disk galaxies will be needed to settle this question.
However, we find the success achieved here encouraging and
suggest that the problem with disk size is alleviated by em-
ploying equations of state for star-forming gas that are stiffer
than isothermal.
Using a suite of isolated galaxy formation simulations, we
have explicitly demonstrated the consequences our treatment
for the ISM has on the formation of disk galaxies. Simulations
with an effectively isothermal ISM yield gravitationally un-
stable disks and fail to produce large, smooth stellar disks. In
contrast, the multiphase model supplies enough pressure sup-
port to the star-forming gas to prevent catastrophic Toomre
instability, allowing stable galactic disks to form. We have
also verified that these conclusions do not depend on details
of our numerical integration.
Our work differs in a number of respects from earlier nu-
merical attempts to study disk galaxy formation. We have
used a novel formulation of smoothed particle hydrodynamics
that, by construction, conserves both energy and entropy si-
Disk Galaxy Formation in a ΛCDM Universe
13
Springel & Hernquist 2003a), so Q > 1 and the disk would be
stable, explaining why our "multiphase-ism" models in § 6
produce realistic disks. Moreover, for thin disks of gas, the
fastest growing unstable mode has a wavelength which is a
significant fraction of the disk size (e.g. Binney & Tremaine
1987) and, therefore, the growth of the instability is sensi-
tive only to the macroscopic equation of state for the ISM.
Whether or not the effective equation of state we have adopted
(i.e. Figure 1) for the ISM is appropriate in detail is uncertain,
but this question is largely independent of the exact proper-
ties of star-forming gas. In particular, we expect our results to
generalize to other models for the ISM with similar effective
equations of state.
Having identified one exponential disk galaxy in our simu-
lation, future work is needed to extend our analysis to other
galaxies. For example, a good statistical sample of simulated
disks is needed before we can characterize the distribution
of disk sizes and their morphological types, as well as the
photometric and kinematic properties of the galaxies. This
work will require larger simulations of substantial dynamic
range in order to obtain adequate numbers of simulated
disk galaxies, but we find the initial successes described
here and recently by other authors highly encouraging. A
comparison with the wealth of observational data on the
structural parameters of galaxies in the local universe will
then show whether hydrodynamical simulations are finally
producing realistic disk galaxies.
This work was supported in part by NSF grants ACI 96-
19019, AST 98-02568, AST 99-00877, AST 00-71019, and
AST-0206299, and NASA ATP grants NAG5-12140, NAG5-
13292, and NAG5-13381. NY acknowledges support from
JSPS Special Research Fellowship (02674). The simulations
were performed at the Center for Parallel Astrophysical Com-
puting at the Harvard-Smithsonian Center for Astrophysics.
multaneously, even when smoothing lengths vary in response
to changes in density (Springel & Hernquist 2002). While we
consider it unlikely that previous numerical work on galaxy
formation with SPH was strongly affected by inaccuracies
in e.g. entropy conservation, it is clearly preferable to use a
fully conservative scheme that mitigates against any numeri-
cal "overcooling."
More important, we have formulated the consequences of
feedback using a subresolution model for star formation in a
multiphase ISM. In this scheme, the local supernova feedback
modifies the effective equation of state for star-forming gas,
leading to an efficient self-regulation of star formation and a
stabilizing effect on highly overdense gas. This aspect of our
modeling feeds directly into the dynamics of forming disks
by modifying the pressure gradient term in the hydrodynamic
equations of motion.
Our formulation of the ISM physics has a number of advan-
tages over earlier treatments of star formation and feedback.
Other workers have implemented techniques to allow feed-
back to have a significant dynamical impact by, e.g. deposit-
ing thermal energy into gas surrounding star-forming regions
and artificially delaying the radiative loss of this energy by
the gas for an ad hoc time interval. While similar in spirit
to our method, these approaches make the comparison of re-
sults from different calculations challenging because the con-
sequences of feedback are difficult to summarize in a concise,
quantitative form. When expressed in terms of an effective
equation of state, as done here, the dynamical implications of
feedback become easier to interpret. A discussion in terms of
the effective equation of state is also relatively insensitive to
the detailed physics responsible for pressurizing the gas, and
enables one to relate the description of feedback to observa-
tions of the ISM.
In order to demonstrate the sensitivity of galaxy formation
to the physics of star-forming gas, we employed a simple
multiphase model of the ISM (Springel & Hernquist 2003a),
where the gas is imagined to be thermally unstable at suf-
ficiently high densities and exists in two distinct phases, in
pressure equilibrium. In detail, it is not plausible that this is
a true picture of the ISM of disk galaxies, because this model
ignores, e.g. turbulent motion, magnetic fields, and cosmic
rays. However, we believe that the basic points of our mod-
eling do not depend on these complications. The evolution
of the different models shown in § 6 depends mainly on the
effective equation of state for the star-forming gas on scales
which characterize the Toomre instability. According to the
Toomre criterion, a thin sheet of gas in differential rotation
will be gravitationally unstable if
csκ
πGΣ
< 1 ,
Q ≡
(10)
- 1
75 ,
where cs is the sound speed, κ is the epicyclic frequency, and
Σ is the surface density. In terms of an effective temperature
of the gas,
Q ≈ 1.1 T 1/2
5 κ35 Σ
(11)
where κ35 ≡ κ/(35 km sec- 1kpc- 1) and Σ75 ≡ Σ/(75 M⊙ pc- 2)
are values characteristic of the Milky Way in the solar neigh-
borhood, and T5 ≡ T /(105 K). If the gas obeyed an isothermal
equation of state with T5 = 0.1, then Q ∼ 0.3, and the disk
would be violently unstable, as for example in the case of our
"iso-ism" simulations described in § 6. With our equation of
state, however, the effective temperature at densities charac-
teristic of a forming galaxy would be T5 ∼ 1 (see Fig. 1 of
14
Robertson et al.
REFERENCES
Abadi, M. G., Navarro, J. F., Steinmetz, M., & Eke, V. R. 2003a, ApJ, 591,
Nagamine, K., Springel, V., & Hernquist, L. 2003a, MNRAS, accepted
-- . 2003b, ApJ, 597, 21
Binney, J. & Tremaine, S. 1987, Galactic Dynamics, Princeton University
-- . 2003b, MNRAS, accepted (astro-ph/0305409)
Nagamine, K., Cen, R., Hernquist, L., Ostriker, J.P. & Springel, V. 2003c,
Contardo, G., Steinmetz, M., & Fritze-v. Alvensleben, U. 1998, ApJ, 507,
Nagamine, K., Springel, V., Hernquist, L. & Machacek, M. 2003d, MNRAS,
499
Press
497
Courteau, S. 1997, AJ, 114, 2402
Cousins, A. W. J. 1978, Monthly Notes of the Astronomical Society of South
Africa, 37, 8
Davé, R., Hernquist, L., Katz, N. & Weinberg, D.H. 1999, ApJ, 511, 521
Freeman, K. C. 1970, ApJ, 160, 811
Fukugita, M., Ichikawa, T., Gunn, J. E., Doi, M., Shimasaku, K., & Schneider,
D. P. 1996, AJ, 111, 1748
Giovanelli, R., Haynes, M. P., Herter, T., Vogt, N. P., Wegner, G., Salzer, J. J.,
da Costa, L. N., & Freudling, W. 1997, AJ, 113, 22
Governato, F., et al. 2002, preprint (astro-ph/0207044)
Han, M., & Mould, J. R. 1992, ApJ, 396, 453
Hernquist, L. 1993, ApJ, 404, 717
Hernquist, L. & Springel, V. 2003, MNRAS, 341, 1253
Hultman, J. & Pharasyn, A. 1999, A&A, 347, 769
Johnson, H. L., & Morgan, W. W. 1953, ApJ, 117, 313
Katz, N. 1992, ApJ, 391, 502
Katz, N., & Gunn, J. E. 1991, ApJ, 377, 365
Katz, N., Weinberg, D.H. & Hernquist, L. 1996, ApJS, 105, 19
Kauffmann, G., et al. 2003, MNRAS, 341, 33
Kennicutt, R. C. 1989, ApJ, 344, 685
-- . 1998, ApJ, 498, 541
Kroupa, P. 2001, MNRAS, 322, 231
Martin, C. L., & Kennicutt, R. C. 2001, ApJ, 555, 301
Mathewson, D. S., Ford, V. L., & Buchhorn, M. 1992, ApJS, 81, 413
McKee, C. F., & Ostriker, J. P. 1977, ApJ, 218, 148
Mo, H. J., Mao, S., & White, S. D. M. 1998, MNRAS, 295, 319
(astro-ph/0302187)
ApJ, submitted (astro-ph/0311294)
submitted (astro-ph/0311295)
Navarro, J. F. 1998, preprint (astro-ph/9807084)
Navarro, J. F., & Benz, W. 1991, ApJ, 380, 320
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1995, MNRAS, 275, 56
Navarro, J. F., & Steinmetz, M. 1997, ApJ, 478, 13
-- . 2000, ApJ, 538, 477
Navarro, J. F., & White, S. D. M. 1993, MNRAS, 265, 271
-- . 1994, MNRAS, 267, 401
O'Shea, B.W., Nagamine, K., Springel, V., Hernquist, L. & Norman, M.L.
2003, ApJ, submitted (astro-ph/0312651)
Power, C., Navarro, J. F., Jenkins, A., Frenk, C. S., White, S. D. M., Springel,
V., Stadel, J., & Quinn, T. 2003, MNRAS, 338, 14
Sommer-Larsen, J., & Dolgov, A. 2001, ApJ, 551, 608
Sommer-Larsen, J., Götz, M., & Portinari, L. 2002, Ap&SS, 281, 519
-- . 2003, ApJ, 596, 47
Sommer-Larsen, J., Gelato, S., & Vedel, H. 1999, ApJ, 519, 501
Springel, V., & Hernquist, L. 2002, MNRAS, 333, 649
-- . 2003a, MNRAS, 339, 289
-- . 2003b, MNRAS, 339, 312 (for an updated version, see astro-ph/0206395)
Steinmetz, M., & Müller, E. 1994, A&A, 281, L97
Steinmetz, M., & Müller, E. 1995, MNRAS, 276, 549
Steinmetz, M., & Navarro, J. F. 1999, ApJ, 513, 555
Toomre, A. 1964, ApJ, 139, 1217
Tormen, G. 1997, MNRAS, 290, 411
Yepes, G., Kates, R., Khoklov, A. & Klypin, A. 1997, MNRAS, 284, 235
|
astro-ph/0609412 | 1 | 0609 | 2006-09-14T16:02:18 | Small-Scale X-ray Variability in the Cassiopeia A Supernova Remnant | [
"astro-ph"
] | A comparison of X-ray observations of the Cassiopeia A supernova remnant taken in 2000, 2002, and 2004 with the Chandra ACIS-S3 reveals the presence of several small scale features (<= 10 arcsec) which exhibit significant intensity changes over a 4 year time frame. Here we report on the variability of six features, four of which show count rate increases from ~ 10% to over 90%, and two which show decreases of ~ 30% -- 40%. While extracted 1-4.5 keV X-ray spectra do not reveal gross changes in emission line strengths, spectral fits using non-equilibrium ionization, metal-rich plasma models indicate increased or decreased electron temperatures for features showing increasing or decreasing count rates, respectively. Based on the observed count rate changes and the assumption that the freely expanding ejecta has a velocity of ~ 5000 km/s at the reverse shock front, we estimate the unshocked ejecta to have spatial scale variations of 0.02 - 0.03 pc, which is consistent with the X-ray emitting ejecta belonging to a more diffuse component of the supernova ejecta than that seen in the optically emitting ejecta, which have spatial scales ~ 0.001 pc. | astro-ph | astro-ph | Submitted to the Astronomical Journal
Preprint typeset using LATEX style emulateapj v. 11/26/04
6
0
0
2
p
e
S
4
1
1
v
2
1
4
9
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
SMALL-SCALE X-RAY VARIABILITY IN THE CASSIOPEIA A SUPERNOVA REMNANT
Daniel J. Patnaude1 & Robert A. Fesen 2
Submitted to the Astronomical Journal
ABSTRACT
A comparison of X-ray observations of the Cassiopeia A supernova remnant taken in 2000, 2002,
and 2004 with the Chandra ACIS-S3 reveals the presence of several small scale features (≤ 10′′) which
exhibit significant intensity changes over a 4 year time frame. Here we report on the variability of
six features, four of which show count rate increases from ∼ 10% to over 90%, and two which show
decreases of ∼ 30% -- 40%. While extracted 1-4.5 keV X-ray spectra do not reveal gross changes
in emission line strengths, spectral fits using non-equilibrium ionization, metal-rich plasma models
indicate increased or decreased electron temperatures for features showing increasing or decreasing
count rates, respectively. Based on the observed count rate changes and the assumption that the
freely expanding ejecta has a velocity of ∼ 5000 km s−1 at the reverse shock front, we estimate the
unshocked ejecta to have spatial scale variations of 0.02 − 0.03 pc, which is consistent with the X-ray
emitting ejecta belonging to a more diffuse component of the supernova ejecta than that seen in the
optically emitting ejecta, which have spatial scales ∼ 10−3 pc.
Subject headings: ISM: individual (Cassiopeia A) - supernova remnants - ISM: dynamics
1. INTRODUCTION
Hwang et al. 2000).
Cassiopeia A (Cas A) is currently the youngest
known Galactic supernova remnant (SNR) with an es-
timated explosion date no earlier than around 1670
AD (Thorstensen et al. 2001; Fesen et al. 2006a). The
remnant consists of optical, infrared, and X-ray emis-
sion arranged in a shell roughly 4′ diameter (≃ 4
pc at 3.4+0.3
−0.1 kpc; Reed et al. 1995) consisting of
undiluted supernova (SN) ejecta rich in O, Si, Ar,
Ca, and Fe heated by the remnant's ≃ 3000 km
s−1 reverse shock (Chevalier & Kirshner 1978, 1979;
Douvion et al. 1999; Hughes et al. 2000; Willingale et al.
2003; Hwang & Laming 2003; Laming & Hwang 2003;
Morse et al. 2004).
Viewed in X-rays, the remnant's brightest emission is
concentrated in a 30′′ thick, ≃ 110′′ radius emission ring
dominated by thermal plasma arising from shock heated
SN debris. Farther out beyond this emission ring, lies
fainter filamentary X-ray emission (radius ≃ 150′′) as-
sociated with the current position of the remnant's ≈
6000 km s−1 forward blast wave as it moves through the
surrounding circumstellar material (Gotthelf et al. 2001;
DeLaney & Rudnick 2003).
If the remnant's expanding SN ejecta were uniform in
density and chemical composition, X-ray emission from
the shocked ejecta would appear as a uniform ring of
emission with little spectral variation. What is actually
observed, however, is a patchy, irregular ring of clumpy
emission reflecting the inhomogeneous nature of the ex-
panding SN debris as it interacts with the reverse shock
front. Due both to strong ejecta clumping caused by ra-
diative cooling instabilities and the turbulent mixing of
different chemical layers of the progenitor star during the
SN explosion, the resulting X-ray emission morphology is
structurally and chemically complex (Hughes et al. 2000;
1 Smithsonian Astrophysical Observatory, 60 Garden St, Cam-
bridge, MA 02138
2 6127 Wilder Laboratory, Physics & Astronomy Department,
Dartmouth College, Hanover, NH 03755
Although Cas A's reverse shock front has been located
to arcsecond accuracy in a few regions through high-
resolution optical images (Fesen et al. 2001; Morse et al.
2004), it has only been broadly localized through a sharp
increase in X-ray emission in azimuthal averages of se-
lected portions of the remnant (Gotthelf et al. 2001). On
small spatial scales, evidence for the reverse shock inter-
acting with the ejecta would be manifested in X-rays as
a rise in emission as the SN ejecta is shocked and decel-
erated. Post-shock cooling of the metal-rich ejecta could
lead to equally dramatic localized emission declines.
Here we present an analysis of archival Chandra images
of Cas A which show a number of substantial, small-
scale brightness changes between 2000 and 2004 which
we interpret as marking regions of reverse shock passage.
We report on spectral analyses for six specific regions
and find some plasma temperature changes in the X-ray
emitting ejecta. In §2 we present the relevant Chandra
observations and subsequent data reduction.
In §3 we
discuss the results from our analysis, and summarize our
conclusions in §4.
2. OBSERVATIONS AND DATA REDUCTION
Cas A was observed with the ACIS-S3 chip as part of
the Chandra Guest Observer program for 50 ksec both
on 30 January 2000 and 6 February 2002, and for 1
Msec as a Chandra Very -- Large Project during the first
half of 2004 (see Hwang et al. 2000; Gotthelf et al. 2001;
DeLaney & Rudnick 2003 for details regarding the 2000
and 2002 observations, and Hwang et al. 2004 for the
2004 observations.). The 0.′′492 CCD pixel scale of ACIS
undersamples the telescope's ≃ 0.′′5 resolution. We re-
processed these data using the latest version (3.2) of the
Chandra Calibration Database (CalDB). This reprocess-
ing focused on recalculating the aspect solution for better
positional accuracy. Version 3.2 of the CalDB was also
used in the generation of detector response matrices used
in the spectral fitting, which is discussed below.
Both direct visual comparisons of the three epoch im-
2
Patnaude & Fesen
ages and image differences and subtractions were used
to identify several small scale regions where the detected
flux appeared to vary with time. Six selected regions
showing significant flux variations in the total ACIS band
(0.3 -- 10.0 keV) are marked in Figure 1 with enlarge-
ments shown for the years 2000, 2002, and 2004 in Fig-
ure 2. The coordinates, sizes of the extracted flux re-
gions, and count rates are given in Table 1. The quoted
rates for each epoch are normalized to the observed
count rate of the remnant's central X-ray point source
(Tananbaum 1999; Aschenbach 1999; Pavlov & Zavlin
1999).
Spectra were extracted for the six regions listed in Ta-
ble 1. Background subtraction was done by selecting
regions close to the knots. The background spectra were
binned such that each detector channel contained a min-
imum of 10 counts. The background spectra were then
fit to models appropriate for a thermal plasma in non-
equilibrium ionization. More specifically, since we are
investigating regions in the interior of the SNR, the back-
ground contains contributions from swept up circumstel-
lar material as well as diffuse emission from shocked
supernova ejecta around the regions of interest. This
method of background subtraction is preferred to using
ACIS Blank-Sky images since scattered X-rays from the
SNR are the dominant source of background here. Fur-
thermore, while the source spectra were found to vary,
the background spectra comes from the diffuse emission
around the knots and does not vary.
In order to minimize the effects of Chandra's az-
imuthally varying point spread function which are man-
ifest in the combined 2004 observations (Hwang et al.
2004), we limited ourselves to using a 50 ksec subset
(ObsID 5196; 8 February 2004) of the 1 Msec observation
when extracting the 2004 spectra, since the observational
parameters for these data closely match those of the 2000
(ObsID 114) and 2002 (ObsID 1952) observations. The
extracted source and background spectra were fit using
version 11.2 of the X-ray spectral fitting package XSPEC.
The six regions of interest are relatively small in area
(. 30 arcsec2) with total observed counts of 5000 -- 20,000
for the 0.3 -- 10 keV energy a range for each of the three
epoch observations. We grouped the extracted spectra
such that there was a minimum of 10 counts in any de-
tector channel. Moreover, since the goal of this analysis
was not concerned with looking for specific abundance
variations in the ejecta, we chose to limit our spectral
fitting to NEI models (Borkowski et al. 2001) contain-
ing mainly Si, S, and Ar (and Mg and Ca, when there
were prominent emissions visible at ∼ 1.2 and 3.8 keV,
respectively) and limited the fits to between 1.0 and 4.5
keV. For all models, the ratio (Si/Si⊙) was fixed at 1,
while the relative abundances of the other included ele-
ments were allowed to vary. Finally, we chose to freeze
the abundances of all other modeled elements (H, He,
C, N, O, Ne, Ca, Fe, and Ni) at zero. We justify this
by noting that these elements do not have emission lines
in the energy band we are interested in. We do note,
however, that by not including these elements, we may
under-predict the contribution to the flux from the un-
derlying continuum. In Table 2, we list the fitted plasma
temperature for each epoch and each region, and show
the data and the resulting fits in Figures 3 -- 5.
All regions, with the exception of Region 5, are well
modeled as a thermal spectrum. Region 5 appears to
possess purely nonthermal emission and is likely filamen-
tary emission arising from the remnant's forward shock
front viewed in tangent like several other interior pro-
jected filamentary features (Hughes et al. 2000; Region
'D') and common in the remnant's southwestern quad-
rant (DeLaney et al. 2004).
3. RESULTS AND DISCUSSION
3.1. Count Rate Variations
A comparison of 2000 -- 2004 Chandra images of Cas
A shows fairly steady X-ray emission flux from the rem-
nant's shock heated ejecta on both large and small spa-
tial scales. However, there are a few small features in
the main emission ring that do show significant changes
in brightness over this time interval. Most of these show
increases rather than decreases although this may be due
in part to a detection bias against finding changes in dif-
fuse patches of emission, especially if they are faint.
As can be seen qualitatively in Figure 2 and quantita-
tively in Table 1, between 2000.1 and 2004.1 four small
regions in Cas A's X-ray emission structure exhibited sig-
nificant increases in intensity while two regions showed a
decrease. Specifically, Regions 1, 3, and 4 each show in-
creases of ≈ 10% -- 20% between 2000 and 2004, while Re-
gion 2 went from near invisibility in 2000 to being among
the brightest features in the remnant, with a near dou-
bling of its count rate between 2002 and 2004. In 2000,
the count rate in this small region is due mainly to the
diffuse emission in the area around and toward the knot
in Region 2. As can seen in the 2000 and 2002 images,
a new and relatively bright knot emerges here, becoming
twice as bright just two years later.
We note that fine-scale flux changes in the remnant
are not necessarily monotonic. Region 3, for example,
showed both an increase and a decrease over four years;
nearly a 40% increase from 2000 to 2002 but then a 14%
decrease from 2002 to 2004. The short time span of
data taken in early 2004, taken as part of the 1Msec
Very Large Program on Cas A, does not permit us to
investigate whether this decrease continued after Febru-
ary 2004, but future observations could investigate this
question.
Most of these flux changes occur on relatively small
spatial scales, of order a few arcseconds. For example,
the white box in Figure 2 centered on the emerging bright
spot in Region 2 is just 3′′ × 3′′ and yet encloses most of
the observed count rate change. Similarly, the majority
of the brightness change for the northeastern limb emis-
sion knot shown in Region 1 occurs along a segment of
the knot's structure extending 1" - 3" to the northwest.
In contrast to the brightening seen for Regions 1 -- 4,
the X-ray emission detected in Regions 5 and 6 show de-
creases in brightness. Shown in the bottom two rows of
Figure 2, the count rates in these regions drop signifi-
cantly from 2002 to 2004; a ∼ 40% decrease for Region
5 and ∼ 20% in Region 6. For Region 5, the measured
40% change in flux represents an average over several
small, individual features some of which exhibited large
and small changes in brightness.
While it has been known for some time that a layer
of contamination is forming over the ACIS detector thus
reducing its overall efficiency (Plucinsky et al. 2003), the
expected flux reduction from this effect is not enough to
Small-Scale X-ray Variability of Cas A
3
explain the overall decreases we see here. Nonetheless, to
avoid uncertainties due to detector sensitivity degrada-
tion, we list in Table 2 measured count rates normalized
to the observed count rate for the remnant's central X-
ray point source. Conversion of the observed count rates
into intensity changes can be estimated using the X-ray
point source observed FX = 8 × 10−13 erg cm−2 (for
0.6 − 6 keV; see Fesen, Pavlov, & Sanwal 2006 and refs.
therein).
3.2. Spectral Variations
We performed spectral fits for each region to investi-
gate whether rapid changes in the observed count rate
also reflected changes in spectral properties of the emis-
sion features. In general, we find that the modeled spec-
tra are relatively constant from year to year (cf., Region
1, Fig. 3, left, and Regions 4, Fig. 4, right), but there are
exceptions. For example, Region 2 (Fig. 3, right) which
showed the most extreme flux brightening, shows some
evidence for the emergence of helium-like silicon and sul-
fur emission starting in the 2002 spectrum. From our
spectral fits for this region, we also found that kT in-
creased in value from ∼ 1.0 -- 1.5 keV between 2002 and
2004, combined with an increase in net from 4.6 -- 53 ×
1010 cm−3 s over the same time period. Such large in-
creases in flux and ionization timescale suggest that the
ejecta in this location has indeed recently encountered
the remnant's reverse shock front.
Anderson et al. (1994) predict that ejecta knots which
are strongly decelerated by the reverse shock will show
both a marked increase in their thermal emission (like
that seen here), coupled with an increase in synchrotron
emission. We have compared 6 cm VLA radio observa-
tions (∼ 0.′′4 resolution) taken in April 2000 and again
in April 2001 and note a qualitative increase in the radio
emission around Region 2 (Delaney 2006) which is com-
parable to the increase in X-ray emission. This increase
in the radio and X-ray flux is consistent with the inter-
pretation of Anderson et al. (1994). However, we do not
find any corresponding changes in the radio emission for
the other five regions we present here, contrary to what is
expected if we interpret these flux changes in the context
of Anderson et al. (1994). In particular, Region 1, which
like Region 2, also shows a change in kT (from 0.8 -- ∼
1.1 keV) between 2000 and 2004 , does not exhibit a cor-
responding increase in the 6 cm radio emission (Delaney
2006).
On the other hand, model fits for Region 4, a large dif-
fuse patch along the eastern rim and which showed about
the same level of X-ray flux changes as seen for Region
1, exhibited nearly a constant electron temperature over
the four year period. Moreover, Region 3 (Fig. 4, left)
did not exhibit the same sort of strong, monotonic kT
changes as the other three regions. However, based on
the large change in the modeled kT between 2000 and
2002 (Table 2) from ∼ 0.8 -- 1.5 keV, along with the emer-
gence of a broadening near helium-like sulfur (tentatively
identified here as Si Lyβ) suggests that the ejecta here
may be undergoing the same sort of rapid ionization seen
in Region 2. Another possible spectral change seen for
Region 3 is the apparent broadening of the silicon lines in
the 2002 and 2004 spectra. Combined with the increase
in kT , this suggests that the ejecta is in the process of
ionizing up to the hydrogen-like state for silicon.
From its spectrum, Region 4 (Fig. 4; right) appears
to be nearly in ionization equilibrium, and as noted
above (Table 2) the fitted temperature stays nearly con-
stant between 2000 and 2004. Modeling suggests that its
spectrum can be equally well modeled by either a NEI
plasma with an ionization age ∼ 1013 cm−3 s, or by a
collisional ionization model, such as a Raymond -- Smith
plasma (Raymond & Smith 1977). The high fitted ion-
ization age is puzzling because it implies electron densi-
ties of ∼ 103 cm−3; densities which are as high as those
seen in the optical knots. However, visual inspection of
the spectrum for Region 4 (Fig. 4; right) shows that it is
clearly different from regions which do not exhibit ion-
ization equilibrium (i.e., Regions 1 and 3 for instance).
Morse et al. (2004) proposed that Cas A's ejecta con-
sists of cool, optically bright emission knots embedded in
a warm, tenuous ejecta medium which is responsible for
the observed X-ray emission. This general description is
supported by our model fits (Table 2). Were the ejecta
completely uniform, then a single temperature, single
ionization age would adequately describe the spectra in
all regions at each epochs. However, since we see that
the fitted temperature rises in Regions 1 -- 3, the rate at
which the ejecta is being shocked in these regions is not
constant due to density inhomogeneities in the preshock
material.
While the four regions which showed increases in in-
tensity appear to be associated with ejecta which has
recently crossed the reverse shock, Regions 5 and 6 show
flux decreases and therefore do not indicate reverse shock
locations. The interior projected thin filament seen in
Region 5 has a non-thermal spectrum and is likely as-
sociated with the forward and not the reverse shock
front (DeLaney et al. 2004). The observed changes in
its brightness probably reflects changes in path length
along interconnecting tangents of the wave front of the
forward shock as it interacts with the local circumstellar
medium on the near-facing side of the remnant.
In contrast, emission seen in Region 6 (Fig. 5, right)
which also showed a decrease in brightness, may repre-
sent ejecta cooling in the post-shock flow following re-
verse shock passage. Its projected location in the rem-
nant is consistent with this identification. While it lies at
approximately the same radial distance from the center
of expansion as the brightening Regions 1 -- 4, it lies just
behind (eastward) a patch of optically bright emission
ejecta along the remnant's southwestern limb.
If this
feature is indeed fading due to cooling, this would im-
ply it was shocked much earlier than in the other regions
in the area and would be consistent with its location in-
side the steady X-ray and optical emission features found
farther out from the remnant's center.
3.3. Scale Length of X-ray Emitting Knots
Rapid intensity changes in small regions throughout
the remnant offer some insight into the structure of the
lower density component of Cas A's SN ejecta. Abrupt
increases (or decreases) in the observed count rate
strongly suggest that the X-ray emitting ejecta is non-
uniform on fairly small scales. This conclusion has long
been clear from optical observations (Fesen et al. 2001)
but not from X-ray observations which are sensitive to
a much lower density, higher temperature component of
the ejecta but which likely dominate the remnant's over-
4
Patnaude & Fesen
all mass budget (Vink et al. 1996; Willingale et al. 2003;
Laming & Hwang 2003). The velocity of the freely ex-
panding ejecta clumps at the specific regions we investi-
gated is not known, but a rough estimate can be made
based on the distance of the reverse shock from the cen-
ter of expansion. The reverse shock is located rrs ∼ 95′′
(≃ 1.6 pc at a distance of 3.4 kpc) from the center of
expansion (Gotthelf et al. 2001) and unshocked, freely
expanding ejecta at this distance have been expanding
for tage ≃ 330 yr. Therefore, we estimate that the un-
shocked ejecta velocity is no more than vej ≈ rrs/tage ≈
5000 km s−1.
The observed intensity for the small features we stud-
ied show significant changes on time scales of ∆t ∼ 2 -- 4
yrs. This suggests a scale length for spatial variations
in the ejecta no larger than the distance covered by a
∼ 3000 km s−1 reverse shock moving through ∼ 5000
km s−1 expanding ejecta over the course of a few years;
i.e., ∆t × vej ≈ 0.02 − 0.03 pc or about 1 -- 2 arcsecond at
3.4 kpc. This scale length is consistent with the actual
sizes of the variable flux features we see in the remnant.
In contrast, the fine-scale structure of optical knots in
Cas A are of order ∼ 0.′′2−0.′′3 in size (Fesen et al. 2001),
roughly an order of magnitude smaller than the spatial
scale for variations seen in the clumpy component of the
X-ray bright ejecta.
4. CONCLUSIONS
With the exception of the rapidly evolving remnant as-
sociated with SN 1987A (Burrows et al. 2000; Park et al.
2004), no remnant has been previously reported to show
significant flux changes across portions of its X-ray emit-
ting structure. This is due in part to the limited number
of high-resolution images available on young SNRs cover-
ing a significant time span, and the relatively small num-
ber of very young Galactic or LMC/SMC SNRs known.
Inspection of X-ray data taken of the Cassiopeia A su-
pernova remnant with Chandra ACIS-S3 in 2000, 2002,
and 2004 reveals the presence of several small scale fea-
tures (≤ 10′′) which exhibit significant intensity changes
over this 4 year time frame. Here we described six such
features four of which had count rate increases from ∼
10% to over 90%, while two others showed decreases of
∼ 30% -- 40%. Whereas the majority of the six variable
flux features exhibited no gross spectral changes in the
1 -- 4 keV energy band during the four year period, one
region which had the greatest increase in brightness also
showed hints of slight emission line changes as well as an
increase in its 6 cm radio emission. Spectral fits using
non-equilibrium ionization, metal-rich plasma models in-
dicate increased or decreased electron temperatures for
features showing increasing or decreasing count rates, re-
spectively.
A fading emission filament projected near the rem-
nant's interior (Region 5) showed a completely nonther-
mal spectrum and thus may be associated with the for-
ward shock front. It decreasing intensity may be the re-
sult of changing path lengths along the edges of a wavy
forward shock as it interacts with the local,
inhomo-
geneous circumstellar medium. Based on the observed
count rate changes and the assumption of freely expand-
ing 5000 km s−1 ejecta at the reverse shock front, we
estimate the unshocked ejecta to have spatial scale vari-
ations of 0.02 − 0.03 pc, consistent with the X-ray emit-
ting ejecta belonging to a more diffuse component of the
supernova ejecta.
In general, bright X-ray and optical knots are not co-
incident. This is not unexpected given the differences in
their temperatures and densities. Moreover, the sizes of
the X-ray knots (∼ 1′′ − 5′′) appear to be an order of
magnitude larger than the optical knots (∼ 0.′′2 − 0.′′3;
Fesen et al. 2001). However, the remnant's optical knots
can also appear and brighten on timescales as short as a
few years, not unlike the timescales we observed for these
X-ray knots. Given the newly observed short term vari-
ability seen here in the X-ray, future X-ray and optical
observations of selected knots and filaments might pro-
vide valuable information regarding the dynamical and
radiative evolution of supernova ejecta knots over a range
of densities, sizes, and postshock temperatures.
Cassiopeia A is currently the youngest and one of the
most studied Galactic SNRs. The small yet noticeable
changes we found combined with future X-ray observa-
tions of this bright X-ray remnant may help shed light
not only on the remnant's near-term X-ray evolution, but
also help us better understand the fine-scale structure,
dynamics and chemistry of debris from core-collapse su-
pernovae in general.
The authors wish to thank Paul Plucinsky for help-
ful advice regarding the accuracy of the ACIS response
matrices and Tracey Delaney for discussions regarding
the emergence of radio knots coincident with the X-ray
knots. We would also like to thank the referee, Parviz
Ghavamian, for several valuable comments and sugges-
tions. This work was partially supported by the Chandra
X-ray Observatory through the Archival Research Award
AR4-5005X.
REFERENCES
Anderson, M. C., Jones, T. W., Rudnick, L., Tregillis, I. L., &
Kang, H. 1994, ApJ, 421, L31
Aschenbach, B. 1999, IAU Circ. 7249
Borkowski, K. J., Lyerly, W. J., & Reynolds, S. P. 2001, ApJ, 548,
820
Burrows, D. N., et al. 2000, ApJ, 543, L149
Chevalier, R. A., & Kirshner, R. P. 1978, ApJ, 219, 931
Chevalier, R. A., & Kirshner, R. P. 1979, ApJ, 233, 154
DeLaney, T., & Rudnick, L. 2003, ApJ, 589, 818
Delaney, T. private communication
DeLaney, T., Rudnick, L., Fesen, R. A., Jones, T. W., Petre, R.,
& Morse, J. A. 2004, ApJ, 613, 343
Douvion, T., Lagage, P. O., & Cesarsky, C. J. 1999, A&A, 352,
L111
Fesen, R. A., Morse, J. A., Chevalier, R. A., Borkowski, K. J.,
Gerardy, C. L., Lawrence, S. S., & van den Bergh, S. 2001, AJ,
122, 2644
Fesen, R. A., Hammell, M. C.,Morse, J., Chevalier, R. A.,
Borkowski, K. J., Dopita, M. A., Gerardy, C. L., Lawrence, S. S.,
Raymond, J. C., & van den Bergh, S. 2006a, ApJ, 645, 283
Fesen, R. A., Pavlov, G. G., & Sanwal, D. 2006b, ApJ, 636, 848
Gotthelf, E. V., Koralesky, B., Rudnick, L., Jones, T. W., Hwang,
U., & Petre, R. 2001, ApJ, 552, L39
Hughes, J. P., Rakowski, C. E., Burrows, D. N., & Slane, P. O.
2000, ApJ, 528, L109
Hwang, U., Holt, S. S., & Petre, R. 2000, ApJ, 537, L119
Hwang, U., & Laming, J. M. 2003, ApJ, 597, 362
Hwang, U., et al. 2004, ApJ, 615, L117
Small-Scale X-ray Variability of Cas A
5
TABLE 1
ACIS-S3 Count Rates
Region RA (J2000) Dec (J2000) Extraction
Normalized Count Ratea
Count Rate Change
(h:m:s.s)
(◦
′
′′)
Region
(2000.1)
(2002.1)
(2004.1b )
(∆2004−2000 )
1
2
3
4
5
6
23:23:34.0
23:23:24.4
23:23:13.4
23:23:11.2
23:23:26.9
23:23:18.0
+58:49:56.2
+58:47:13.2
+58:48:10.0
+58:48:53.2
+58:48:19.0
+58:48:37.0
5′′ × 4′′
3′′ × 3′′
5′′ × 3′′
5′′ × 5′′
3′′ × 10′′
4′′ × 4′′
1.84±0.02
0.48±0.02
0.93±0.01
2.06±0.02
1.56±0.02
1.04±0.01
2.04±0.01
0.57±0.01
1.28±0.02
2.24±0.02
1.21±0.02
0.83±0.01
2.09±0.01
0.93±0.01
1.10±0.02
2.51±0.01
0.95±0.02
0.71±0.02
+14%
+93%
+18%
+22%
−40%
−31%
aValues listed are detected counts s−1 normalized to Cas A's X-ray point source rate of 0.12±0.01 cts s−1.
bEpoch 2004 count rates are calculated from Chandra Observation ID 5196, which has the same roll angle as the 2000.1
and 2002.1 observations.
Spectral Fit Temperature Variations
TABLE 2
Region
1
2
3
4
6
2000.1
0.76+0.02
−0.03
· · ·
0.84+0.06
−0.05
0.94+0.02
−0.03
0.71+0.04
−0.03
kT (keV)
2002.1
0.86+0.09
−0.06
1.09+0.32
−0.28
1.43+0.07
−0.10
0.92+0.04
−0.02
0.63+0.02
−0.02
2004.1
1.05+0.12
−0.10
1.46+0.17
−0.12
1.45+0.08
−0.12
0.94+0.02
−0.05
0.55+0.02
−0.04
Kamper, K. & van den Bergh, S. 1976, ApJS, 32, 351
Laming, J. M., & Hwang, U. 2003, ApJ, 597, 347
Raymond, J. C., & Smith, B. W. 1977, ApJS, 35, 419
Reed, J. E., Hester, J. J., Fabian, A. C., & Winkler, P. F. 1995,
133, 403
Morse, J. A., Fesen, R. A., Chevalier, R. A., Borkowski, K. J.,
Gerardy, C. L., Lawrence, S. S., & van den Bergh, S. 2004, ApJ,
614, 727
Park, S., Zhekov, S. A., Burrows, D. N., Garmire, G. P., & McCray,
R. 2004, ApJ, 610, 275
Pavlov, G. G., & Zavlin, V. E. 1999, IAU Circ. 7270
Peimbert, M., & van den Bergh, S. 1971, ApJ, 167, 223
Plucinsky, P. P., et al. 2003, Proc. SPIE, 4851, 89
ApJ, 440, 706
Tananbaum, H. 1999, IAU Circ. 7246
Thorstensen, J. R., Fesen, R. A., & van den Bergh, S. 2001, AJ,
122, 297
Vink, J., Kaastra, J. S., & Bleeker, J. A. M. 1996, A&A, 307,
Willingale, R., Bleeker, J. A. M., van der Heyden, K. J., & Kaastra,
J. S. 2003, A&A, 398, 1021
6
Patnaude & Fesen
Fig. 1. -- A Chandra image of Cas A in 2004 (Observation ID 5196). The marked regions correspond to those listed in Table 1 and
shown in Figure 2.
Small-Scale X-ray Variability of Cas A
7
Fig. 2. -- Close-up of the six regions listed in Table 1 for each epoch. The boxes correspond to the spectral extraction regions used in
the fits in Table 2 and shown in Figures 3 -- 5.
8
Patnaude & Fesen
Fig. 3. -- Left: Spectral fits to Region 1 for 2000, 2002 and 2004. Right: Region 2 data for 2000 and spectral fits to Region 2 for 2002
and 2004.
Fig. 4. -- Left: Spectral fits to Region 3 for 2000, 2002 and 2004. Right: Same as left, for Region 4.
Small-Scale X-ray Variability of Cas A
9
Fig. 5. -- Left: Spectral fits to Region 5 for 2000, 2002, and 2004. Right: Same as left for Region 6.
|
0707.4419 | 1 | 0707 | 2007-07-30T13:57:20 | OVRO N2H+ Observations of Class 0 Protostars: Constraints on the Formation of Binary Stars | [
"astro-ph"
] | We present the results of an interferometric study of the N2H+(1--0) emission from nine nearby, isolated, low-mass protostellar cores, using the OVRO millimeter array. The main goal of this study is the kinematic characterization of the cores in terms of rotation, turbulence, and fragmentation. Eight of the nine objects have compact N2H+ cores with FWHM radii of 1200 -- 3500 AU, spatially coinciding with the thermal dust continuum emission. The only more evolved (Class I) object in the sample (CB 188) shows only faint and extended N2H+ emission. The mean N2H+ line width was found to be 0.37 km/s. Estimated virial masses range from 0.3 to 1.2 M_sun. We find that thermal and turbulent energy support are about equally important in these cores, while rotational support is negligible. The measured velocity gradients across the cores range from 6 to 24 km/s/pc. Assuming these gradients are produced by bulk rotation, we find that the specific angular momenta of the observed Class 0 protostellar cores are intermediate between those of dense (prestellar) molecular cloud cores and the orbital angular momenta of wide PMS binary systems. There appears to be no evolution (decrease) of angular momentum from the smallest prestellar cores via protostellar cores to wide PMS binary systems. In the context that most protostellar cores are assumed to fragment and form binary stars, this means that most of the angular momentum contained in the collapse region is transformed into orbital angular momentum of the resulting stellar binary systems. | astro-ph | astro-ph |
OVRO N2H+ Observations of Class 0 Protostars: Constraints on
the Formation of Binary Stars
Xuepeng Chen, Ralf Launhardt, and Thomas Henning
Max Planck Institute for Astronomy, Konigstuhl 17, D-69117 Heidelberg, Germany;
[email protected]
ABSTRACT
We present the results of an interferometric study of the N2H+ (1 -- 0) emission
from nine nearby, isolated, low-mass protostellar cores, using the OVRO millime-
ter array. The main goal of this study is the kinematic characterization of the
cores in terms of rotation, turbulence, and fragmentation. Eight of the nine ob-
jects have compact N2H+ cores with FWHM radii of 1200 − 3500 AU, spatially
coinciding with the thermal dust continuum emission. The only more evolved
(Class I) object in the sample (CB 188) shows only faint and extended N2H+
emission. The mean N2H+ line width was found to be 0.37 km s−1. Estimated
virial masses range from 0.3 to 1.2 M(cid:12). We find that thermal and turbulent en-
ergy support are about equally important in these cores, while rotational support
is negligible. The measured velocity gradients across the cores range from 6 to
24 km s−1 pc−1. Assuming these gradients are produced by bulk rotation, we find
that the specific angular momenta of the observed Class 0 protostellar cores are
intermediate between those of dense (prestellar) molecular cloud cores and the
orbital angular momenta of wide PMS binary systems. There appears to be no
evolution (decrease) of angular momentum from the smallest prestellar cores via
protostellar cores to wide PMS binary systems. In the context that most pro-
tostellar cores are assumed to fragment and form binary stars, this means that
most of the angular momentum contained in the collapse region is transformed
into orbital angular momentum of the resulting stellar binary systems.
Subject headings:
ISM: globules -- ISM: individual (CB 68, CB 188, CB 224,
CB 230, CB 244, IRAS 03282+3035, IRAS 04166+2706, L 723 VLA2, RNO 43) --
ISM: kinematics and dynamics -- ISM: molecules -- stars: formation -- stars:
millimeter
-- 2 --
1.
INTRODUCTION
A major gap in our understanding of star formation concerns the origin of binary stars.
Binary/multiple systems appear to be the preferred outcome of the star formation process,
but at present we do not understand how this occurs (Mathieu et al. 2000). In the past two
decades, statistical properties of binary stars are gradually comprehended through numerous
observations and theoretical simulations, but are still subject to extensive ongoing studies
(see a review by Duchene et al. 2007). The most recent observational results and theoretical
models for binary star formation can be found in the proceedings of IAU Symposium 200
(Zinnecker & Mathieu 2001), Launhardt (2004), and reviews in Protostars & Protoplanets
V (Reipurth, Jewitt, & Keil 2007).
Different formation scenarios for binary stars have been proposed, of which the classical
ideas of capture and fission are no longer considered major processes. Fragmentation of
rotating cloud cores with initially flat density profiles, immediately after a phase of free-fall
collapse, is generally considered to be the most efficient mechanism, leading to systems with
a wide variety of properties (see reviews by Bodenheimer et al. 2000 and Tohline 2002).
These properties are largely determined by the accretion process which, in turn, strongly
depends on the initial conditions of the cloud cores, e.g., the initial distributions of mass
and angular momentum (see Bate & Bonnell 1997). Some of the theoretical predictions,
e.g., that close binary systems are likely to have mass ratios near unity (see Bate 2000), are
already indirectly supported by statistical studies of evolved binary systems (see Halbwachs
et al. 2003).
However, our current knowledge on the formation of binary stars mainly relies on obser-
vations of main sequence (MS) and pre-main sequence (PMS) stars and the constraints they
put on the models. The observational link between initial conditions in a molecular cloud
and the final star systems formed therein is still missing. Furthermore, as multiple systems
certainly undergo dynamical evolution, important information about the formation phase is
lost in the final systems. Direct observations of prestellar cores and protostars are needed
therefore to answer a number of questions, e.g., how common is binarity/multiplicity in the
protostellar phase? What makes a prestellar core to fragment in the collapse phase? How
is angular momentum distributed? Are there differences between cores forming binaries and
those forming single stars? Unfortunately, direct observations of protostellar stages, when
the main collapse has started but no optical or infrared emission emerges from the protostar
through the opaque infalling envelope, were long hampered by the low angular resolution of
millimeter (mm) telescopes and the results were mostly interpreted in terms of single star
formation. Only the recent advance of large mm interferometers has enabled us to directly
observe the formation phase of binary stars, although the number of known systems is still
-- 3 --
very small and no systematic observational studies of the initial fragmentation process do
exist yet (see the review by Launhardt 2004). The systematic study by Looney et al. (2000)
was successful in detecting a number of protostellar binaries by mm dust continuum emis-
sion, but did not provide any kinematic information. Only recently these authors published
velocity fields of 3 objects from their sample (Volgenau et al. 2006).
To search for binary protostars and derive kinematic properties of these systems, we
have started a program to observe, at high angular resolution, a number of isolated low-mass
prestellar and protostellar molecular cloud cores, conducted at the Owens Valley Radio Ob-
servatory (OVRO) millimeter array (this work; hereafter called Paper I) and now continued
with ATCA (Australia Telescope Compact Array) and PdBI (IRAM Plateau de Bure In-
terferometer) arrays (Paper II&III, in prep.). In this paper we present N2H+ results for 9
protostellar cores observed with OVRO. In section 2 we describe the target list, observations,
and data reduction. Observational results are presented in section 3. We give a detailed de-
scription of individual sources and discuss the implications of our results for binary star
formation models in section 4. The main conclusions of this study are summarized in section
5.
2. OBSERVATIONS AND DATA REDUCTION
For this survey we selected a number of well-isolated and nearby low-mass protostellar
cores. Priority was given to sources which were already well-characterized either by our own
previous observational data or in the literature. Most sources are Class 0 objects, which
represent the youngest protostars at an age of a few × 104 yr. While they may already be
too evolved to represent the true initial conditions, these sources provide an opportunity to
probe the earliest and most active stage of the star formation process, where most of the
initial information is still preserved. The target list and basic properties of the sources are
summarized in Table 1.
For studying the gas kinematics and derive rotation curves, we choose to observe the
N2H+ (1-0) hyperfine structure line complex at 93.1378 GHz. N2H+ is known to be a se-
lective tracer of cold, dense, and quiescent gas and is particularly suitable for studying the
structure and kinematics of cold star-forming cores (Turner & Thaddeus 1977; Womack,
Ziurys, & Wychoff 1992; Bachiller 1996; Caselli et al. 2002). It is the most reliable tracer
of the gas kinematics in pre- and protostellar cores for three reasons: (1) compared to other
molecules, it depletes much later and more slowly onto grains (Bergin & Langer 1997), (2)
it is formed where CO is depleted and thus traces only the dense cores and not the wider
envelope and is thus perfectly suited for interferometric observations, and (3) with its seven
-- 4 --
hyperfine components within 17 km s−1 (including optically thin and moderately optically
thick components; Caselli et al. 1995) it provides much more precise and reliable kinematic
information than a single line, even with a moderate signal-noise ratio.
Observations were carried out with the OVRO array of six 10.4 m telescopes during 7
observing seasons between 1999 and 2002. Five different array configurations (C, L, E, H,
and U) were used, with baselines ranging from 18 to 480 m. All antennas were equipped with
cooled SIS receivers which provided average system temperatures of ∼ 300−400 K at the
observing frequency. A digital correlator was centered at 93.1378 GHz. Spectral resolution
and bandwidth were ∼ 0.2 km s−1 and 25 km s−1, respectively. Amplitude and phase were
calibrated through frequent observations of quasars nearby to each source, typically every
20 minutes, resulting in an absolute position uncertainty of ≤ 0.(cid:48)(cid:48)2. The flux density scale
was calibrated by observing Neptune and Uranus. The estimated uncertainty is < 20%.
Observing parameters are summarized in Table 2. The thermal dust continuum emission
was measured simultaneously with the N2H+ and other line observations using a separate
continuum correlator. The combined continuum results will be published in a separate paper
(Launhardt et al., in prep.).
The raw data were edited and calibrated using the MMA software package (Scoville et
al. 1993), and synthesized images were produced using MIRIAD (Sault et al. 1995) and
its CLEAN algorithm, with "robust" weighting of the visibilities (Briggs et al. 1999). The
cleaned and restored maps have effective synthesized beam sizes of 4−8(cid:48)(cid:48) and 1σ rms levels
of ∼ 70 mJy/beam (see Table 2). Further analysis and figures were done with the GILDAS1
software package. With Class (part of GILDAS), we have developed a semi-automatic fitting
routine which allows the derivation of reliable and very accurate velocity fields from the 7-
hyperfine component line complex. Assuming that bulk rotation is the dominating motion,
the velocity fields are then used to derive the rotation axis and the specific angular momentum
of the cores.
3. RESULTS
3.1. Morphology of N2H+ Cores
N2H+ emission is detected from all nine targeted objects. Figure 1 shows the distribu-
tion of the velocity-integrated intensity of N2H+ towards the nine cores. The emission was
integrated over all seven components, using frequency masks that completely cover velocity
1http://www.iram.fr/IRAMFR/GILDAS
-- 5 --
gradients within the sources. All objects are spatially associated with mm dust continuum
sources, indicative of embedded protostars and their accretion disks. The positions of the
mm continuum sources are indicated by crosses in Figure 1 (Launhardt et al. 2007, in prep.).
We find that in seven of the nine objects the mm continuum source lies within the half max-
imum level of the N2H+ emission. The two exceptions are CB 188 (the only Class I object
in our sample) and CB 244. This good general agreement indicates that N2H+ cannot be
significantly depleted like, e.g., CO and CS (see also Bergin et al. 2001; Caselli et al. 1999).
We also measured mean FWHM radii of the integrated N2H+ emission. The mean
FWHM core radii R were measured as A1/2/π, where A is the core area at the half maximum
level, corrected for the beam size. Except for CB 188, all sources exhibit quite compact N2H+
emission regions with mean FWHM radii of 1200−3500 AU (see Table 4). The average radius
in our sample is (cid:104)R(cid:105) = 2000 ± 800 AU, which is much lower than the average value found
by Caselli et al. (2002; hereafter CBMT02) for their sample of starless cores (∼ 10000 AU)
observed with single-dish observations (beam size ∼ 54(cid:48)(cid:48)), but similar to the radius of Class
0 sources IRAM 04191+1522 (∼ 2400 AU) and NGC 1333 IRAS 4B (∼ 1800 AU) observed
with PdBI by Belloche et al. (2002) and Di Francesco et al. (2001), respectively.
However, when viewed in detail, all sources in our sample show a complex, often multi-
peaked structure. For example, IRAS 03282, IRAS 04166, and CB 230 each have two sep-
arated peaks in the region enclosed by the half maximum intensity level, with the mm
continuum source located between the two peaks. In RNO 43, CB 68, CB 224, and CB 244
the main peak of N2H+ emission offsets by 5−10(cid:48)(cid:48) from the mm continuum sources. Only
in L723 VLA2 there is no positional discrepancy between the N2H+ and mm continuum
emission. The individual sources are discussed in detail in Section 4.1.
3.2. Masses and Column Densities
Figure 2 shows the N2H+ spectra at the position of maximum intensity in each map.
The N2H+ (1−0) line complex consists of 7 hyperfine structure components, which have
been detected in all sources. However, in several sources like, e.g., L723 VLA2 (see below),
the line width is larger than the separation between hyperfine components, so that some
lines are blended. The hyperfine fitting program in CLASS (Forveille et al. 1989), with
the frequencies adopted from Caselli et al. (1995) and weights adopted from Womack et
(1992), has been used to determine LSR velocities (VLSR), intrinsic line width ((cid:52)v;
al.
corrected for instrumental effects), total optical depths (τtot), and excitation temperatures
(Tex). These parameters are listed in Table 3. Here τtot is the sum of the peak optical depths
of the seven hyperfine components (see Benson & Myers 1989). The optical depth of the
-- 6 --
main N2H+ (J F1 F = 1 2 3 −→ 0 1 2) component, which is equal to 0.259τtot, is found to
be small (≤ 0.5) at the intensity peak for all sources. Hence the N2H+ emission can be
considered optically thin everywhere. The excitation temperature, Tex, was calculated to be
4.1−5.7 K at the peak positions, using a main-beam efficiency ηB = 0.7 (Padin et al. 1991).
The average (cid:104)Tex(cid:105) ∼ 4.9 K in our sample is similar to what has been found with single-dish
observations for dense cores by CBMT02 (∼ 5.0 K).
Assuming that the observed N2H+ line widths are not dominated by systematic gas
motions, the virial mass of the cores has been calculated as:
5
R(cid:52)v2
m
αvirG
,
8ln2
Mvir =
(1)
where G is the gravitational constant, R is the FWHM core radius, and (cid:52)vm is the line
width of the emission from an "average" particle with mass mave = 2.33 amu (assuming
gas with 90% H2 and 10% He). The coefficient αvir = (3 − p)/(5 − 2p), where p is the
power-law index of the density profile, is corrected for deviations from constant density (see
Williams et al. 1994). In our calculations, we assume p = 1.5 and αvir = 0.75 (see Andr´e,
Ward-Thompson, & Barsony 2000). (cid:52)vm can be derived from the observed spectra by
(cid:52)v2
m = (cid:52)v2
kTex
mH
1
mav
− 1
mobs
(
),
obs + 8ln2
(2)
where (cid:52)vobs is the observed mean line width (obtained through Gaussian fitting to the
distribution of line widths vs. solid angle area; see Table 3 and §3.4) and mobs is the mass of
the emitting molecule (here we use mN2H + = 29 amu). The corresponding hydrogen density
of the nH2 has been calculated assuming a uniform density spherical core with radius R
(given in Table 4). The derived virial masses Mvir in our sample range from 0.3 to 1.2 M(cid:12),
with a mean value of 0.6 M(cid:12). The corresponding hydrogen densities nH2 range from 7.4 to
81 × 105 cm−3. Both Mvir and nH2 values for each source are listed in Table 4.
The N2H+ column density has been calculated independently from the line intensity
using the equation given by Benson et al. (1998):
τ(cid:52)vTex
(cm−2),
N (N2H +) = 3.3 × 1011
1 − e−4.47/Tex
(3)
where τ is the total optical depth, (cid:52)v is the intrinsic line width in km s−1, and Tex is the
excitation temperature in K. The gas-phase N2H+ mass of the core can then be calculated
from MN2H + ≈ N (N2H +)peak × mN2H + × d2 × ΩF W HM , where d is the distance from the
sun and ΩF W HM is the area enclosed by the contour level at 50% of the peak value for each
core.
-- 7 --
From the ratio of N2H+ mass to virial mass, we derived the average fractional abundance
of N2H+ in each core: X(N2H +) = MN2H +/MH2, where MH2 = Mvir/1.36, the factor 1.36
accounting for He and heavier elements. The values are listed in Table 4. The average value
(cid:104)X(N2H +)(cid:105) ∼ 3.3 × 10−10 in our sample is close to that found by CBMT02 for their sample
of dense cores (∼ 3 × 10−10).
3.3. Velocity Fields
Based on the hyperfine fitting program in Class, we have developed a semi-automated
quality control and spectra fitting routine that computes the mean radial velocity, line width,
and line intensity at each point of the map where N2H+ is detected (> 2σ noise). Figure 3
shows the mean velocity fields for the 8 Class 0 objects in our sample, obtained from this
line fitting routine. The N2H+ emission from CB 188, the only evolved (Class I) object in
our sample, is too faint and fuzzy to give a reliable velocity field.
In Class 0 protostars, the effects of infall, outflow, rotation, and turbulence are generally
superimposed. Of these, turbulence and infall normally broaden the lines but do not produce
systematic velocity gradients. On the other hand, systematic velocity gradients are usually
dominated by either rotation or outflow.
In Fig. 3, we therefore also show the outflow
information for each source. We want to mention that many other studies usually do not
take this kind of caution. Figure 3 shows that 5 objects (IRAS 04166, RNO 43, L723 VLA2,
CB 230, and CB 224) have well-ordered velocity fields with symmetrical gradients, while 3
objects (IRAS 03282, CB 68, and CB 244) have more complex velocity fields. Two objects
(RNO 43 and CB 230) with well-ordered velocity fields have gradients roughly perpendicular
to the axis of outflow, while 2 objects (IRAS 04166 and L732 VLA2) have gradients basically
parallel and one object (CB 224) anti-parallel to the outflow direction.
Assuming the mean direction of the bulk angular momentum is preserved in the collapse
from the core to the disk, we would expect outflows to emerge perpendicular to the rotation
velocity gradients. Thus, if the velocity field is dominated by rotation, we would see the
velocity gradient to be perpendicular to the outflow axis, like seen in RNO 43 and CB 230.
If gradients are parallel to the outflow axis and with the same orientation, it is likely that we
see outflow motions rather than rotation, like in IRAS 04166 and L723 VLA2. In these cases,
a possible underlying rotation velocity gradient must be smaller than the observed effective
gradient. Following these arguments, we treat those gradients parallel to the outflow as
upper limits to the rotation gradient. The details of the velocity field for each source are
described in §4.1.
-- 8 --
A least-squares fitting of velocity gradients has been performed for the objects in our
sample using the routine described in Goodman et al.
(1993; hereafter GBFM93). The
fitting results are summarized in Table 5. Listed are in column (2) the mean velocity of the
cores, in columns (3) and (4) the magnitude of the velocity gradient g and its direction Θg
(the direction of increasing velocity, measured east of north), and in column (5) the total
velocity shift across the core gr (i.e., the product between g and core size R).
3.4. Line Widths
Figure 4 shows the distribution of line widths vs. solid angle area in the maps for the
9 protostars in our sample. The mean line width for each source is then derived through
Gaussian fitting to the distribution and is listed in Table 3. We find that the mean line
width in our sample is 0.29 − 0.51 km s−1, with an average of ∼ 0.37 km s−1.
The FWHM thermal line width for a gas in LTE at kinetic temperature TK is given by
(cid:52)v2
th = 8ln2
kTK
mobs
,
(4)
width ((cid:52)vN T = (cid:112)(cid:52)v2
where k is the Boltzmann constant and mobs is the mass of the observed molecule. Assuming
a kinetic gas temperature of 10 K (see Benson & Myers 1989), the thermal contribution to
the N2H+ line width is ∼ 0.13 km s−1. The typical non-thermal contribution to the line
th) is then ∼ 0.35 km s−1 , which is about 2.5 times larger
than the thermal line width (see discussion in §4.2).
obs − (cid:52)v2
Figure 5 shows the spatial distribution of the N2H+ line width for those 6 Class 0 sources
which are not highly elongated (axial ratios ≤ 2). In most cases the line widths are roughly
constant within the interiors of the cores and broader line widths occur only at the edges.
This is consistent with both the NH3 observations by Barranco & Goodman (1998) and the
physical picture described by Goodman et al. (1998) that the star-forming dense cores are
"velocity coherent" regions of nearly constant line width. The exception, CB 230, is discussed
in §4.1.8
-- 9 --
4. DISCUSSION
4.1. Description of Individual Sources
4.1.1.
IRAS 03282+3035
IRAS 03282 is one of the youngest known Class 0 protostars (Andr´e et al. 2000). It is
located in the western part of the Perseus molecular cloud complex at a distance of ∼ 300 pc
(Motte & Andr´e 2001). Its bolometric luminosity and total envelope mass are estimated to
be Lbol ∼ 1.2 L(cid:12) and Menv ∼ 2.9 M(cid:12), respectively (Shirley et al. 2000). IRAS 03282 drives
a highly collimated bipolar molecular outflow at P.A. ∼ −37◦ with kinematic age ∼ 104 yr
(Bachiller et al. 1994).
The N2H+ intensity map of IRAS 03282 shows a molecular cloud core that is hour-glass
shaped and elongated along P.A. ∼ 30 degree, i.e., nearly perpendicular to the axis of the
large-scale CO outflow (see Fig. 3). We interpret the hourglass shape as effect of the outflow,
which re-releases CO from grain surface back into the gas phase. CO, in turn, destroys the
N2H+ molecule (Aikawa et al. 2001). Our OVRO 1.3 mm dust continuum images reveal two
mm sources located between the two N2H+ emission peaks (see Fig. 3). The mass ratio and
the angular separation of this binary protostar are 0.23 and 1.(cid:48)(cid:48)5 (∼ 450 AU at a distance of
300 pc), respectively (Launhardt et al. 2007, in prep.; hereafter LSZ07). The velocity field
of IRAS 03282 does not show a symmetrical gradient and the red-shifted N2H+ emission lobe
to the northwest of the mm sources matches exactly the red-shifted 13CO lobe, implying that
the N2H+ emission in this region is affected by the outflow. Correspondingly, we treat the
observed total velocity gradient as upper limit to rotation (see §3.3). Figure 5 shows the line
width distribution of IRAS 03282. The average line width across the source is ∼ 0.4 ± 0.1
km s−1 and reaches 0.7 km s−1 towards the southern edge of the core.
4.1.2.
IRAS 04166+2706
IRAS 04166 is a Class 0 protostar associated with the small dark cloud B 213 in the
Taurus molecular cloud complex at a distance ∼ 140 pc (Mardones et al. 1997). Its bolo-
metric luminosity and total envelope mass are Lbol ∼ 0.4 L(cid:12) and Menv ∼ 1.0 M(cid:12), respectively
(Shirley et al. 2000). IRAS 04166 drives a highly collimated, extremely high velocity (up to
50 km s−1) bipolar molecular outflow at P.A. ∼ 30◦ (Tafalla et al. 2004).
The N2H+ intensity map of IRAS 04166 shows the same hourglass shape and orientation
relative to the outflow as IRAS 03282. We interpret it in the same way. The N2H+ velocity
-- 10 --
field of IRAS 04166 shows a symmetric gradient, which however matches the red- and blue-
shifted CO outflow lobes (see Fig. 3). As for IRAS 03282, we treat the observed total velocity
gradient as upper limit to rotation.
4.1.3. RNO 43 MM
The millimeter continuum source RNO 43 MM is associated with the dark cloud L 1582B
and the very cold IRAS point source 05295+1247 (detected only at 60 and 100 µm). The
cloud is physically connected to the λ Ori molecular ring located at a distance of ∼ 400 pc
(Zinnecker et al. 1992). RNO 43 MM is the origin of a 3.4 pc long Herbig-Haro flow (HH 243,
HH 244, and HH 245; Reipurth et al. 1997) and drives a collimated bipolar molecular outflow
at P.A. ∼ 60◦ (Arce & Sargent 2004). From its spectral energy distribution (SED) and the
ratio of sub-mm to bolometric luminosity, RNO 43MM has been classified a Class 0 source
(Bachiller 1996). Its bolometric luminosity and total envelope mass are estimated to be Lbol
∼ 6.0 L(cid:12) and Menv ∼ 0.4 M(cid:12), respectively (Zinnecker et al. 1992).
Our N2H+ intensity map of RNO 43MM shows a V-shaped core with two lobes, extend-
ing ∼ 20(cid:48)(cid:48) to the north and west, respectively (see Fig. 1). The western lobe exhibits much
higher (redder) mean velocities than the rest of the core, and is outside the plot scale of
Fig. 3 (11 up to 20 km s−1). This jump in velocity probably means that the western lobe
belongs to another molecular cloud layer in this direction. In the inner core region (included
at 50% contour level), the velocity field exhibits a symmetric structure with a gradient of
∼ 5.8 km s−1 pc−1 at P.A. ∼ −13◦, approximately perpendicular to the axis of outflow,
suggesting rotation (see Fig. 3). The total velocity shift across the inner core is ∼ 0.4 km
s−1.
4.1.4. CB 68
CB 68 (LDN 146) is a small Bok globule located in the outskirts of the ρ Oph dark cloud
complex at a distance of 160 pc (Clemens & Barvainis 1988; Launhardt & Henning 1997).
The dense core of the globule exhibits strong, extended, centrally peaked sub-mm/mm dust
continuum emission (Launhardt & Henning 1997; Launhardt et al. 1998; Vall´ee, Bastien, &
Greaves 2000) and is associated with the cold IRAS point source 16544 -- 1604. The central
source, which was classified as a Class 0 protostar, drives a weak, but strongly collimated
bipolar molecular outflow at P.A. ∼ 142◦ (Wu et al. 1996; Mardones et al. 1997; Vall´ee et
al. 2000).
-- 11 --
Our N2H+ intensity map shows a compact source of about 9(cid:48)(cid:48) FWHM radius, which
peaks very close to the mm continuum position.
In addition, it exhibits two armlike ex-
tensions to northeast and southwest. The velocity field of CB 68 is relatively complicated
and shows no systematic gradient which could be interpreted as rotation. There is also no
clear correlation with the outflow. As for IRAS 03282, we derive only an upper limit for the
rotation velocity gradient.
4.1.5. L723 VLA2
L723 is a small isolated dark cloud located at a distance of 300±150 pc (Goldsmith et
al. 1984). The IRAS point source at the center of the cloud core is associated with strong
far-infrared (Davidson 1987), sub-mm (Shirley et al. 2000) and mm continuum emission
(Cabrit & Andr´e 1991; Reipurth et al. 1993) and was classified as Class 0 protostar with
a bolometric luminosity Lbol ∼ 3.0 L(cid:12) and a circumstellar mass Menv ∼ 1.2 M(cid:12) (LSZ07).
Anglada et al. (1991) detected two radio continuum sources (VLA1 and VLA2) with 15(cid:48)(cid:48)
separation, both located within the error ellipse of the IRAS position. However, only VLA2
was found to be associated with dense gas in the cloud core (Girart et al. 1997). Centered
at the IRAS position is a large quadrupolar molecular outflow with two well-separated pairs
of red and blue lobes: a larger pair at P.A. ∼ 110◦ and a smaller pair at P.A. ∼ 30◦ (Lee
et al. 2002 and references therein). While several scenarios have been proposed to explain
the quadrupolar morphology of the large-scale CO outflow, the discovery of a thermal radio
jet at the position of VLA2 (Anglada et al. 1996) and of a large-scale H2S(1) bipolar flow
(Palacios & Eiroa 1999), both aligned with the larger pair of CO outflow lobes at P.A. ∼
110◦, as well as new high-resolution CO observations (Lee et al. 2002) clearly favor the
presence of two independent flows. Indeed, our OVRO mm continuum observations reveal
two compact sources in L723 VLA2 separated by ∼ 3.(cid:48)(cid:48)2 (960 AU at a distance of 300 pc),
supporting strongly the scenario that the quadrupolar outflow is driven by a binary protostar
system (see Launhardt 2004).
The N2H+ intensity map of L723 VLA2 shows a compact source of about 8(cid:48)(cid:48) FWHM
diameter, which peaks between the two mm continuum sources. In addition, it shows a long
extension to the northwest along the direction of larger outflow. The mean velocity difference
between the two continuum positions is ∼ 0.4 km s−1. However, the overall velocity field
of L723 VLA2 shows that the red-shifted N2H+ emission matches exactly the red-shifted
emission of two outflows and the velocity gradient is basically in the same direction as
the outflow, suggesting that the outflow has a strong effect on the N2H+ emission. As for
IRAS 03282, we therefore treat the observed total velocity gradient as upper limit to rotation.
-- 12 --
The line widths distribution of L723 VLA2 is shown in Fig. 5. The line widths are about
0.5 km s−1 in the northwest extension, 0.6−1.0 km s−1 across the core, and reach 1.5 km s−1
towards the eastern edge of the core. We suggest that the relatively large line widths in this
source are the result of strong outflow-envelope interaction.
4.1.6. CB 188
CB 188 is a small dark globule at a distance of ∼ 300 pc, which harbors a Class I YSO
(Launhardt & Henning 1997). A small (≤ 2(cid:48)) CO outflow with overlapping red and blue
lobes was detected by Yun & Clemens (1994), suggesting the YSO is seen close to pole-on. In
our N2H+ intensity map, CB 188 shows a complex clumpy structure (see Fig. 1). It also has
the weakest N2H+ emission in our survey (also see the spectra in Fig. 2). This could imply
that N2H+ is destroyed during the evolution from Class 0 to Class I, e.g., due to release of
CO from grain surfaces in outflows (see Aikawa et al. 2001). We do not further discuss this
source in this paper.
4.1.7. CB 224
CB 224 is a Bok globule located at a distance of ∼ 400 pc (Launhardt & Henning
1997). Two mm sources were detected in our OVRO survey at an angular separation of 20(cid:48)(cid:48)
(LSZ07). The northeast mm source (not shown here) is associated with a cold IRAS source
20355+6343 (3.9 L(cid:12); Launhardt & Henning 1997), but has no N2H+ emission detected. The
southwest source shown in our images, which is classified a Class 0 object (LSZ07), drives a
collimated 13CO bipolar outflow at P.A. ∼ −120◦ (Chen et al. 2007, in prep.).
The N2H+ intensity map of CB 224 shows the same hourglass morphology perpendicular
to the outflow as IRAS 03282 and IRAS 04166 (but slightly more asystematic in intensity).
There is a clear and systematic velocity gradient across the core, but the lines of constant
velocity are curved in a "C" shape (see Fig. 3). The velocity field cannot be solely interpreted
by rotation. The CO outflow seems to have no effect on the N2H+ velocity field because
it is oriented in the opposite direction. We speculate that the velocity field is due to a
combination of rotation and core contraction. The HCO+ observations in De Vries et al.
(2002) indeed indicated signs of infall motions in CB 224.
-- 13 --
4.1.8. CB 230
CB 230 (L 1177) is a small, bright-rimmed Bok globule at a distance of 400±100 pc (Wolf
et al. 2003). The globule contains a protostellar core of total mass ∼ 5M(cid:12) and exhibits
signatures of mass infall (Launhardt et al. 1997, 1998, 2001). Magnetic field strength and
projected direction in the dense core are B = 218 ± 50 mG and P.A. = −67◦, respectively
(Wolf et al. 2003). The dense core is associated with a large-scale collimated CO outflow at
P.A. = 7◦ of dynamical age ∼ 2 × 104 yr (Yun & Clemens 1994; Chen et al. 2007, in prep.).
The Mid-IR image of CB 230 suggests the presence of two deeply embedded YSOs separated
by ∼10(cid:48)(cid:48) (Launhardt 2004). Only the western source was detected at 1.3 mm and 3 mm dust
continuum, suggesting that the mass of a possible accretion disk around eastern source is
below the detection limit. Two bright near-infrared reflection nebulae are associated with
the embedded YSOs, but the stars are not directly detected at wavelengths shorter than 5
µm. CB 230 is probably a transition object between Class 0 and I.
Our N2H+ intensity map shows that the molecular cloud core is elongated East-West.
The velocity field map shows a clear velocity gradient across the core of ∼ 8.8 km s−1 pc−1
increasing from east to west along the long axis, i.e., roughly perpendicular to the outflow
axis. This strongly supports the view that the two MIR sources form a protobinary system
which is embedded in the N2H+ core. This core in turn rotates about an axis perpendicular
to the connecting line between the two protostars and parallel to the main outflow. Fig. 5
shows that the line width distribution exhibits a strong peak at the position of the 3 mm
continuum source. Together with the velocity field shown in Fig. 3, this can be understood as
the result of Keplerian rotation. Morphology and kinematics of this source will be discussed
more detailed in another paper.
4.1.9. CB 244
CB 244 (L 1262) is a Bok globule located at a distance of ∼ 180 pc (Launhardt & Henning
1997). It is associated with a faint NIR reflection nebula and a cold IRAS point source, and
was classified as Class 0 protostar. Its bolometric luminosity and total envelope mass are
estimated to be Lbol ∼ 1.1 L(cid:12) and Menv ∼ 3.3 M(cid:12), respectively (Launhardt & Henning 1997).
CB 244 drives a bipolar molecular outflow at P.A. ∼ −130◦ (Yun & Clemens 1994; Chen et
al. 2007, in prep.).
The N2H+ intensity map of CB 244 shows an elongated structure in the direction from
Northwest to Southeast, approximately perpendicular to the direction of outflow (see Fig. 3).
The velocity field of CB 244 exhibits a complicated structure and maybe dominated by effects
-- 14 --
other than rotation. As for other sources, we assume that the total observed velocity gradient
puts an upper limit to the rotation. The distribution of line widths for this source shows
two distinct peaks (see Fig. 4). The smaller line widths originate from the southeastern
part of the core where the protostar is embedded, while the larger widths are found in the
northwestern extension only (see Fig. 1). Therefore, we adopted the smaller peak value as
representative for CB 244.
4.2. Turbulent Motions
At a kinetic gas temperature of 10 K, the typical non-thermal line width in our sample
(0.35 km s−1) is about 2.5 times lager than the thermal line width (0.13 km s−1) (see §3.4).
The origin of the non-thermal line width in such cores is subject of an ongoing debate in the
literature. Generally, turbulence is suggested to be the main contribution (see e.g., Barranco
& Goodman 1998 and Goodman et al. 1998), but infall, outflow, and rotation may also
contribute to the non-thermal line width.
With the exception of L723 VLA2, there appears to be no spatial correlation between
regions of increased line width and outflow features (see Fig. 5). CB 230 may represent a
special case where Keplerian rotation of a large circumstellar disk causes the large non-
thermal line width in the central region (see Fig. 5). To avoid a bias from localized line-
broadening due to outflows and/or Keplerian rotation, we estimated the observed mean line
width through Gaussian fitting to the distribution of line widths vs. area in the maps (see
Fig. 4). The non-thermal contribution to these mean line widths are listed in Table 3. We
assume that these mean non-thermal line widths are dominated by turbulence. The thermal
FWHM line width of an "average" particle of mass 2.33 mH, which represents the local sound
speed, would be ∼ 0.44 km s−1 at 10 K. The mean observed non-thermal line width is ∼ 1.3
times smaller than this value, which means that turbulence in these cores is subsonic.
Figure 6 shows the distribution of non-thermal line width (cid:52)vN T with core size R for
the objects in our sample, together with the dense cores from GBFM93 and CBMT02. It
shows that high-level (supersonic) turbulence normally occurs in large-scale cores only. In
the cores traced by NH3 (R > 20000 AU ∼ 0.1 pc), the non-thermal line widths decrease with
core size with a power-low index ∼ 0.2, while in those traced by N2H+ (CBMT02; 5000 AU
< R < 20000 AU), the line widths decrease with an index of ∼ 0.5. This suggests that non-
thermal motions are more quickly damped in smaller cores (see also Fuller & Myers 1992).
Comparing our data with CBMT02, we find that the relation between line width and core
size no longer holds at R < 10000 AU, suggesting the inner parts of dense cores are "velocity
coherent". The radius at which the gas becomes "coherent" is less than 0.1 pc, as suggested
-- 15 --
in Goodman et al. (1998). Also note that the mean non-thermal line width in our sample
(0.35 km s−1) is even larger than the widths in some larger scale dense cores (see Fig. 6). We
speculate that the heating from an internal protostar, as well as related activities, like e.g.,
infall and/or outflow, contribute to the line widths in these protostellar cores.
4.3. Systematic Gas Motions: Fast Rotation of Protostellar Cores
For the fragmentation model of binary star formation, some initial angular momentum
must be present; otherwise the cores will collapse onto a single star. The source of this
angular momentum is generally suggested to be the bulk rotation of the core. As suggested
in the earlier study by GBFM93, rotation is a common feature of dense cores in molecular
clouds. In our observations, most objects show well-ordered velocity fields with symmetric
gradients and the motions of several objects could be interpreted as bulk rotation. The
preliminary results from our ATCA and PdBI observations show similar well-ordered velocity
fields (Paper II&III, in prep.), supporting the view in GBFM93.
Up to now only a few Class 0 objects have been studied in detail kinematically. Typical
examples are the nearby, isolated object IRAM 04191+1522 (IRAM 04191 for short; see
Belloche et al. 2002), or NGC 1333 IRAS 4A (IRAS 4A for short; see Di Francesco et al.
2001 and Belloche et al. 2006), which is actually a binary protostar at a separation of ∼
600 AU (Looney et al. 2000). Below, we analyze our 8 Class 0 objects together with these
two sources. Of the 8 Class 0 targets, CB 230, IRAS 03828, and L723 VLA2 are resolved as
binary protostars by our mm observations, while IRAS 04166, RNO 43, CB 68, CB 224, and
CB 244 remain single or unresolved. We treat RNO 43 and CB 230 as objects which provide
reliable rotation velocity gradients and take the other 6 measurements as upper limits. The
ASURV2 software package was used for the statistical analysis of the results. This package
performs a "survival analysis" which takes into account upper limits and allows to compute
a statistical sample mean. It must, however, be noted that two real measurements and 6
upper limits are not sufficient to derive statistically significant correlations.
The velocity gradients derived in our sample range from 5.8 to ≤ 24.2 km s−1 pc−1
(see Table 5). The mean value derived by the cumulative Kaplan-Meier (KM) estimator
in ASURV is 7.0±0.8 km s−1 pc−1. This result is consistent with the gradients derived in
IRAM 04191 (∼ 7 km s−1 pc−1) and IRAS 4A (∼ 9.3 km s−1 pc−1), but it is much larger than
the velocity gradients of dense cores derived from single-dish observations in GBFM93 and
2ASURV Rev.1.2 (LaValley, Isobe & Feigelson 1992) is a software package which implements the methods
presented in Feigelson & Nelson (1985). For details see http://astrostatistics.psu.edu/.
-- 16 --
CBMT02 (1−2 km s−1 pc−1). The correlation between velocity gradient (g) and core size
(R) is shown in Figure 7. It shows a clear trend with smaller cores (R < 5000 AU) having
larger velocity gradients. Taking into account only RNO 43 and CB 230, the entire dataset
could be fitted by a relation of g ∝ R−0.6±0.1, which is steeper than the slope of ∝ R−0.4
obtained by GBFM93 for the larger scale cores only. As expected, smaller (more evolved)
protostellar cores rotate much faster than larger (prestellar) cores.
4.4. Constraints on Angular Momentum
Assuming the velocity gradients are due to core rotation, the specific angular momentum
J/M of the objects can be calculated with the following equation (see GBFM93):
J/M = αrotωR2 =
2
g
5 + 2α
sini
R2 ≈ 1
4
gR2,
(5)
where the coefficient αrot = 2/(5 + 2p), p is the power-law index of the radial density profile
(here we adopt p = 1.5), g is the velocity gradient, and i is the inclination angle to the line
of sight direction. Here we assume sin i to be 1 for all sources. The derived specific angular
momenta J/M for our 8 Class 0 protostars are listed in column (6) in Table 5. We derived
values between < 0.12 and 0.45 × 10−3 km s−1 pc, with a KM sample mean of 0.21±0.1 ×
10−3 km s−1 pc.
In Figure 8 we show the distribution of specific angular momentum vs. size scale for
Class 0 single (unresolved) and binary protostars (this work), together with molecular cloud
cores, PMS binary stars, and single stars, etc. The data of NH3 dense cores and N2H+
starless cores are from GBFM93 and CBMT02, respectively. For PMS binary stars, the
GMBD×q/(1 + q)2 (where MB is
specific orbital angular momenta are derived as J/M =
the total stellar mass, D is the separation, and q is the mass ratio). Data of Class I and T
Tauri binaries are from Chen et al. (Paper IV, in prep.)3; data of very low-mass (< 0.1 M(cid:12))
and brown dwarf binaries are from Burgasser et al. (2007). The specific angular momentum
of Class I single stars is derived from J/M = vR, where v of 38 km s−1 and R of 2.7 R(cid:12)
are mean values from the sample of Class I stars in Covey et al. (2005) and adopted as
representative values here.
√
As shown in Fig. 8, there is a strong correlation between J/M and size scale in dense
cores (≥ 5000 AU). The data can be fitted with a power-law correlation of J/M ∼ R1.7±0.1,
3The masses of Class I and T Tauri binaries are dynamic masses. The angular momentum plotted in
Fig. 8 for these binaries does not include the angular momentum from the stellar rotation.
-- 17 --
which is consistent with the index 1.6 ± 0.2 obtained by GBFM93. This means that in
these dense cores the angular velocity is locked at a constant value to the first order, which
is generally explained by the mechanism of magnetic braking (see Basu & Mouschovias
1994). In addition, Fig. 8 shows that the mean J/M value in this region is considerably (∼
2 magnitudes) larger than the typical orbital angular momentum of T Tauri binary systems.
However, stars are not formed from the entire cloud (large dense cores), but only from the
dense inner R ≤ 5000 AU part of the cores that undergoes dynamical gravitational collapse
and decouples from the rest of the cloud. Fig. 8 shows that such cores have almost the same
specific angular momentum as the widest (few hundred AU) PMS binaries. But, there is
a gap in size scale between the smallest prestellar cores and the widest binaries. We find
that the specific angular momenta of Class 0 protostellar cores are located in this gap, but
nearly indistinguishable from those of wide PMS binary systems, i.e., angular momentum
is basically maintained (conserved) from the smallest prestellar cores via protostellar cores
to wide PMS binary systems. In the context that most protostellar cores are assumed to
fragment and form binary stars, this means that most of the angular momentum contained
in the collapse region is transformed into orbital angular momentum of the resulting stellar
binary system.
4.5. Energy Balance
In this section we estimate the contribution of different terms to the total energy balance
of the protostellar cores in our sample. In Table 6 we summarize the basic equations and
the estimated ratios for the rotational, thermal, and turbulent energy to the gravitational
potential energy. Here we assume a mean kinetic gas temperature ∼ 10 K for all objects. The
masses and radii used in the equations are virial masses and FWHM radii listed in Table 4.
The ratios of thermal and turbulent energy to gravitational potential energy ((cid:104)βtherm(cid:105)
≈ 0.26 and (cid:104)βturb(cid:105) ≈ 0.15) show that in these protostellar cores both thermal and turbulent
contribution together appear to dominate the support of the cores, but the thermal contri-
bution ∼ 2 times outweighs turbulence. The estimated βrot values range from < 0.004 to
0.017, with a KM sample mean of ∼ 0.007±0.002, which is much lower than βtherm and βturb.
This suggests that rotation is not dominating in the support of the protostellar cores. When
we apply the equilibrium virial theorem 2[Etherm + Eturb + Erot] + Egrav = 0 (in the absence
of magnetic fields), we find that all the Class 0 protostellar cores in our sample are slightly
supercritical (see Table 6).
Although the rotation energy is relatively small, it is thought to play an important role
in the fragmentation process (see reviews by Bodenheimer et al. 2000 and Tholine 2002). In
-- 18 --
general, if βrot is very large (βrot ≥ 0.25 − 0.3), a gas cloud can be stable against dynamical
collapse, potentially inhibiting fragmentation and star formation. However, if βrot is very
small, a cloud will not have enough rotational energy to experience fragmentation. Boss
(1999) has shown that rotating, magnetized cloud cores fragment when βrot > 0.01 initially4.
It should be noted that three sources in our sample might have a βrot > 0.01. Of these, CB 230
(βrot ∼ 0.017) and L723 VLA2 (βrot < 0.014) have been resolved as binary protostars. On
the other hand, IRAS 03282, a resolved binary protostar, has a very low βrot (∼ 0.004),
and IRAS 04166, a single (unresolved) protostar, has βrot value close to 0.01. Thus, our
observations can neither confirm nor disprove a relation between βrot and fragmentation. This
could be due to low-number statistics combined with observational uncertainties. Figure 9
shows the distribution of βrot with core size R for our sample, together with the cores from
GBFM93 and CBMT02. It shows that βrot is roughly independent of R, as suggested before
by GBFM93.
5. Summary and Conclusions
We present N2H+ (1-0) observations of 9 isolated low-mass protostellar cores using the
OVRO mm array. The main conclusions of this work are summarized as follows:
(1) N2H+ emission is detected in all target objects and the emission is spatially consistent
with the thermal dust continuum emission. The mean excitation temperature of the N2H+
line is ∼ 4.9 K. The mean FWHM core radius is (cid:104)R(cid:105) = 2000± 800 AU. The derived virial
masses of the N2H+ cores in our sample range from 0.3 to 1.2 M(cid:12), with a mean value of
0.6 M(cid:12). The corresponding mean hydrogen number densities range from 106 to 107 cm−3.
The N2H+ column densities in our sample range from 0.6 to 2.8 × 1012 cm−2, with a mean
value of 1.4 × 1012 cm−2. The average fractional abundances of N2H+, calculated by relating
the N2H+ column densities derived from the line strength to the virial masses, was found to
be ∼ 3.3 × 10−10. This is consistent with the result obtained in other surveys with single-dish
observations.
(2) The observed mean line widths range from 0.29 to 0.51 km s−1, with a mean value
of 0.37 km s−1. The non-thermal contribution is about 2.5 times larger than the thermal
line width, suggesting that the protostellar cores in our sample are not purely thermally
supported. We find that line widths are roughly constant within the interiors of the cores
and larger line widths only occur at the edges of the cores. We conclude that turbulence is
4Machida et al. (2005) find that the fragmentation does occur in rotating, magnetized clouds when βrot
≥ 0.04, considering magnetic fields suppress fragmentation.
-- 19 --
not negligible but subsonic in the protostellar cores.
(3) We derive the N2H+ velocity fields of eight Class 0 protostellar cores. CB 230 and
RNO 43 show symmetrical velocity gradients that can be explained by bulk rotation. In L723
VLA2, IRAS 03282, and IRAS 04166, outflow-envelope interaction appears to dominate the
velocity fields. CB 68, CB 224, and CB 244 show complicated velocity fields, which could
be affected by infall or large-scale turbulence. We argue that in these cores the observed
velocity gradients provide an upper limit to any underlying bulk rotation.
(4) The velocity gradients over the cores range from 6 to 24 km s−1, with a mean
value of ∼ 7 km s−1 pc−1. This is much larger than what has been found in single-dish
observations of prestellar cores, but agrees with recent interferometric observations of other
Class 0 protostellar cores. Assuming these gradients are due to rotation, the comparison
between gradients and core sizes suggests that smaller (evolved) protostellar cores rotate
much faster than larger dense (prestellar) cores. The data could be fitted by a relation of
g ∝ R−0.6±0.1.
(5) We find that in terms of specific angular momentum and size scale Class 0 protostellar
cores fill the gap between dense molecular cloud cores and PMS binary systems. There
appears to be no evolution (decrease) of angular momentum from the smallest prestellar
cores via protostellar cores to wide PMS binary system. In the context that most protostellar
cores are assumed to fragment and form binary stars, this means that most of the angular
momentum contained in the collapse region is transformed into orbital angular momentum
of the resulting stellar binary system.
(6) Both thermal and turbulent energy together dominate the support against gravity,
but the thermal contribution is about 2 times larger than turbulence. All protostellar cores
in our sample are found to be slightly virially supercritical.
(7) The ratio βrot of rotational energy to gravitational energy is relatively small in our
sample, ranging from 0.004 to 0.02, with a mean value of 0.007. We find that βrot values in
our sample show no clear correlation with observed binary protostars. On the other hand,
the three identified binary protostars are also not distinguished by βturb values. This could
be due to low-number statistics combined with observational uncertainties.
We thank the anonymous referee for many insightful comments and suggestions. The
Owens Valley millimeter-wave array was supported by NSF grant AST 9981546. Funding
from NASA's Origins of Solar Systems program (through grant NAG5-9530) is gratefully
acknowledged. Research at Owens Valley on the formation of young stars and planets was
also supported by the Norris Planetary Origins Project. We want to thank A. I. Sargent, who
-- 20 --
was directly involved in the early stages of this project, for fruitful discussions. We thank the
OVRO staff for technical support during the observations. We also thank A. Goodman and
S. Schnee for providing the VFIT routine and thank E. Feigelson for providing the ASURV
code.
Aikawa, Y., Ohashi, N., Inutsuka, S. I. et al. 2001, ApJ, 552, 639
REFERENCES
Andr´e, P., Ward-Thompson, D., & Barsony, M. 2000, in Protostars and Planets IV, ed. V.
Mannings, A. P. Boss, & S. S. Russell (Tucson: Univ. Arizona Press), 59
Anglada, G., Estalella, R., Rodriguez, L. F., Torrelles, J. M., Lopez, R., & Canto, J. 1991,
ApJ, 376, 615
Anglada, G., Rodriguez, L. F., & Torrelles, J. M. 1996, ApJ, 473, L123
Arce, H. G., & Sargent, A. I. 2004, ApJ, 624, 232
Bachiller, R. 1996, ARA&A, 34, 111
Bachiller, R., Terebey, S., Jarrett, T., Martin-Pintado, J., Beichman, C. A., & van Buren,
D. 1994, ApJ, 437, 296
Barranco, J. A, & Goodman, A. 1998, ApJ, 504, 207
Basu, S., & Mouschovias, T. C. 1994, ApJ, 432, 720
Bate, M. R. 2000, MNRAS, 314, 33
Bate, M. R., & Bonnell, I. A. 1997, MNRAS, 285, 33
Belloche, A., Andr´e, P., Despois, D., & Blinder, S. 2002, A&A, 393, 927
Belloche, A., Hennebelle, P., & Andr´e, P. 2006, A&A, 453, 145
Benson, P. J., Caselli, P., & Myers, P. C., 1998, ApJ, 506, 743
Benson, P. J., & Myers, P. C. 1989, ApJS, 71, 89
Bergin, E. A., Ciardi, D. R., Lada, C. J., Alves, J., & Lada, E. A. 2001, ApJ, 557, 209
Bergin, E. A., Langer, W. D. 1997, ApJ, 486, 316
-- 21 --
Bodenheimer, P., Burkert, A., Klein, R. I., & Boss, A. P. 2000, in Protostars and Planets
IV, ed. V. Mannings, A. P. Boss, & S. R. Russell (Tucson: Univ. Arizona Press), 675
Boss, A. P. 1999, ApJ, 520, 744
Briggs, D. S., Schwab, F. R., & Sramek, R. A. 1999, ASPC, 180, 127
Burgasser, A. J., Reid, I. N., Siegler, N. et al. 2007, in Protostars and Planets V, ed. B.
Reipurth, D. Jewitt, & K. Keil (Tucson: Univ. Arizona Press), 427
Cabrit, S., & Andr´e, P. 1991, ApJ, 379, 25
Caselli, P., Benson, P. J., Myers, P., & Tafalla, M. 2002, ApJ, 572, 238 (CBMT02)
Caselli, P., Walmsley, C. M., Tafalla, M., Dore, L., & Myers, P. C. 1999, ApJ, 523, L165
Caselli, P., Myers, P. C., Thaddeus, P. 1995, ApJ, 455, L77
Clemens, D. P., & Barvainis, R. 1988, ApJS, 68, 257
Covey, K. R., Greene, T. P., Doppmann, G. W., & Lada, C. J. 2005, AJ, 129, 2765
Davidson, J. A. 1987, ApJ, 315, 602
De Vries, Ch. H., Narayanan, G., & Snell, R. L. 2002, ApJ, 577, 798
Di Francesco, J., Myers, P. C., & Wilner, D. J. et al. 2001, ApJ, 562, 770
Duchene, G., Delgado-Donate, E., Haisch, K. E., Loinard, L., & Rodr´ıguez, L. 2007, in
Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tucson: Univ.
Arizona Press), 379
Feigelson, E. D., & Nelson, P. I., 1985, ApJ, 293, 192
Forveille, T. Guilloteau, S., & Lucas, R. 1989, CLASS Manual (Grenoble: IRAM)
Fuller, G. A., & Myers, P. C. 1992, ApJ, 384, 523
Girart, J. M., Estalella, R., Anglada, G., Torrelles, J. M., Ho, P. T. P., & Rodriguez, L. F.
1997, IAU Symp. 182: Herbig-Haro Flows and the Birth of Stars, 182, 266
Goldsmith, P. F., Snell, R. L., Hemeon-Heyer, M., & Langer, W. D. 1984, ApJ, 286, 599
Goodman, A. A., Barranco, J. A., Wilner, D. J., & Heyer, M. H. 1998, ApJ, 504, 223
Goodman, A. A., Benson, P. J., Fuller, G. A., & Myers, P. C. 1993, ApJ, 406, 528 (GBFM93)
-- 22 --
Halbwachs, J. L., Mayor, M., Udry, S., & Arenou, F. 2003, A&A, 397, 159
Launhardt R. 2001, in The Formation of Binary Stars, IAU Symp. 200, ed. H. Zinnecker, &
R. D. Mathieu (San Francisco: ASP), 117
Launhardt, R. 2004, in IAU Symp. 221, Star Formation at High Angular Resolution, ed.
M. G. Burton, R. Jayawardhana, & T. L. Bourke (San Francisco: ASP), 213
Launhardt, R., Evans, N. J., Wang, Y., Clemens, D. P., Henning, T., & Yun, J. L. 1998,
ApJS, 119, 59
Launhardt, R., & Henning, T. 1997, A&A, 326, 329
LaValley, M., Isobe, T., & Feigelson, E. D., 1992, "ASURV", Bull. Am. Astron. Soc.
Lee, C., Mundy, L. G., Stone, J. M., & Ostriker, E. C. 2002, ApJ, 576, 294
Looney, L. W., Mundy, L. G., & Welch, W. J. 2000, ApJ, 529, 477
Machida, M. N., Matsumoto, T., Hanawa, T., & Tomisaka, K. 2005, MNRAS, 362, 382
Mardones, D., Myers, P. C., Tafalla, M., Wilner, D. J., Bachiller, R., & Garay, G. 1997,
ApJ, 489, 719
Mathieu, R. D., Ghez, A. M., Jensen, E. L. N., & Simon, M. 2000, in Protostars and Planets
IV, ed. V. Mannings, A. P. Boss, & S. R. Russell (Tucson: Univ. Arizona Press), 703
Motte, F., & Andr´e, P. 2001, A&A, 365, 440
Padin, S., Scott, S. L., Woody, D. P., et al. 1991, PASP, 103, 461
Palacios, J., & Eiroa, C. 1999, A&A, 346, 233
Reipurth, B., Bally, J., & Devine, D. 1997, AJ, 114, 2708
Reipurth, B., Chini, R., Krugel, E., Kreysa, E., & Sievers, A. 1993, A&A, 273, 221
Reipurth, B., Jewitt, D., & Keil, K. (ed.) 2007, Protostars and Planets V (Tucson: Univ.
Arizona Press)
Sault, R. J., Teuben, P. J., & Wright, M. C. H. 1995, in ASP Conf. Ser. 77, Astronomical
Data Analysis Software and Systems IV, ed. R. A. Shaw, H. E. Payne, & J. J. E.
Hayes (San Francisco: ASP), 443
Scoville, N. Z., Carlstrom, J. E., Chandler, C. J. et al. 1993, PASP, 105, 1482
-- 23 --
Shirley, Y. L., Evans II, N. J., Rawlings, J. M. C., & Gregersen, E. M. 2000, ApJS, 131, 249
Tafalla, M., Santiago, J., Johnstone, D., & Bachiller, R. 2004, A&A, 423, L21
Tohline, J. E. 2002, ARA&A, 40, 349
Turner, B. E., & Thaddeus, P. 1977, ApJ, 211, 755
Vall´ee, J. P., Bastien, P., & Greaves, J. S. 2000, ApJ, 542, 352
Volgenau, N. H., Mundy, L. G., Looney, L. W., & Welch, W. J. 2006, ApJ, 651, 301
Wang, Y., Evans II, N. J., Zhou, S., & Clemens, D. P. 1995, ApJ, 454, 217
Williams, J. P., de Geus, E. J., & Blitz, L. 1994, ApJ, 428, 693
Wolf, S., Launhardt, R., & Henning, T. 2003, ApJ, 592, 233
Womack, M., Ziurys, L. M., & Wyckoff, S. 1992, ApJ, 387, 417
Wu, Y., Huang, M., & He, J. 1996, A&AS, 115, 283
Yun, J. L. & Clemens, D. P. 1994a, AJ, 108, 612
Yun, J. L. & Clemens, D. P. 1994b, ApJS, 92, 145
Zinnecker, H., & Mathieu, R. D. (ed.) 2001, IAU Symposium. 200, The Formation of Binary
stars (San Francisco: ASP)
Zinnecker, H., Bastien, P., Arcoragi, J. P., & Yorke, H. W. 1992, A&A, 265, 726
This preprint was prepared with the AAS LATEX macros v5.2.
-- 24 --
Table 1. Basic properties of target sources
Source
Name
Associated
IRAS source
D
[pc]
Lbol
[L(cid:12)]
Tbol Menv
[M(cid:12)]
[K]
Outflow
Infall
Class
Refs.
03282+3035
IRAS 03282
IRAS 04166
04166+2706
RNO 43 MM 05295+1247
16544−1604
CB 68
19156+1906
L723 VLA2
19179+1129
CB 188
20355+6343
CB 224
CB 230
21169+6804
23238+7401
CB 244
300
140
400
160
300
300
400
400
180
1.2
0.4
6.0
1.6
3.3
2.6
16
8.2
1.1
23
91
74
47
56
2.9
1.0
0.4
3.5
7.3
0.7
6.6
5.1
3.3
y, bipol, coll.
y, bipol, coll.
y, bipol, coll.
y, bipol, coll.
y, quadrupol.
y, bipol
y, bipol
y, bipol, coll.
y, bipol
y
y
y
y
y
0
0
0
0
0
I
1,2,3
1,3,4
5,6
7,8
3,9
8,10
0/I
0/I
0
11,12,13
10,11,12
8,10,14
aThe IRAS source is not always associated with the mm source.
Note. -- References. -- (1) Mardones et al. 1997; (2) Bachiller et al. 1994; (3) Shirley et al. 2000; (4)
Tafalla et al. 2004; (5) Zinnecker et al. 1992; (6) Arce & Sargent 2004; (7) Vall´ee et al. 2000; (8) Launhardt
& Henning 1997; (9) Lee et al. 2002; (10) Yun & Clemens 1994; (11) Wolf et al. 2003; (12) Launhardt et al.
1998; (13) Chen et al. in prep.; (14) Wang et al. 1995
Table 2. Target list and summary of observations
Source
Name
Other
Name
R.A. & Dec. (1950)a
[h : m : s, ◦ : (cid:48) : (cid:48)(cid:48)]
Array
configuration
...
B 213
...
...
IRAS 03282
IRAS 04166
RNO 43 MM L 1582B
CB 68
L723 VLA2
CB 188
CB 224
CB 230
CB 244
L 1100
L 1177
L 1262
L 146
03:28:15.2, +30:35:14
04:16:37.8, +27:06:29
05:29:30.6, +12:47:35
16:54:27.2, −16:04:44
19:15:41.8, +19:06:45
19:17:54.1, +11:30:02
20:35:30.6, +63:42:47
21:16:53.7, +68:04:55
23:23:48.5, +74:01:08
CLE
CLU
CLEH
CLH
CLE
CLEU
CL
CL
CL
[kλ]
UV coverage HPBWb
[arcsecs]
5.5×4.3
4.8×4.2
5.3×4.7
8.4×5.0
5.5×4.5
4.4×4.1
5.4×5.1
7.1×6.4
6.3×5.2
3 -- 36
3 -- 145
3 -- 68
3 -- 62
3 -- 36
3 -- 127
3 -- 36
3 -- 36
3 -- 36
rms
[mJy/beam]
76
45
52
71
81
55
97
63
86
aReference position for figures and tables in the paper.
bSynthesized FWHM beam size with robust weighting
-- 25 --
Table 3. Parameters from N2H+ (1-0) spectral fittinga
Source
Name
IRAS 03282
IRAS 04166
RNO 43
CB 68
L723 VLA2
CB 188
CB 224
CB 230
CB 244
VLSR
(km s−1)
6.89±0.01
6.64±0.02
9.77±0.01
5.26±0.04
11.00±0.02
7.23±0.02
−2.65±0.01
2.78±0.01
3.24±0.01
Tex
(K)
4.98±0.05
4.08±0.13
5.15±0.25
4.25±0.04
5.74±0.21
4.58±0.08
4.61±0.04
5.49±0.13
4.93±0.11
τtot
1.38±0.08
1.76±0.18
0.77±0.01
1.38±0.09
0.78±0.03
0.73±0.04
1.75±0.08
1.54±0.04
1.09±0.16
(cid:52)v
(km s−1)
0.51±0.03
0.34±0.01
0.41±0.02
0.37±0.04
1.07±0.04
0.38±0.03
0.50±0.03
0.36±0.01
0.38±0.02
b
(cid:52)vmean
(km s−1)
0.29±0.01
0.32±0.01
0.39±0.01
0.33±0.01
0.51±0.01
0.31±0.01
0.47±0.01
0.29±0.01
0.45±0.02
c
(cid:52)vN T
(km s−1)
0.26±0.02
0.29±0.02
0.37±0.02
0.30±0.02
0.49±0.02
0.28±0.02
0.45±0.02
0.26±0.02
0.43±0.03
aValue at the intensity peak. The error represents 1 σ error in the hyperfine fitting.
bMean line width for each object (see §3.4).
cNon-thermal line width for each object at the given gas temperature (10 K; see §3.4).
Table 4.
Size, density, and mass
Source
R
[AU]
Mvir
[M(cid:12)]
(cid:104)nH2(cid:105)a
[×105 cm−3]
N (N2H +)
[×1012cm−2]
MN2H+
[×10−10 M(cid:12)]
X(N2H +)
[×10−10]
1703
IRAS 03282
IRAS 04166
1232
RNO 43 MM 3530
1440
CB 68
1272
L723 VLA2
1763
CB 188
1329
CB 224
CB 230
2762
1774
CB 244
0.42
0.30
1.21
0.37
0.65
0.44
0.57
0.71
0.72
22.6
43.6
7.4
33.7
84.4
21.6
64.7
9.0
34.9
1.81
0.96
0.82
0.93
2.84
0.57
1.95
1.52
0.96
aMean density, computed from R and Mvir (see §3.2)
1.57
0.31
2.32
0.51
1.89
0.46
2.07
3.86
0.69
5.11
1.41
2.60
1.83
3.96
1.42
4.97
7.41
1.30
Table 5. Velocity gradient and specific angular momentum
Source
IRAS 03282
IRAS 04166
RNO 43 MM
CB 68
L723 VLA2
CB 224
CB 230
CB 244
mean velocity
[km s−1]
g
[km s−1 pc−1]
6.91
6.64
9.72
5.15
11.03
-2.67
2.69
3.51
<6.6
<12.5
5.8±0.1
<10.3±0.2
<24.2
<11.2
8.8±0.1
<22.9
Θg
[degree]
70.8±1.2
-134.2±1.7
-23.1±1.0
161.2±0.8
-139.2±0.2
-82.7±1.0
-54.0±0.4
51.2±0.2
gr
[km s−1]
J/M
[×10−3 km s−1 pc]
0.11
0.15
0.21
0.15
0.30
0.15
0.24
0.41
<0.12
<0.12
0.45
<0.13
<0.24
<0.12
0.42
<0.45
-- 26 --
Table 6. Energy balance
Source
grav (×1035J)
Ea
βb
rot
βc
therm
βd
turb
IRAS 03282
IRAS 04166
RNO 43 MM
CB 68
L723 VLA2
CB 224
CB 230
CB 244
1.34
0.99
5.51
1.29
4.36
3.19
2.40
3.91
<0.004
<0.007
0.009
<0.006
<0.014
<0.004
0.017
<0.03
0.33
0.33
0.24
0.31
0.16
0.19
0.31
0.20
0.11
0.14
0.16
0.14
0.19
0.19
0.11
0.18
vir
βe
−0.11
−0.05
−0.19
−0.08
−0.27
−0.23
−0.12
−0.18
aGravitational potential energy: Egrav = αvirGM 2/R.
bβrot = Erot
Egrav
= 1
2
for all the objects.
cβtherm = Etherm
Egrav
αrot
αvir
GM = 0.17 g2
ω2R3
sin2i
R3
GM . Here we assume sini = 1
and Etherm is the thermal energy estimated from
(k is the Boltzman constant and µ = 2.33 is the mean
Etherm = 3
2 M kT
µmH
molecular weight).
dβturb = Eturb
Egrav
and Eturb is the turbulent energy estimated from the non-
(cid:52)vturb
thermal line width Eturb = 3/2M σ2
8ln2 . Here we assume that
the non-thermal line widths at the given temperature in our sample are from
turbulence.
N T (σ2
N T =
eβvir = 2(βrot + βtherm + βturb) − 1.
-- 27 --
Fig. 1. -- Maps of the N2H+ (1−0) intensity integrated over the seven hyperfine components
for 9 protostars. The unit of the scale is [Jy beam−1 km s−1]. The contours start at ∼
3 σ with steps of ∼ 2 σ. Beam sizes are shown as grey ovals in each map. The crosses
in the maps represent the peaks of 3 mm dust continuum emission (except IRAS 03282, in
which the crosses indicate the peaks of 1.3 mm dust emission). All the positions of dust
emission are selected from Launhardt et al. 2007 (in prep.). The asterisk in CB 230 marks
the positions of the secondary protostar observed at 7 µm with ISOCAM (not detected at
3 mm, see Launhardt 2001).
-- 28 --
Fig. 2. -- N2H+ spectra at the peak position of the nine observed protostars. Gray dotted
curves show the hyperfine structure line fitting. Fitting results are given in Table 3.
-- 29 --
Fig. 3. -- N2H+ velocity field maps of 8 Class 0 protostars. The unit of the scale is km s−1.
The contours in each map are from Fig. 1, but range from 30% to 99% of the peak intensity
by the step of 10%. The crosses in each are same as them in Fig. 1. The red and blue arrows
show the directions of the red- and blue-shifted outflow for each source (see text).
-- 30 --
Fig. 4. -- Diagrams of the correlation between line width and solid angle area for the nine
protostars in our sample. Black solid curves and values in each map show the results of
Gaussian fitting.
0.250.300.350.400.450.500.550.600.65051015200.29(0.01)kms-1IRAS03282W(arcsec2)0.260.280.300.320.340.360.38024681012140.32(0.01)kms-1IRAS041660.250.300.350.400.450.500.550.600.65051015202530RNO430.389(0.002)kms-10.300.350.400.450.500.550.600.650.700.75010203040506070CB680.33(0.01)kms-1W(arcsec2)0.30.40.50.60.70.80.91.01.11.21.31.41.5024681012141618202224L723VLA20.51(0.01)kms-10.260.280.300.320.340.36051015202530354045500.31(0.01)kms-1CB1880.250.300.350.400.450.500.550.600.65024681012CB2240.467(0.003)kms-1W(arcsec2)Linewidth(kms-1)0.270.300.330.360.390.420.45010203040506070CB2300.290(0.001)kms-1Linewidth(kms-1)0.30.40.50.60.70.80.91.01.11.21.31.41.51.60510152025301.06(0.01)kms-10.45(0.02)kms-1CB244Linewidth(kms-1) -- 31 --
Fig. 5. -- Line widths distribution of 6 Class 0 protostars in our sample. The unit of the
scale is km s−1. The crosses in each map are same as them in Fig. 1 and the contours in each
map are same as them in Fig. 3. Solid and dashed lines in each map show the direction of
blue- and red-shifted outflow, respectively.
-- 32 --
Fig. 6. -- Non-thermal line width (cid:52)VNT vs. size R of dense molecular cloud cores. Data of
NH3 (open squares) and N2H+ (open triangles) dense cores are adopted from GBFM93 and
CBMT02, respectively. Solid lines marked 0.13, 0.17 and 0.44 km s−1 represent thermal line
widths of N2H+, NH3, and 2.33 mH mass "average" particle at 10 K, respectively. The fit to
the GBFM93 and CBMT02 data results in power-law indexes of 0.21±0.05 and 0.51±0.15,
respectively. The levels (p-values) of statistical significance are < 0.01% (GBFM93) and ∼
0.1% (CBMT02), respectively.
1000100001000000.110.44 km s-10.17 km s-10.13 km s-1V ~ R0.51 (0.15)V ~ R0.21 (0.05)20000 AU5000 AU VNT (km s-1)R (AU) NH3 dense cores (GBFM93) N2H+ dense cores (CBMT02) Class 0 protostars (this work) -- 33 --
Fig. 7. -- Mean velocity gradient vs.
size R of dense molecular cloud cores. Data of
NH3 (open squares) and N2H+ (open triangles) dense cores are adopted from GBFM93 and
CBMT02, respectively. Data of IRAM 04191 and NGC1333 IRAS4A (asterisks) are adopted
from Belloche et al. (2002; 2006). Solid line shows the fitting result with a power-law index
of −0.6±0.1. The level of statistical significance is ∼ 0.1%.
100010000100000110R-0.6IRAM 04191IRAS 4A NH3 dense cores (GBFM93) N2H+ dense cores (CBMT02) Class 0 protostars (this work) Gradient (km s-1 pc-1)R (AU) -- 34 --
Fig. 8. -- Distribution of specific angular momentum J/M vs. size R of molecular cloud
cores, protostars, and stars. For the dense cores, the specific angular momenta are derived
from J/M = 2/5 (R2 × g) (see text). The data of the NH3 dense cores are from Goodman et
√
al. (1993); the data of the N2H+ starless cores are from Caselli et al. (2002). For the binary
GM D×q/(1 + q)2 (see
stars, the orbital specific angular momenta are derived from J/M =
text). The data of Class I binaries and T Tauri binaries are from Chen et al. (2007 IV, in
prep.), and the masses are all dynamic masses; The data of very low-mass (< 0.1 M(cid:12)) and
brown dwarf binaries are from Burgasser et al. (2007).
-3-2-10123456121314151617181920IRAS4AClass0protostars100AU5000AUPMSbinariesJ/M~R0.6DensecoresJ/M~R1.7IRAM04191LatetypeMSstarsClassIsinglestarsNH3densecoresN2H+starlesscoresClassIbinariesTTSbinariesVLM&BDbinariesClass0protostars(thiswork)LargeMolecularcloudcoresLog(J/M)(m2s-1)Log(size)(AU) -- 35 --
Fig. 9. -- Ratio of rotational to gravitational energy βrot vs. size R. Data of NH3 (open
squares) and N2H+ (open triangles) dense cores are adopted from GBFM93 and CBMT02,
respectively. Data of IRAM 04191 and NGC1333 IRAS4A (asterisks) are adopted from
Belloche et al. (2002; 2006).
1000100001000001E-30.010.11IRAS 4AIRAM 04191L723 VLA2CB 230 NH3 dense cores (GBFM93) N2H+ dense cores (CBMT02) Class 0 protostars (this work) rotR (AU) |
astro-ph/9904365 | 1 | 9904 | 1999-04-26T20:25:55 | The fluid mechanics of dark matter formation: Why does Jeans's (1902 & 1929) theory fail? | [
"astro-ph"
] | Jeans's (1902 & 1929) linear gravitational instability criterion gives truly spectacular errors in its predictions of cosmological structure formation according to Gibson's (1996) new nonlinear theory. Scales are determined by viscous or turbulent forces, or by diffusivity, at Schwarz length scales L_SV, L_ST, or L_SD, respectively, whichever is larger. By these new criteria, void formation begins in the plasma epoch soon after matter dominates energy, at L approx L_SV = (gamma nu / rho G)^1/2 scales corresponding to protosuperclusters, decreasing to protogalaxies at the plasma-gas transition, where gamma is the rate-of-strain of the expanding universe, nu is the kinematic viscosity, rho is the density, and G is Newton's gravitational constant. Condensation of the primordial gas occurs at mass scales a trillion times less than the Jeans mass to form a `fog' of micro-brown-dwarf (MBD) particles that persist as the galactic baryonic dark matter, as reported by Schild (1996) from quasar-microlensing studies. Nonbaryonic dark matter condensation is prevented by its enormous diffusivity at scales smaller than L_SD = (D^2 /rho G)^1/2, where D is the diffusivity, so it forms outer halos of galaxies, cluster-halos of galaxy clusters, and supercluster-halos. | astro-ph | astro-ph |
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
12 (12.03.4; 12.04.1; 12.05.1; 12.07.1; 12.12.1)
The fluid mechanics of dark matter formation
Why does Jeans's (1902 & 1929) theory fail?
Carl H. Gibson
Departments of Applied Mechanics and Engineering Sciences and
Scripps Institution of Oceanography, University of California at San Diego,
La Jolla, CA 92093-0411, USA
email: [email protected]
Received July 1, 1998;
Abstract. Jeans's (1902 & 1929) gravitational instability criterion gives truly spectac-
ular errors in its predictions of cosmological structure formation according to Gibson's
(1996) new theory. It is suggested that the linear perturbation stability analysis leading to
the acoustic Jeans length scale criterion L ≥ LJ ≡ Vs/(ρG)1/2 is quite irrelevant to grav-
itational condensation, which is intrinsically nonlinear and nonacoustic. Instead, conden-
sation is limited by viscous or turbulent forces, or by diffusivity, at Schwarz length scales
LSV , LST , or LSD, respectively, whichever is larger, independent of LJ and the sound
speed Vs. By these new criteria, cosmological structure formation begins in the plasma
epoch soon after matter dominates energy, at L ≈ LSV ≡ (γν/ρG)1/2 scales correspond-
ing to protosuperclusters, decreasing to protogalaxies at the plasma-gas transition, where
γ is the rate-of-strain of the expanding universe, ν is the kinematic viscosity, ρ is the den-
sity, and G is Newton's gravitational constant. Condensation of the primordial gas occurs
at mass scales a trillion times less than the Jeans mass to form a 'fog' of micro-brown-
dwarf (MBD) particles that persist as the galactic baryonic dark matter, as reported by
Schild (1996) from quasar-microlensing studies. Nonbaryonic dark matter condensation
is prevented by its enormous diffusivity at scales smaller than LSD ≡ (D2/ρG)1/2, where
D is the diffusivity, so it forms outer halos of galaxies, cluster-halos of galaxy clusters,
and supercluster-halos.
Key words: Cosmology:theory -- dark matter -- early Universe -- gravitational lensing --
large-scale structure formation of the Universe
Send offprint requests to: C. H. Gibson
2
Carl H. Gibson: The fluid mechanics of dark matter formation
1. Introduction
Jeans's theory fails because it is linear. Linear theories typically fail when applied to
nonlinear processes; for example, applications of the linearized Navier-Stokes equations
to problems of fluid mechanics give laminar solutions, but observations show that actual
flows are turbulent when the Reynolds number is large.
Cosmology (e.g.; Peebles 1993, Padmanabhan 1993) relies exclusively on Jeans's the-
ory in modelling the formation of structure by gravitational instability. Because LJ for
baryonic (ordinary) matter in the hot plasma epoch following the Big Bang is larger than
the Hubble scale of causality LH ≡ ct, where c is the speed of light and t is the time, no
baryonic structures can form until the cooling plasma forms neutral gas. Star and galaxy
formation models invented to accommodate both Jeans's theory and the observations of
early structure formation have employed a variety of innovative maneuvers and concepts.
Nonbaryonic "cold dark matter" was invented to permit nonbaryonic condensations in
the plasma epoch with gravitational potential wells that could guide the early formation
of baryonic galaxy masses. Fragmentation theories were proposed to produce M⊙-stars
rather than Jeans-superstars at the 105M⊙ proto-globular-cluster Jeans mass of the hot
primordial gas.
Both of these concepts have severe fluid mechanical difficulties according to the
Gibson 1996 theory. Cold dark matter cannot condense at the galactic scales needed
because its nonbaryonic, virtually collisionless, nature requires it to have an enor-
mous diffusivity with supergalactic LSD length scales. Fragmentation theories (e.g.,
Low and Lynden-Bell 1976) are based on a faulty condensation premise that implies
large velocities and a powerful turbulence regime that would produce a first gener-
ation of large stars with minimum mass determined by the turbulent Schwarz scale
LST ≡ ε1/2/(ρG)3/4, where ε is the viscous dissipation rate of the turbulence, with a
flurry of starbursts, supernovas, and metal production that is not observed in globu-
lar star clusters. The population of small, long-lived, metal-free, globular cluster stars
observed is strong evidence of a quiet, weakly turbulent formation regime.
2. Jeans's acoustic theory
Jeans considered the problem of gravitational condensation in a large body of nearly con-
stant density, nearly motionless gas. Viscosity and diffusivity were ignored. The density
and momentum conservation equations were linearized by dropping second order terms
after substituting mean plus fluctuating values for the density, pressure, gravitational po-
tential, and velocity. Details of the derivation are given in many cosmological texts (e.g.;
Carl H. Gibson: The fluid mechanics of dark matter formation
3
Kolb and Turner 1994, p342) so they need not be repeated here. The mean gravitational
force ∇φ is assumed to be zero, violating the Poisson equation
∇2φ = 4πGρ,
(1)
where φ is the gravitational potential, in what is known as the Jeans swindle. Cross-
differentiating the linearized perturbation equations produces a single, second order dif-
ferential equation satisfied by Fourier modes propagating at the speed of sound Vs. From
the dispersion equation
ω2 = V 2
s k2 − 4πGρ,
(2)
(4πGρ/V 2
where ω is the frequency and k is the wavenumber, a critical wavenumber kJ =
s )1/2 exists, called the Jeans wavenumber. For k less than kJ , ω is imaginary
and the mode grows exponentially with time. For k larger than kJ , the mode is a propa-
gating sound wave. Density was assumed to be a function only of pressure (the barotropic
assumption).
Either the barotropic assumption or the linearization of the momentum and density
equations are sufficient to reduce the problem to one of acoustics. Physically, sound
waves provide density nuclei at wavecrests that can trigger gravitational condensation if
their time of propagation λ/Vs for wavelength λ is longer than the gravitational free fall
time τg ≡ (ρG)−1/2. Setting the two times equal gives the Jeans gravitational instability
criterion: gravitational condensation occurs only for λ ≥ LJ .
Jeans's analysis fails to account for the effects of gravity, diffusivity, or fluid me-
chanical forces upon nonacoustic density maxima and density minima; that is, points
surrounded on all sides by either lower or higher density. These move approximately
with the fluid velocity, not Vs, (Gibson 1968). The evolution of such zero gradient
points and associated minimal gradient surfaces is critical to turbulent mixing the-
ory (Gibson et al. 1988). Turbulence scrambles passive scalar fields such as tempera-
ture, chemical species concentration and density to produce nonacoustic extrema, sad-
dle points, doublets, saddle lines and minimal gradient surfaces. A quasi-equilibrium
develops between convection and diffusion at such zero gradient points and minimal
gradient surfaces that is the basis of a universal similarity theory of turbulent mixing
(Gibson 1991) analogous to the universal similarity theory of Kolmogorov for turbulence.
Just as turbulent velocity fields are damped by viscosity at the Kolmogorov length scale
LK ≡ (ν/γ)1/2, where ν is the kinematic viscosity and γ is the rate-of-strain, scalar fields
like temperature are damped by diffusivity at the Batchelor length scale LB ≡ (D/γ)1/2,
where D is the molecular diffusivity. This prediction has been confirmed by laboratory ex-
periments and numerical simulations (Gibson et al. 1988) for the range 10−2 ≤ P r ≤ 105,
where the Prandtl number P r ≡ ν/D.
4
Carl H. Gibson: The fluid mechanics of dark matter formation
On cosmological length scales, density fields scrambled by turbulence are not nec-
essarily dynamically passive but may respond to gravitational forces. In the density
conservation equation
∂ρ/∂t + vi(∂ρ/∂xi) = Deff ∂ 2ρ/∂xj∂xj
(3)
the effective diffusivity of density Deff ≡ D − L2/τg is affected by gravitation in the
vicinity of minimal density gradient features, and reverses its sign to negative if the fea-
ture size L is larger than the diffusive Schwarz scale LSD (Gibson and Schild 1998a).
LSD ≡ (D2/ρG)1/4 is derived by setting the diffusive velocity vD ≈ D/L of an isodensity
surface a distance L from a minimal gradient configuration equal to the gravitational
velocity vg ≈ L/τg. Thus, nonacoustic density maxima in a quiescent, otherwise ho-
mogeneous, fluid are absolutely unstable to gravitational condensation, and nonacoustic
density minima are absolutely unstable to void formation, on scales larger than LSD.
Jeans believed from his analysis (Jeans 1929) that sound waves with λ ≥ LJ would
grow in amplitude indefinitely, producing unlimited kinetic energy from his gravitational
instability. This is clearly incorrect, since any wavecrest that collects a finite quantity of
mass from the ambient fluid will also collect its zero momentum and become a nonacoustic
density nucleus. From the enormous Jeans mass values indicated at high temperature,
he believed he had proved his speculation that the cores of galaxies consisted of hot
gas (emerging from other Universes!) and not stars, which could only form in the cooler
(smaller LJ ) spiral arms, thrown into cold outer space by centrifugal forces of the spinning
core. The concepts of pressure support and thermal support often used to justify Jeans's
theory are good examples of bad dimensional analysis, lacking any proper physical basis.
3. Fluid mechanical theory
Gravitational condensation on a nonacoustic density maximum is limited by either dif-
fusion or by viscous, magnetic or turbulent forces at diffusive, viscous, magnetic, or
turbulent Schwarz scales LSX, whichever is largest, where X is D, V, M, T , respectively
(Gibson 1996, Gibson and Schild 1998a). Magnetic forces are assumed to be unimpor-
tant for the cosmological conditions of interest. Gravitational forces Fg ≈ ρ2GL4 equal
viscous forces FV ≈ ρνγL2 at LSV ≡ (νγ/ρG)1/2, and turbulent forces FT ≈ ρ(ε)2/3L8/3
at LST ≡ ε1/2/(ρG)3/4. Kolmogorov's theory is used to estimate the turbulent forces as
a function of length scale L.
The criterion (Gibson 1996, 1997a, 1997b; Gibson and Schild 1998a, 1998b) for grav-
itational condensation or void formation at scale L is therefore
L ≥ (LSX )max ; X = D, V, M, T.
(4)
Carl H. Gibson: The fluid mechanics of dark matter formation
5
4. Structures in the plasma epoch
Without the Jeans constraint, structure formation begins in the early stages of the hot
plasma epoch after the Big Bang when decreasing viscous forces first permit gravitational
decelerations and sufficient time has elapsed for the information about density varia-
tions to propagate; that is, the decreasing viscous Schwarz scale LSV becomes smaller
than the increasing Hubble scale LH ≡ ct, where c is the velocity of light. Low levels
of temperature fluctuations of the primordial gas indicated by the COsmic microwave
Background Experiment (COBE) satellite (δT /T ≈ 10−5) constrain the velocity fluc-
tuations δv/c ≪ 10−5 to levels of very weak turbulence. Setting the observed mass of
superclusters ≈ 1046 kg equal to the Hubble mass ρL3
H computed from Einstein's equa-
tions (Weinberg 1972, Table 15.4) indicates the time of first structure was ≈ 1012 s, or
30 000 y (Gibson 1997b).
Setting LH ≈ 3 1020 m (10 kpc) = LSV gives ν ≈ 6 1027 m2 s−1 with ρ ≈ 10−15 kg m−3
and γ ≈ 1/t = 10−12 rad s−1. Such a large viscosity suggests a neutrino-electron-proton
coupling mechanism, presumably through the Mikheyev-Smirnov-Wolfenstein (MSW)
effect (Bahcall 1997), supporting the Neutrino-98 claim that neutrinos have mass.
The viscous condensation mass ρL3
SV decreases to about 1042 kg (Gibson 1996) as
the Universe expands and cools to the plasma-gas transition at t ≈ 1013 s, or 300 000 y,
based on Einstein's equations to determine T and ρ and assuming the usual dependence of
viscosity ν on temperature T (Weinberg 1972). Assuming gravitational decelerations that
are possible always occur, we see that protosupercluster, protocluster, and protogalaxy
structure formation should be well underway before the emergence of the primordial gas.
5. Primordial fog formation
The first condensation scales of the primordial gas mixture of hydrogen and helium are
the maximum size Schwarz scale, and an initial length scale LIC ≡ (RT /ρG)1/2 equal to
the Jeans scale LJ but independent of Jeans's linear perturbation stability analysis, and
acoustics, where R is the gas constant of the mixture. From the ideal gas law p/ρ = RT
we see that density increases can be compensated by pressure increases with no change
in temperature in a uniform temperature gas, and that gravitational forces Fg ≈ ρ2GL4
will dominate the resulting pressure gradient forces Fp ≈ pL2 = ρRT L2 for length scales
T ≈ 3 000 K gives a condensation mass ρL3
L ≥ LIC . Taking R ≈ 5 000 m2 s−2 K−1, ρ ≈ 10−18 kg m−3 m (Weinberg 1972) and
IC ≈ 105M⊙, the mass of a globular cluster of
stars. Because the temperature of the primordial gas was observed to be quite uniform
by COBE, we can expect the protogalaxy masses of primordial gas emerging from the
plasma epoch to immediately fragment into proto-globular-cluster (PGC) gas objects on
LIC ≈ 3 1017 m (10 pc) scales, with subfragments at (LSX )max.
6
Carl H. Gibson: The fluid mechanics of dark matter formation
The kinematic viscosity ν of the primordial gas mixture decreased by a factor of
about a trillion from plasma values at transition, to ν ≈ 2.4 1012 m2 s−1 assuming the
density within the PGC objects are about 10−17 kg m−3. Therefore, the viscous Schwarz
scale LSV ≈ (2.4 1012 10−13/10−17 6.7 10−11)1/2 = 1.9 1013 m, so the viscous Schwarz
SV ρ = 6.8 1022 kg, or MSV = 6.8 1024 kg using ρ = 10−18. The turbulent
Schwarz mass MST ≈ 8.8 1022 kg assuming 10% of the COBE temperature fluctuations
mass MSV ≈ L3
are due to turbulent red shifts ([(δv/c)/(δT /T )] = 10−1) as a best estimate.
We see that the entire universe of primordial H-He gas turned to fog soon after the
plasma-gas transition, with primordial fog particle (PFP) mass values in the range 1023
to 1025 kg depending on the estimated density and turbulence levels of the gas. The
time required to form a PFP is set by the time required for void regions to grow from
minimum density points and maximum density saddle points to surround and isolate
the condensing PFP objects (Gibson and Schild 1998a). Voids grow as rarefaction waves
with a limiting maximum velocity Vs set by the second law of thermodynamics, so the
minimum PFP formation time is τP F P ≤ (LSX )max/Vs, or about 103 y. Full condensation
of the PFP to form a dense core near hydrostatic equilibrium requires a much longer time,
near the gravitational free fall time τg ≈ 2 106 y.
Radiation heat transport during the PFP condensation period before the creation of
dust should have permitted cooling to temperatures near those of the expanding universe.
After about a billion years hydrogen dew point and freezing point temperatures (20-13 K)
would be reached, forming the micro-brown-dwarf conditions expected for these widely
separated (103 −104 AU) small planetary objects that comprise most of the baryonic dark
matter of the present universe and the materials of construction for the stars and heavy
elements. Because such frozen objects occupy an angle of less than a micro-arcsecond
viewed from their average separation distance, they are invisible to most observations
except by gravitational microlensing, or if a nearby hot star brings these volatile comets
out of cold storage.
6. Observations
A variety of observations confirm the new theory that fluid mechanical forces and diffu-
sion limit gravitational condensation (Gibson 1996), and confute Jeans's (1902 & 1929)
acoustic criterion:
-- quasar microlensing at micro-brown-dwarf frequencies (Schild 1996),
-- tomography of dense galaxy clusters indicating diffuse (nonbaryonic) superhalo dark
matter at LSD scales with Dnb ≈ 1028m2 s−1 (Tyson and Fischer 1995),
-- the Gunn-Peterson missing gas sequestered as PFPs,
-- the dissipation of 'gas clouds' in the Ly − α forest,
Carl H. Gibson: The fluid mechanics of dark matter formation
7
-- extreme scattering events, cometary globules, FLIERS, ansae, Herbig-Haro 'chunks',
etc.
Evidence that the dark matter of galaxies is dominated by small planetary mass
objects has been accumulating from reports of many observers that the multiple images
of lensed quasars twinkle at corresponding high frequencies. After several years spent
resolving a controversy about the time delay between images Q0957+561A,B, to permit
correction for any intrinsic fluctuations of the light intensity of the source by subtraction
of the properly phased images, Schild 1996 announced that the lensing galaxy mass
comprises ≈ 10−6M⊙ "rogue planets" that are "likely to be the missing mass."
Star-microlensing studies from the Large Magellanic Cloud have failed to detect lens-
ing at small planetary mass frequencies, thus excluding this quasar-microlensing popula-
tion as the Galaxy halo missing mass (Alcock et al. 1998, Renault et al. 1998). However,
the exclusion is based on the unlikely assumption that the number density of such small
objects is uniform. The population must have mostly primordial gas composition since
no cosmological model predicts this much baryonic mass of any other material, and must
be primordial since it constitutes the material of construction, and an important stage,
in the condensation of the gas to form stars. Gravitational aggregation is a nonlinear,
self-similar, cascade process likely to produce an extremely intermittent lognormal spatial
distribution of the PFP number density, with mode value orders of magnitude smaller
than the mean. Since star-microlensing from a small solid angle produces a small num-
ber of independent samples, the observations estimate the mode rather than the mean,
resolving the observational conflict (Gibson and Schild 1998b).
7. Conclusions
1. Jeans's gravitational instability criterion L ≥ LJ is irrelevant to gravitational struc-
ture formation in cosmology and astrophysics, and is egregiously misleading in all of
its applications.
2. The correct criterion for gravitational structure formation is that L must be larger
than the largest Schwarz scale; that is, L ≥ (LSX)max, where X is D, V, M, T , de-
pending on whether diffusion or viscous, magnetic or turbulent forces limit the grav-
itational effects.
3. Structure formation began in the plasma epoch with protosupercluster to protogalaxy
decelerations.
4. Gravitational condensations began soon after the plasma-gas transition, forming
micro-brown-dwarfs, clustered in PGCs, that persist as the dominant dark matter
component of inner galactic halos (50 kpc).
8
Carl H. Gibson: The fluid mechanics of dark matter formation
5. The present fluid mechanical theory and its cosmological consequences regarding the
forms of baryonic and nonbaryonic dark matter (Gibson 1996) is well supported by
observations, especially the quasar-microlensing of Schild 1996 and his inference that
the lens galaxy mass of Q0957+561A,B is dominated by small rogue planets (inter-
preted here as PFPs).
6. Star-microlensing studies that rule out MBDs as the Galaxy missing mass
(Alcock et al. 1998, Renault et al. 1998), contrary to the quasar-microlensing evi-
dence and the present theory, are subject to extreme undersampling errors from
their unwarranted assumption of a uniform number density distribution, rather
than extremely intermittent lognormal distributions expected from nonlinear ag-
gregational cascades of such small objects as they form nested clusters and stars
(Gibson and Schild 1998b).
Acknowledgements. Numerous helpful suggestions were provided by Rudy Schild.
References
Alcock, C., et al. 1998, ApJ, 499, L9
Bahcall, J. N. 1997, in Unsolved Problems in Astrophysics, Eds. Bahcall, J. N. and Ostriker, J.
P., Princeton Univ. Press, NJ
Gibson, C. H. 1968, Phys. Fluids, 11, 2305
Gibson, C. H., W. T. Ashurst, and A. R. Kerstein 1988, J. Fluid Mech. 194, 261
Gibson, C. H. 1991, Proc. R. Soc. A, 433, 149
Gibson, C. H. 1996, Appl. Mech. Rev., 49, 299
Gibson, C. H. 1997a, The Identification of Dark Matter, Ed.: N. J. C. Spooner, World Scientific,
114
Gibson, C. H. 1997b, Dark Matter in Astro- and Particle Physics, Ed.: H. V. Klapdor-
Kleingrothaus and Y. Ramachers, World Scientific, 409
Gibson, C. H. and Schild, R. E. 1998a, to be published
Gibson, C. H. and Schild, R. E. 1998b, to be published
Jeans, J. H. 1902, Phil. Trans. R. Soc. Lond. A, 199, 1
Jeans, J. H. 1929, Astronomy and Cosmology, Cambridge Univ. Press, Cambridge, UK
Kolb, E. W. and M. S. Turner 1994, The Early Universe, Addison-Wesley Publishing Company
Low, C. and Lynden-Bell, D. 1976, MNRAS, 176, 367
Padmanabhan, T. 1993, Structure Formation in the Universe, Cambridge University Press,
Cambridge, UK
Peebles, P. J. E. 1993, Principles of Physical Cosmology, Princeton Univ. Press, Princeton, NJ
Renault, C. et al. 1998, A&A, 329, 522
Schild, R. 1996, ApJ, 464, 125
Tyson, J. A. and P. Fischer 1995, ApJ, 446, L55
Weinberg, S. 1972, Gravitation and Cosmology: Principles and Applications of the General
Theory of Relativity, John Wiley and Sons, New York
|
astro-ph/0511594 | 1 | 0511 | 2005-11-19T21:48:03 | A Spectroscopic Study of the Environments of Gravitational Lens Galaxies | [
"astro-ph"
] | (Abridged) We present the first results from our spectroscopic survey of the environments of strong gravitational lenses. The lens galaxy belongs to a poor group of galaxies in six of the eight systems in our sample. We discover three new groups associated with the lens galaxies of BRI 0952-0115 (five members), MG 1654+1346 (seven members), and B2114+022 (five members). We more than double the number of members for another three previously known groups around the lenses MG 0751+2716 (13 total members), PG 1115+080 (13 total members), and B1422+231 (16 total members). We determine the kinematics of the six groups, including their mean velocities, velocity dispersions, and projected spatial centroids. The velocity dispersions of the groups range from 110 +170, -80 to 470 +100, -90 km/s. In at least three of the lenses -- MG0751, PG1115, and B1422 -- the group environment significantly affects the lens potential. These lenses happen to be the quadruply-imaged ones in our sample, which suggests a connection between image configuration and environment. The lens galaxy is the brightest member in fewer than half of the groups. Our survey also allows us to assess for the first time whether mass structures along the line of sight are important for lensing. We first show that, in principle, the lens potential may be affected by line-of-sight structures over a wide range of spatial and redshift offsets from the lens. We then quantify real line-of-sight effects using our survey and find that at least four of the eight lens fields have substantial interloping structures close in projection to the lens, and at least one of those structures (in the field of MG0751) significantly affects the lens potential. | astro-ph | astro-ph | Draft version May 15, 2019
Preprint typeset using LATEX style emulateapj v. 6/22/04
A SPECTROSCOPIC STUDY OF THE ENVIRONMENTS OF GRAVITATIONAL LENS GALAXIES
Ivelina Momcheva and Kurtis Williams
Steward Observatory, 933 N. Cherry Ave., Tucson, AZ 85721
Department of Physics and Astronomy, Rutgers University, Piscataway, NJ 08854
Charles Keeton
and
Ann Zabludoff
Steward Obs., 933 N. Cherry Ave., Tucson, AZ 85721
Draft version May 15, 2019
ABSTRACT
−80 to 470+100
We present the first results from our spectroscopic survey1 of the environments of strong gravi-
tational lenses. The lens galaxy belongs to a poor group of galaxies in six of the eight systems in
our sample. We discover three new groups associated with the lens galaxies of BRI 0952−0115 (five
members), MG 1654+1346 (seven members), and B2114+022 (five members). We more than double
the number of members for another three previously known groups around the lenses MG 0751+2716
(13 total members), PG 1115+080 (13 total members), and B1422+231 (16 total members). We
determine the kinematics of the six groups, including their mean velocities, velocity dispersions, and
projected spatial centroids. For the newly discovered groups, we quantify these properties for the first
time. For the other three groups, the increased membership allows us to make more robust estimates
of the kinematic properties of the groups than previously possible. The velocity dispersions of the
groups range from 110+170
−90 km s−1. The higher velocity dispersions (for the richer groups
MG0751, PG1115, and B1422) are consistent with those of nearby X-ray luminous groups, while the
others (for the poorer groups BRI0952, MG1654, and B2114) are more typical of nearby dynamically
younger groups. The lens galaxy is the brightest member in fewer than half of the groups. In general,
the brightest group galaxy is an early-type galaxy that lies off the center of the potential and occupies
an orbit indistinguishable from the other group members. In at least three of the lenses -- MG0751,
PG1115, and B1422 -- the group environment significantly affects the lens potential. These lenses
happen to be the quadruply-imaged ones in our sample, which suggests a connection between image
configuration and environment. Finally, our survey allows us to assess for the first time whether mass
structures along the line of sight are important for lensing. We first show that, in principle, the lens
potential may be affected by line-of-sight structures over a wide range of spatial and redshift offsets
from the lens. We then quantify real line-of-sight effects using our survey and find that at least four
of the eight lens fields have substantial interloping structures close in projection to the lens, and at
least one of those structures (in the field of MG0751) significantly affects the lens potential.
Subject headings: gravitational lensing -- (galaxies:) quasars:
individual (MG 0751+2716, BRI
0952−0115, PG 1115+080, B1422+231, MG 1654+1346, PMN J2004−1349,
B2114+022, HE 2149−2745) -- galaxies: clusters: general -- galaxies: halos
5
0
0
2
v
o
N
9
1
1
v
4
9
5
1
1
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1. INTRODUCTION
2002; Rusin et al.
2005; Treu & Koopmans
The study of strong gravitational lens systems offers
critical constraints on the masses, shapes, evolution,
and substructure of galaxy dark matter halos (e.g.,
Kochanek 1991; Keeton et al. 1998; Metcalf & Madau
2003;
2001; Dalal & Kochanek
Rusin & Kochanek
2004;
Ferreras et al. 2005), on the Hubble constant indepen-
dent of the local distance ladder (e.g., Refsdal 1964;
Kochanek & Schechter 2003), and on the dark energy
density (e.g., Turner 1990; Kochanek 1996a; Chae 2003;
Linder 2004; Mitchell et al. 2005). However, our under-
standing of observed lenses is limited by uncertainties
and biases in the lens models necessary to analyze the
data. Despite improving data for lensed images and
1This paper includes data gathered with the 6.5 meter Magellan
Telescopes located at Las Campanas Observatory, Chile.
lens galaxies, astrophysical applications of lensing are
still hindered by poor knowledge of the environments in
which strong lens systems reside.
Several arguments suggest that lenses have complex
environments. Statistical arguments based on galaxy
demographics imply that at least 25% of lens galax-
ies lie in dense environments such as groups and clus-
ters (Keeton et al. 2000). From spectroscopic obser-
vations, several
lenses are in fact known to lie in
groups (MG 0751+2716, PG 1115+080, B1422+231,
and B1608+656; Tonry & Kochanek 1999; Kundi´c et al.
1997a,b; Tonry 1998; Fassnacht et al. 2004), and sev-
eral others in clusters (RX J0911+0551, Q0957+561,
HST 14113+5221, and MG 2016+112; Kneib et al. 2000;
Young et al. 1981b; Fischer et al. 1998; Soucail et al.
2001). Indirect evidence for the existence of other groups
comes from the large tidal shears required to explain the
image configurations of many four-image (quad) lenses,
2
Momcheva et al.
which presumably come from mass structures near the
lens galaxy or along the line of sight (Keeton et al. 1997).
The range of required shears in quad lenses could re-
flect a range of environment densities, running from
poor groups to rich clusters. Comparisons of the lens-
ing rate in different surveys have also been cited as ev-
idence that many lens galaxies probably lie in groups
(Blandford et al. 2001). Finally, theoretical models pre-
dict that lens galaxies reside in complex environments
that produce substantial shears, although it is not yet
clear whether the models predict shears large enough
to explain real quad lenses (Holder & Schechter 2003;
Dalal & Watson 2004).
If not handled properly, complex environments can in-
ject uncertainties and biases into the astrophysical quan-
tities derived from lens models (see Keeton & Zabludoff
2004, hereafter KZ04). For example, neglecting environ-
ment altogether leads to lens models that, for most pur-
poses,1 are simply wrong. Approximating environmental
effects with a simple shear term leads to models that are
better but still tend to overestimate the Hubble constant,
the velocity dispersion of the lens galaxy, and the dark
energy density ΩΛ, and to underestimate the magnifica-
tions of the lensed images.
In principle, modeling the
full richness of environmental effects can remove these
biases, and may also resolve the long-standing puzzle of
why quad lenses are almost as common as doubles in
statistically complete lens samples (see King et al. 1996;
Kochanek 1996b; Keeton et al. 1997; Rusin & Tegmark
2001; Cohn & Kochanek 2004, KZ04). Such an analysis
requires detailed knowledge of the galaxy populations,
velocity dispersions, and projected spatial centroids of
groups and clusters around lenses in order to determine
how the environments affect the lens potentials. To
date, such observations have mainly been carried out
for the few lenses that reside in clusters, which leaves
many lenses whose environments are known poorly or
not at all. Worse, existing observations cannot charac-
terize the distribution of lens environments, so we can-
not assess environment-related biases in statistical quan-
tities (such as ΩΛ or the quad/double ratio) or ensemble
properties (such as evolution or substructure). While
the environment distribution can be predicted from the-
oretical models (Keeton et al. 1997; Holder & Schechter
2003; Dalal & Watson 2004), disagreements among the
models, and discrepancies between the predicted distri-
butions and the shears required to fit observed lenses,
raise questions about the predictions.
These issues have not been adequately addressed
with observations, because no systematic survey of
lens environments exists.
Surveys of a few lens
fields have been published individually or in pairs
(Young et al. 1981b; Kundi´c et al. 1997a,b; Fischer et al.
1998; Tonry 1998; Tonry & Kochanek 1999; Kneib et al.
2000; Fassnacht & Lubin 2002; Fassnacht et al. 2004;
Soucail et al. 2001). In many cases, though, those sur-
veys only spanned a ∼ 30′′ field around each lens, so they
did not adequately sample group or cluster membership
out to the virial radius (∼ 0.7 Mpc for groups, corre-
1 The important exception is measurements of the total mass
within the Einstein radius, which are largely independent of as-
sumptions built into lens models (e.g., Kochanek 1991; Cohn et al.
2001).
sponding to ∼ 3′ at the redshifts of the lenses we study).
We have undertaken a systematic deep and wide-field
survey of lens fields, and here we present results for the
first eight systems that we have targeted for multi-object
spectroscopy. We characterize the environment within a
∼ 6′ diameter field around each of the eight lenses, and
quantify how those environments affect the lens poten-
tials.
Going beyond the lenses' immediate environments, we
also consider the degree to which massive structures
along the line of sight to a lens affect the lens potential.
The prevalence and importance of interloping structures
in lens fields is poorly understood. Observationally, there
appear to be bound groups along the lines of sight to
B0712+472 (10 members; Fassnacht & Lubin 2002) and
MG 1131+0456 (3 members; Tonry & Kochanek 2000).
Overdensities of galaxies are seen in the fields of sev-
eral other strong lenses (Faure et al. 2004; Morgan et al.
2005), but it is not yet known whether they indicate
massive bound structures, and whether any such struc-
tures are associated with the lens galaxies or lie elsewhere
along the lines of sight. On the theoretical side, stud-
ies have yielded conflicting results as to whether line-
of-sight structures are very important or negligible for
lensing (e.g., Seljak 1994; Bar-Kana 1996; Keeton et al.
1997; Premadi & Martel 2004). We show here that lenses
are, in principle, sensitive to structures over a wide range
of redshifts and projected spatial offsets, so the practi-
cal importance of interloping structures depends only on
how common they are. Our photometric/spectroscopic
pencil-beam survey of
lens fields enables us to self-
consistently identify any prominent structures at all rel-
evant redshifts, and to assess their actual contributions
to observed lenses.
Separate from lensing, an important by-product of our
survey is a sizable sample of poor groups at intermediate
redshifts. Only a few such samples are presently known
(Carlberg et al. 2001; Wilman et al. 2005; Gerke et al.
2005). Groups are important laboratories for studies of
galaxy evolution because they are the most common en-
vironments for galaxies, and are also relatively simple
systems in which the range of mechanisms thought to
drive galaxy evolution (primarily galaxy -- galaxy interac-
tions) is much narrower than in hotter, denser clusters
(Zabludoff & Mulchaey 1998a, hereafter ZM98). Unfor-
tunately, poor groups are notoriously difficult to identify
using conventional methods for finding clusters, due to
their low projected surface densities, faint X-ray lumi-
nosities, and inefficiency for weak lensing. The veloc-
ity dispersions of nearby groups range from σr ∼ 200
km s−1 for systems that are X-ray faint, late-type dom-
inated, dynamically young, and generally similar to the
Local Group; to σr ∼ 300 -- 500 km s−1 for systems that
are richer, X-ray luminous, early-type dominated, and
dynamically more evolved; up to σr ∼ 1000 km s−1 for
rich clusters.
In nearby X-ray luminous groups with
σr ∼ 300 -- 500 km s−1, there is always a giant elliptical
that lies at the center of the group potential, which sug-
gests that such galaxies form in groups via interactions
prior to being accreted by rich clusters (ZM98). Groups
at intermediate redshifts like those we describe here will,
in conjunction with nearby group samples, permit us to
observe the evolution of groups directly.
The organization of this paper is as follows. In §2 we
Lens Galaxy Environments
3
describe our sample of eight lens systems and summarize
previous work on them.
In §3 we present our spectro-
scopic data in the eight lens fields. In §4 we determine
the membership, kinematics, and centroids of the groups,
and use those properties to quantify how the environ-
ments affect the lens models. We also explore the effects
of line-of-sight structures on the lens models. We sum-
marize our results and conclusions in §5. We present the
formalism for computing the convergence and shear aris-
ing from perturbing structures anywhere along the line
of sight in an Appendix. Where necessary, we assume a
cosmology with ΩM = 0.3, ΩΛ = 0.7, and H0 = 70 km
s−1 Mpc−1.
2. THE SAMPLE
Our sample consists of eight known gravitational lens
systems with lens galaxies at intermediate redshifts
0.25 < zl < 0.5. Four of the lenses (MG 0751+2716,
PG 1115+080, B1422+231, and MG 1654+1346) were
suspected from previous studies to have complex environ-
ments. We chose the other four lenses (BRI 0952−0115,
PMN J2004−1349, B2114+022, and HE 2149−2745) be-
cause of the availability of prior imaging and photometry
as well as accessibility from Las Campanas Observatory.
In the remainder of this section we briefly review prior
studies of these eight lenses. The data from this section
are summarized in Table 1.
MG 0751+2716 (hereafter MG0751), discovered as a
part of the MIT -- Greenbank -- VLA search for gravita-
tional lenses,
is a radio lens with four images and a
partial ring (Leh´ar et al. 1993). Optical imaging of the
system by Leh´ar et al. (1997) identified an R = 21.3
galaxy (G3) located 0.2′′ northeast of the brightest ra-
dio spot as the likely lens galaxy. G3 is a satellite
of a much brighter R = 19.1 galaxy (G1) located 6′′
away. Tonry & Kochanek (1999) determined the red-
shifts of the galaxies to be zG1 = 0.3501 ± 0.0003 and
zG3 = 0.3502 ± 0.0003. They also found a nearby emis-
sion line galaxy to have redshift 0.3505 ± 0.0003, indi-
cating that the lens galaxy lies in a small group with at
least three members. Lens models by Leh´ar et al. (1997)
suggest that MG0751 requires more external shear that
can be accounted for by the observed galaxies, which
is consistent with the hypothesis that the lens environ-
ment is complex. The redshift of the source quasar is
zs = 3.200 ± 0.001 (Tonry & Kochanek 1999).
BRI 0952−0115 (hereafter BRI0952) was discovered by
McMahon & Irwin (1992) as a doubly imaged zs = 4.5
optical quasar. The quasar is also detected at millimeter
wavelengths (Omont et al. 1996). Keeton et al. (1998)
found that the lens is a flattened early type galaxy,
and Kochanek et al. (2000) estimated a lens redshift of
zl = 0.41 ± 0.05 based on fundamental plane fitting. Be-
cause the separation between the images is small (0.9′′)
and the lens galaxy is faint (21.9 in F675W; Keeton et al.
1998), the lens redshift has not been determined spectro-
scopically.
PG 1115+080 (hereafter PG1115) is a lens system dis-
covered by Weymann et al. (1980), in which a radio-quiet
quasar at redshift zs = 1.722 is lensed into four im-
ages (Hege et al. 1981). The lens galaxy was first de-
tected by Henry & Heasley (1986); its redshift was esti-
mated by Angonin-Willaime et al. (1993), and later im-
proved by Kundi´c et al. (1997a) and Tonry (1998) to zl =
0.3098±0.0002. Young et al. (1981a) suggested the pres-
ence of a small group of galaxies near the lens. This was
confirmed by Kundi´c et al. (1997a) and Tonry (1998),
who measured the redshifts of a total of four galaxies
within 20′′ of the lens galaxy. Kundi´c et al. (1997a) es-
timated a group velocity dispersion of σr = 270 ± 70
km s−1 from four galaxies, while Tonry (1998) estimated
σr = 326 km s−1 from a slightly different set of four
galaxies. Grant et al. (2004) detected diffuse X-ray emis-
sion that is associated with the group and that has a
temperature kT ∼ 0.8 ± 0.2 keV; this value is consis-
tent with typical values for low-redshift poor groups, but
somewhat high given the measured group velocity disper-
sion and the local σr-TX relation (Mulchaey & Zabludoff
1998, hereafter MZ98). PG1115 is one of nine known
strong lens systems for which the time delay between
different images has been measured, so it can be used
to determine H0. Schechter et al. (1997) measured the
light curves of the different images and estimated the
time delays, and Bar-Kana (1997) then reanalyzed the
data to give more precise results: the delay between im-
ages B and C is tBC = 25.0+3.3
−3.8 days, and the ratio of
the delays between A (actually a combination of the close
images A1 and A2), B, and C is tAC /tBA = 1.13+0.18
−0.17.
PG1115 is one of the lenses with "anomalous" flux ratios
thought to indicate some sort of small-scale structure
in the lens galaxy (e.g., Metcalf & Madau 2001; Chiba
2002; Dalal & Kochanek 2002; Keeton et al. 2005a).
B1422+231 (hereafter B1422) is a four-image lens
discovered by Patnaik et al.
(1992) while searching
for small-separation lenses among flat spectrum ra-
dio sources in the Jodrell Bank -- VLA Astrometric Sur-
vey (JVAS; Patnaik et al. 1992a; Browne et al. 1998;
Wilkinson et al. 1998; King et al. 1999). The source
is a radio loud quasar at zs = 3.62 (Patnaik et al.
1992), and the lens is a luminous elliptical at zl =
0.3374 (Impey et al. 1996; Kundi´c et al. 1997b). The
lens galaxy and five nearby galaxies form a group
at zg = 0.338 with a rest-frame line-of-sight veloc-
ity dispersion of σr = 550 ± 50 km s−1 (Kundi´c et al.
1997b).
Lens models for B1422 require a signifi-
cant shear γ ∼ 0.20 -- 0.26, which may be attributable
to the group environment (Hogg & Blandford 1994;
Keeton et al. 1997; Dobler & Keeton 2005).
Indeed,
from their estimate of the group's velocity dispersion and
centroid, Kundi´c et al. (1997b) estimated γ = 0.23 and
pointed out that the group will also create some conver-
gence κ that may affect the lens potential. Grant et al.
(2004) detected B1422 in X-rays (0.5 -- 2 keV) and de-
termined a temperature of kT = 1.0+∞
−0.3 keV, which
is consistent with the value expected for a poor group
(MZ98). B1422 is another lens with "anomalous" flux
ratios (Mao & Schneider 1998; Chiba 2002).
MG 1654+1346 (hereafter MG1654) was origi-
nally detected in the MIT -- Greenbank -- VLA survey.
Langston et al. (1988, 1989) recognized its unusual struc-
ture in a VLA snapshot and obtained radio and optical
mapping. The source is a zs = 1.74 radio quasar with a
compact core and two extended radio lobes. The south-
west lobe is lensed into a ring by a zl = 0.254 giant
elliptical galaxy (Langston et al. 1988; Kochanek et al.
2000). Langston et al. (1989) noted an enhancement of
the number density of galaxies near the lens; some of the
4
Momcheva et al.
TABLE 1
Gravitational Lens Galaxies
Lens
RAb
Decb
zl
(J2000)
MG 0751+2716
BRI 0952−0115
PG 1115+080
B1422+231
MG 1654+1346
PMN J2004−1349
B2114+022
07:51:41.46 +27:16:31.4
09:55:00.01 −01:30:05.0
11:18:17.00 +07:45:57.7
14:24:38.09 +22:56:00.6
16:54:41.83 +13:46:22.0
20:04:07.07 −13:49:30.7
21:16:50.75 +02:25:46.9
HE 2149−2745
21:52:07.44 −27:31:50.2
0.349a
(0.41)c
0.31b
0.34b
0.254a
-
0.316a
0.59b
0.50b
Ib
[mag]
21.26
21.21
18.92
19.66
17.9
-
18.63
b
zs
3.20
4.50
1.72
3.62
1.74
-
-
∆te
[days]
-
-
25.0±2.0
-
-
-
-
Imagesb, f
R
2
4
4
R
2
2+2
e
kTX
[keV]
-
-
0.8±0.2
1.0+ inf
−0.3
-
-
-
19.56
2.03
103.0±12.0
2
-
Ngrp
d, e
2
-
4
5
-
-
-
-
e
σr
[km s−1]
-
-
270±70
550±50
-
-
-
-
aData from this work
bData from references in the text and from the CASTLES website (http://cfa-www.harvard.edu/castles/).
cPhotometric redshift (Kochanek et al. 2000)
dNumber of previously known group members in addition to the lens galaxy.
eReferences in text.
fR means an Einstein ring.
nearby galaxies are comparable in brightness to the lens
galaxy, suggesting a complex environment.
PMN J2004−1349 (hereafter PMN2004) is a two-image
lens discovered in a search for radio lenses in the south-
ern sky (Winn et al. 2001). The radio spectral index
of the images is typical for radio-loud quasars, so the
source is considered to be a quasar despite the lack of an
optical spectrum and a measured redshift (Winn et al.
2001). Based on photometry, Winn et al. (2001) sug-
gested a lens redshift in the range 0.5 < zl < 1.0. Higher-
resolution imaging by Winn, Hall & Schechter (2003) re-
vealed a spiral lens galaxy (only the fifth one known) and
showed that the color differences between the two im-
ages at optical and near-infrared wavelengths can be ex-
plained by differential extinction. The extinction analy-
sis can be used to infer the lens redshift; it seems to imply
somewhat low values (0.03 . zl . 0.36), but that result
depends on assumptions about the extinction curve.
B2114+022 (hereafter B2114) was discovered as part
of the search for lenses in JVAS (King et al. 1999). Ra-
dio maps show four distinct components within 2.4′′ in
a configuration that is atypical for lenses. Furthermore,
the sources can be divided into two pairs with distinct
radio surface brightnesses and radio spectra: sources A
and D are similar to each other, and sources B and C
are similar to one another, but the two pairs are clearly
different. Ground-based and HST optical imaging and
spectroscopy do not detect the lensed images but re-
veal two lens galaxies at zl1 = 0.3157 and zl2 = 0.5883
(Augusto et al. 2001), suggesting a complex lensing ge-
ometry. Chae, Mao & Augusto (2001) could explain two
of the radio components (A and D) as lensed images us-
ing a two-plane lens model.
It is not known whether
the other components (B and C) are images of the same
source (unlikely), images of a different source, or struc-
ture related to the G1 lens galaxy. No lens models that
explain these components have been published.
HE 2149−2745 (hereafter HE2149) is a doubly im-
aged broad absorption line quasar at redshift zs =
2.033, which was discovered by Wisotzki et al. (1996)
in the Hamburg/ESO wide-angle survey for bright
quasars. Burud et al. (2002) reported a redshift of
zl = 0.495 ± 0.01 for the elliptical lens galaxy, consistent
with the photometric redshift estimate of zl = 0.43+0.07
−0.06
(Kochanek et al. 2000). Burud et al. also measured the
time delay between the two images to be ∆t = 103 ± 12
days. Based on the large number of red non-stellar
objects in R-band images of the field around the lens,
Lopez, Wucknitz & Wisotzki (1998) suggested that the
lens galaxy might be a member of a cluster.
3. SPECTROSCOPIC DATA
We first identified galaxies for follow-up spectroscopy
from two-color, wide-field imaging of each lens field. We
obtained deep images in I and either V or R during the
period from May 2002 to June 2004 using the 36′ × 36′
Mosaic Imager on the 4-m telescopes at Kitt Peak Na-
tional Observatory and Cerro Tololo Inter-American Ob-
servatory. We reduced these images and extracted photo-
metric catalogs following standard methods using IRAF2
and SExtractor (Bertin & Arnouts 1996). A more de-
tailed description of the photometric analysis is presented
in a separate paper (Williams et al. 2005).
We selected spectroscopy targets using a prioritization
scheme based on objects' colors and projected distances
from the lens. Highest priority was given to objects
populating a red sequence in the color -- magnitude dia-
gram that is consistent with the lens redshift, and to
targets that lie within a group-like virial radius of 0.7
Mpc (ZM98), which corresponds to ∼ 3′ over the red-
shift range of our lenses. We applied a magnitude cut
at I = 21.5 to assure reasonable exposure times. This
limiting magnitude corresponds to I ∗ + 4.2 at the low
redshift limit of our sample (z = 0.25) and to I ∗ + 2.8 at
the high redshift limit (z = 0.5), where I ∗ is the observed
magnitude of an L∗ galaxy, adopted from Williams et al.
(2005). We obtained multislit spectroscopy during two
observing runs, March 1 -- 4 and August 30 -- September
2 IRAF is distributed by the National Optical Astronomy Ob-
servatories, which are operated by the Association of Universities
for Research in Astronomy, Inc., under cooperative agreement with
the National Science Foundation.
Lens Galaxy Environments
5
2, 2003, with the Low Dispersion Survey Spectrograph
(LDSS-2; Allington-Smith et al. 1990) at the 6.5-m Mag-
ellan 2 (Clay) telescope at Las Campanas Observatory.
All spectra were taken with the medium blue grism (300
l/mm; 5000A blaze) over a wavelength range of 3900 --
8000A. We used 1.03′′ slitlets, resulting in a spectral res-
olution of ≈ 15A FWHM. Each slitmask had dimensions
∼ 5′ × 7′, typically contained 20 -- 30 targets, and was ob-
served for 4×900 s.
Our sky coverage is shown in Figure 1. Each panel
spans 15′ × 15′ and is centered on the lens galaxy. Rect-
angles show the outlines of our slitmasks. There is a
noticeable difference in the sampling between the two ob-
serving runs. In March 2003, when we observed MG0751,
BRI0952, PG1115, B1422, and MG1654, the masks lay
mostly on top of one another, providing exhaustive cov-
erage of the immediate surroundings of each lens galaxy.
Because most of our fields were far north for Magellan,
we aligned the masks with the paralactic angle to min-
imize the effects of atmospheric dispersion. In contrast,
the masks for MG1654, PMN2004, B2114, and HE2149
in the August 2003 observing run were tiled to cover a
larger area of the sky but still overlap significantly in
the 3′ projected radius around the lens. No attempt was
made to align the masks with the paralactic angle be-
cause the targets were close to or south of the celestial
equator and were therefore observed at relatively low air-
mass.
The figure shows that our sampling is very good within
3′ around the lens for all fields, with a minor excep-
tion in B1422. We observed four masks for each of
MG0751, BRI0952, PG1115 and B1422, five masks for
each of PMN2004, B2114 and HE2149, and seven masks
for MG1654 (which was observed during both runs). We
discuss our spectroscopic completeness below.
We reduced all spectra using standard IRAF proce-
dures and corrected them to the local standard of rest
using the IRAF routines RVCORRECT and DOPCOR.
We determined the radial velocities using the cross-
correlation of absorption lines (XCSAO) and/or using
emission line identifications (EMSAO)3. If both emis-
sion and absorption line velocities were found, the quoted
value is a weighted average of the two. Marc Postman
kindly provided template galaxy spectra (Postman et al.
2002, and private communications). We visually in-
spected every fit to ensure accuracy.
The number of objects observed in each field is listed
in Table 2. Galaxies represent ∼42% of the objects tar-
geted spectroscopically, while stars originally misclassi-
fied as galaxies are ∼22%. "Failed" targets (∼ 32%) are
those for which were unable to obtain velocities. The
majority of failed targets, especially at fainter magni-
tudes, were absorption-line systems for which the signal-
to-noise was too low to allow successful cross-correlation.
Other causes of failures are low surface brightness or poor
astrometry (more problematic in March than August).
The fraction of stars was significantly lower in August
(15%) than in March (26%), thanks to improvements
in star-galaxy separation made between the two runs
(see Williams et al. 2005). The large number of stars
in PMN2004 is due to the high stellar density at this
3 XCSAO and EMSAO are routines in the RVSAO IRAF package
(Kurtz & Mink 1998)
relatively low galactic latitude and longitude (l = 28◦,
b = −22◦).
Figure 2 shows the completeness of our spectroscopy
with respect to the photometry. The solid-line histogram
shows the magnitude distribution for all galaxies in the
photometric catalog projected within 3′ of the lens, while
the shaded histogram shows the subset for which we de-
termined velocities. Our target selection scheme, based
on colors and projected offsets from the lens, misses a
few of the brightest galaxies (many of which are likely to
be foreground objects). We miss a larger fraction of ob-
jects at the faint end, but these are very few and/or less
massive galaxies, which would significantly affect the lens
potential only if very close to the lens galaxy. In total,
there are only four galaxies within 10′′ of the lens galax-
ies (the zone in which a small perturber can have even a
moderate effect; see Fig. 6 below) that are present in our
photometric catalog but have no determined velocities.
Two of those (in B2114 and PMN2004) are below our
spectroscopic magnitude limit of I = 21.5. The other
two galaxies are in the field of HE2149, have I magni-
tudes of 19.97 and 21.08, and lie 7′′ and 10′′ respectively
from the lens. In color -- magnitude space (Williams et al.
2005) they lie on red sequences identified as line-of-sight
structures at z = 0.45 and z = 0.60 (see §4.5).
We were never able to put slits on all of our highest-
priority objects; this limitation is inherent to multislit
spectroscopy. For lensing purposes, the main effect of
spectroscopic incompleteness is to cause us to underesti-
mate environment-related lensing biases (see §§4.4 -- 4.5).
In particular, the fact that we prioritize galaxies thought
to lie at the lens redshift (color selection) means that we
undersample line-of-sight structures and hence underes-
timate their contributions to the lens potential.
As already mentioned, we improved our photometric
catalogs after some of the spectroscopic targets were se-
lected. As a result, there are 15 galaxies whose velocities
we measured that do not actually appear in the final pho-
tometric catalog. (These galaxies lie under bleed trails in
the imaging, so accurate photometry is not possible.) In
addition, there are another 25 galaxies whose velocities
we measured that are (mis-)classified in the final pho-
tometric catalog as stars. We omit all of these galaxies
when comparing the spectroscopic and photometric cat-
alogs for the purpose of understanding our spectroscopic
completeness (i.e., they are excluded from Fig. 2). How-
ever, these remain a part of our spectroscopic catalog
and all analyses based on that catalog.
To estimate the zero-point velocity correction and ex-
ternal velocity errors, we cross-correlate 403 sky spectra
extracted from our data with the same templates used for
the galaxy spectra and find a mean velocity of ¯υ = 40±50
km s−1. We also determine the velocities of 153 of the
serendipitously observed stars and find ¯υ = 30 ± 180
km s−1. Both methods give mean velocities comparable
to or smaller than the dispersion, and much smaller than
the velocities of the objects in the sample. We therefore
conclude that no zero point correction is needed.
Table 3 lists our spectroscopic catalog. For each entry
we give the catalog name, J2000 coordinates calibrated
to USNO-B2.0, projected distance from the lens in ar-
cmin, aperture magnitude within a fixed physical size of
∼6.5 kpc (see Williams et al. 2005), the heliocentric ra-
dial velocity, velocity error, redshift and redshift error.
6
Momcheva et al.
5
0
-5
5
0
-5
-5
0
5
-5
0
5
-5
0
5
-5
0
5
Fig. 1. -- Sky coverage for LDSS-2 multislit spectroscopy in the fields of the eight lens systems in our sample. Each panel is 15′ × 15′,
centered on the lens galaxy. Rectangles show the positions of our slit masks. Each slitmask covers ∼ 5′ × 7′ and includes 20 -- 30 slitlets
parallel to the long side of the mask. The 3′ circle around the lens galaxy corresponds to ∼ 0.7 Mpc, a group-like virial at the redshifs of
the lenses in our sample. We observed four masks for each of MG0751, BRI0952, PG1115, and B1422, five masks for each of PMN2004,
B2114, and HE2149, and seven masks for MG1654. The masks from the first observing run lack a particular sampling pattern but provide
excellent coverage of the expected group virial radius in all cases except B1422, where a small portion remains unsampled. The masks
from the second observing run were tiled in an attempt to maximize the sky coverage while still providing good sampling of the immediate
surroundings of the lens.
TABLE 2
LDSS-2 Observations
Lens
Date
# Galaxiesa # Stars # Failed Total
MG0751
BRI0952
PG1115
B1422
MG1654
March 2003
March 2003
March 2003
March 2003
March 2003
August 2003
PMN2004 August 2003
B2114
August 2003
August 2003
HE2149
aIncluding QSOs.
38
47
47
53
39
20
41
38
41
24
24
28
14
30
6
43
1
8
27
19
11
26
32
35
35
46
66
89
90
86
93
101
61
119
85
115
The last column describes the method from which the
velocity was obtained: 1 for absorption lines, 2 for emis-
sion lines, or 3 for a combination of both. Missing data
means that the object was not in the final photometric
catalog as described above. Such objects have no iden-
tification names and are numbered successively starting
with 90001.
4. RESULTS AND DISCUSSION
4.1. Group Membership
The environments of most strong lenses are not well
characterized. Our first goal is to determine whether
each lens galaxy lies in a group or cluster, and if so to
identify the other member galaxies. Even in cases where
groups were already identified (MG0751, PG1115, and
B1422), the number of group members known previously
ranged from three to six. Our deep, wide-field spectro-
scopic sampling has the potential not only to find new
groups, but also to increase the membership of known
lens groups to the point where robust determinations of
the group velocity dispersions and centroids are possi-
ble. These are essential for understanding how a group
affects the lens potential, as discussed in §4.4. Addition-
ally, a more complete inventory of the brightest (most
massive) group members, and their contributions to the
lens potential, will also greatly improve lens models.
We present the redshift histograms for all eight of our
lens fields in Figure 3. In the left panels, the shaded his-
tograms include all galaxies that lie within a projected ra-
dius of 3′ (a group-like virial radius) around the lens, and
Lens Galaxy Environments
7
Fig. 2. -- Apparent magnitude histograms of galaxies in the photometric catalog (solid line), and of those galaxies within 3′ of the lens
for which redshifts were obtained (shaded). Our target selection scheme, based on colors and projected offsets from the lens, misses a few
of the brightest galaxies. We miss a larger fraction of objects at the faint end, but these are presumably less massive galaxies that do not
contribute significantly to the lens potential.
Lens Field Galaxy Properties -- EXAMPLE
TABLE 3
ID
α
δ
[hh:mm:ss]
[dd:mm:ss]
b
[']
I
[mag]
∆cz
cz
[km s−1]
z
∆z
Spectral
Type
MG0751: MWKZ GAL
... 9809
... 9739
... 9726
... 9582
... 9606
... 9570
... 9560
... 9333
... 9297
... 9274
... 9225
... 9239
... 9100
... 9047
... 9120
... 9049
... 8794
... 9006
... 8816
... 8669
... 8682
... 8673
... 8476
... 7921
... 8385
... 8257
... 8288
07:51:30.20
07:51:30.77
07:51:31.00
07:51:32.04
07:51:32.28
07:51:32.49
07:51:32.73
07:51:34.61
07:51:34.66
07:51:35.24
07:51:35.43
07:51:35.75
07:51:36.66
07:51:36.96
07:51:36.98
07:51:37.73
07:51:37.92
07:51:38.04
07:51:40.32
07:51:41.07
07:51:41.50
07:51:41.84
07:51:43.25
07:51:43.26
07:51:44.18
07:51:45.32
07:51:45.50
27:14:43.1
27:17:55.7
27:14:16.7
27:12:59.2
27:12:58.9
27:17:38.6
27:14:42.1
27:15:45.5
27:13:37.7
27:17:37.1
27:17:07.9
27:15:22.8
27:19:39.8
27:19:27.9
27:18:40.6
27:18:06.5
27:16:12.9
27:17:33.7
27:16:22.1
27:19:43.7
27:16:31.9
27:16:29.2
27:17:53.3
27:16:06.3
27:16:39.7
27:17:02.4
27:18:51.4
3.10
2.76
3.24
3.30
4.10
2.29
2.67
1.71
3.28
1.76
1.48
1.72
3.31
3.10
2.37
1.78
0.85
1.28
0.31
3.20
0.00
0.09
1.41
0.58
0.61
0.99
2.49
20.9
20.6
20.5
19.0
20.8
20.2
21.0
19.5
18.4
19.6
19.0
20.9
18.4
18.3
20.9
20.4
0.0
20.1
20.9
18.9
20.1
21.5
19.8
0.0
20.7
19.4
20.4
168070
105970
175630
91080
106950
104680
106300
91220
28910
105350
79820
104960
74510
104220
72100
71940
167590
168620
104940
147320
104810
167090
169470
124780
112430
105110
60380
30
70
50
30
100
30
30
50
40
110
60
80
50
40
60
50
100
90
100
120
120
10
210
30
200
60
100
0.56023
0.35323
0.58543
0.30360
0.35650
0.34893
0.35433
0.30407
0.09637
0.35117
0.26607
0.34987
0.24837
0.34740
0.24033
0.23980
0.55863
0.56207
0.34980
0.49107
0.34937
0.55697
0.56490
0.41593
0.37477
0.35037
0.20127
0.000083
0.000237
0.000173
0.000113
0.000320
0.000113
0.000113
0.000153
0.000133
0.000380
0.000213
0.000280
0.000160
0.000140
0.000203
0.000163
0.000320
0.000307
0.000323
0.000390
0.000410
0.000037
0.000687
0.000103
0.000657
0.000200
0.000317
2
2
3
3
1
3
2
1
1
1
3
1
1
2
2
2
1
1
1
1
1
2
1
3
1
1
1
Note. -- This table is published in its entirety in the electronic edition. A portion is shown here
for guidance regarding its form and content. Objects with missing data were not found in our final
photometric catalog for reasons explained in the text.
8
Momcheva et al.
Ntot indicates the number of galaxies in the histogram.
This number excludes a few high-redshift AGNs that fall
outside the range of the plot. The histograms include
all galaxies with spectroscopic redshifts both from our
sample and from the literature. In particular, we have
added the following 16 galaxies to our catalog: G1, G6,
and G7 in MG0751 (Tonry & Kochanek 1999); the lens
galaxy GL as well as Gx in PG1115 (Tonry 1998); the
lens galaxy G1 as well as G2, G3, G4, G6, G8, G9, G10,
and Gx in B1422 (Kundi´c et al. 1997b; Tonry 1998); the
zl = 0.59 lens galaxy in B2114 (Augusto et al. 2001); and
the zl = 0.495±0.01 lens galaxy in HE2149 (Burud et al.
2002). We have not included the lens galaxy in BRI0952
(because it only has a photometric redshift estimate) or
in PMN2004 (because no good redshift estimate exists).
The vertical line shows the position of the lens galaxy
as listed in Table 1. In six cases (see Table 4), there is
clearly a peak in redshift space at or near the lens galaxy
redshift.
We determine the group membership by applying
a pessimistic 3σ clipping algorithm (as suggested by
Yahil & Vidal 1977) to any redshift peak containing a
lens galaxy. This procedure removes the galaxy most de-
viant from the mean redshift and recalculates the mean
and velocity dispersion of the distribution. If the omit-
ted galaxy is more than three new standard deviations
away from the recomputed mean, it is rejected. This loop
is executed until an omitted galaxy is not rejected. We
use statistical bi-weight estimators, which are more ro-
bust for small sample sizes than the standard estimators
(Beers et al. 1990), to calculate the location (mean red-
shift) and scale (velocity dispersion). In Figure 3, Ngrp
is the number of group members determined by the 3σ
clipping algorithm.
The right-hand panels in Figure 3 show a cut within
±5000 km s−1 of the lens galaxy velocity to present a bet-
ter view of the lens groups themselves. The peak height
is lower than in the full histogram because the bin size
is smaller, but groups are still easily recognizable in six
cases: MG0751, BRI0952, PG1115, B1422, MG1654 and
B2114. The groups in BRI0952, MG1654, and B2114
are new discoveries. For the previously known groups,
we have increased the number of group members from
three to 13 in MG0751, from five to 13 in PG1115, and
from six to 16 in B1422. The projected spatial distribu-
tions of galaxies in the six groups are shown in Figure
4.
In the case of BRI0952, we suggest that the lens
galaxy may belong to the five-member group at zg =
0.422, which is consistent with its photometric redshift
estimate of zl = 0.41 ± 0.05.
In B1422, the group
seems to consist of two clumps -- one around the lens
galaxy and the other to the northeast of it (see Figure
4). Zabludoff & Mulchaey (1998b) have found analogous
substructure in nearby groups. B1422 is our best sam-
pled group (16 members) and we expect that increasing
the membership of the other groups might reveal similar
clumpiness. In B2114, the group is associated with the
foreground lens galaxy (the first vertical dashed line).
There is a second peak slightly in front of the group
around the lens galaxy, but the galaxies are significantly
offset from the lens on the sky and thus not important
for lensing. We classify as "group members" only the
galaxies in the peak at the lens redshift.
Our spectroscopic findings agree well with the expec-
tations set by our photometry (see Williams et al. 2005).
All six lenses where we find groups show a compelling
red sequence at the lens redshift. In HE2149, the color --
magnitude diagram shows a well defined red sequence
corresponding to z ∼ 0.28, where we see a prominent
line-of-sight structure in Figure 3. In PMN2004, the lack
of compelling structures in either the color -- magnitude
diagram or the redshift histogram suggests that there
are no significant structures along the line of sight, al-
though it is difficult to draw firm conclusions from non-
detections (especially given incomplete spectroscopy). In
any case, the agreement between our photometric and
spectroscopic results reassures us that we understand the
data and their implications. These results are discussed
further in Williams et al. (2005).
4.2. Group Kinematics
The group velocity dispersion provides a key observ-
able probe of the group potential and its effect on
lens models. Group velocity dispersions based on small
member catalogs are uncertain and may be biased, be-
cause poor sampling of the underlying velocity distribu-
tion tends to underestimate the true velocity dispersion
(ZM98). Our more extensive member catalogs now al-
low us to measure the group velocity dispersions more
accurately than was previously possible.
To determine the mean velocities, ¯υ, and line-of-sight
velocity dispersions, σr, of the six groups in our sam-
ple, we use bi-weight estimators of location and scale
(Beers et al. 1990) because of their superiority at de-
weighting tails in the velocity distribution. In BRI0952,
which has only five members, the bi-weight estimator
routine fails, so we use the standard method for cal-
culating the mean velocity and velocity dispersion from
Danese et al. (1980). In general, standard and bi-weight
methods yield similar means and velocity dispersions.
We use all known group members, including the 10 found
in the literature, to determine these kinematic properties.
We apply a standard 1/(1 + zg) cosmological correction
to the velocity dispersions (Beers et al. 1990). The kine-
matic properties of the groups are presented in Table 4.
The six groups have velocity dispersions ranging from
110+170
−80 to 470+100
−90 km s−1. In the nearby universe, this
range of velocity dispersions describes systems running
from dynamically young, unrelaxed systems like the Lo-
cal Group, up to more dynamically relaxed, X-ray lumi-
nous groups. We are probably seeing a similar range of
groups in our intermediate-redshift sample. It is impor-
tant to note that the total number of galaxies for which
velocities were obtained is similar in all six fields, and
that Ngrp is roughly correlated with σr, suggesting that
the differences between the measured σr values are real
and not due to variable sampling.
As a further check on the accuracy of our group ve-
locity dispersions, we can determine the X-ray tempera-
ture that would be derived by combining our measured
σr values with the σr-TX relation for nearby groups and
clusters (ZM98), and compare that with the temperature
measured directly from X-ray observations. Only two of
the groups in our sample have been observed in X-rays
with Chandra: PG1115 and B1422 (Grant et al. 2004).
The X-ray temperatures expected from the σr-TX rela-
tion (1.5 keV and 1.7 keV, respectively) are consistent
Lens Galaxy Environments
9
TABLE 4
Group Kinematic Properties
Lens
Ntot Ngrp
αcen
σα,cen
[′′]
δcen
σδ,cen
[′′]
υmin
υmax
[ km s−1]
¯υ
δ ¯υ
[ km s−1]
z
σr,grp
δσr,grp
[ km s−1]
MG0751
BRI0952
PG1115
B1422
MG1654
B2114
39
44
48
57
59
38
13
5
13
16
7
5
07:51:40.7
09:54:56.1
11:18:16.8
14:24:41.0
16:54:39.3
21:16:51.4
±11 +27:16:53 ±9
±34 −01:29:58 ±27
±11 +07:45:36 ±9
±11 +22:55:42 ±9
±27 +13:47:15 ±20
±34 +02:10:59 ±27
104220
125000
92120
100640
75390
93000
105970
127000
93960
102810
76210
95000
104980 ±100
±30
126510
92970 ±110
101540 ±130
75750 ±100
94240
±80
0.3499
0.4217
0.3090
0.3385
0.2525
0.3141
320
170a
440
470
200
110
+170
−110
+150
−100
+90
−80
+100
−90
+120
−80
+170
−80
aVelocity dispersion calculated in the manner of Danese et al. (1980), instead of using the statistical bi-weight estimator of scale
(Beers et al. 1990). See text.
Fig. 3. -- (Left) Galaxy redshift distributions of the eight fields in our sample. The bin size is 1000 km s−1. The shaded histogram
includes all galaxies that lie within a projected radius of 3′ about the lens. The vertical dashed lines show spectroscopic lens galaxy
redshifts from the literature. The vertical dotted line for BRI0952 shows a photometric estimate of the lens galaxy redshift. In PMN2004,
the vertical dot-dashed lines show two different model-implied estimates of the lens galaxy redshift (Winn, Hall & Schechter 2003). Ntot is
the total number of galaxies included in the histogram, while Ngrp is the total number of group members. (Right) A close-up of the range
±5000 km s−1 centered on the mean group velocity. The bin size is 500 km s−1. The shaded histogram shows confirmed group members.
within the 95% confidence limit with the observed val-
ues (0.8 keV and 1.0 keV, respectively). This suggests
that our values of σr for these groups are reasonable.
4.3. Group Centroids
The projected offset of the lens galaxy from the group
centroid is another key ingredient in estimating the con-
tribution of a group to the lens potential. The position
of the brightest group galaxy relative to the spatial and
kinematic centroid of the group is also an important con-
10
Momcheva et al.
Fig. 3. -- continued.
straint on models of giant elliptical formation (ZM98).
(The lens galaxy may or may not be the brightest group
galaxy, as discussed below.)
In this section we calcu-
late the projected spatial and kinematic centroids of the
groups by averaging the sky positions and velocities, re-
spectively, of all group members. For members whose
velocities were added from the literature, we use coordi-
nates from our own photometric catalog when possible in
order to maintain a consistent coordinate system. Nei-
ther the projected spatial centroid nor the mean velocity
is weighted by the luminosity, because we do not want
to introduce an a priori bias toward the brightest group
galaxy or assume implicitly that the mass-to-light ratios
for all group members are the same. Nevertheless, the
luminosity-weighted centroids are within 2σ of the un-
weighted centroids for all groups except B1422 (where
four bright galaxies close to the lens pull the luminosity-
weighted centroid 3.5σ away from the unweighted cen-
troid).
4.3.1. Projected Offset Between the Spatial Centroid
and the Lens Galaxy
If a group around a lens has a common dark matter
halo, the effects of that halo on the lens potential are
sensitive to any projected spatial offset between the halo
centroid and the lens galaxy. In particular, the offset de-
termines the relative importance of convergence (gravi-
tational focusing) and shear (tidal distortions) from the
group halo.
(See §4.4 and the Appendix for details.)
While we obviously want to determine the offset in each
lens/group system, we also seek to understand the distri-
bution of offsets because that affects the distribution of
convergence and shear, which in turn affects statistical
applications like constraining the dark energy or under-
standing the quad/double ratio. For MG0751, PG1115,
and B1422, the new members we have found allow us to
measure the offsets more precisely. For the newly discov-
ered groups around BRI0952, MG1654, and B2114, we
are able determine the offsets for the first time.
Figure 4 shows that in some cases there is a clear offset
between the lens galaxy and the projected group spatial
centroid.
In B1422, the offset is substantial; the lens
galaxy lies outside the 2σ errors for the group centroid
and is not the galaxy closest to the group centroid. In
MG0751, PG1115 and MG1654 the lens galaxy is only
marginally within the 2σ centroid errorbars, and is also
not the galaxy closest to the centroid. We use these
spatial offsets in our calculations of the shear and the
convergence due to the lens environment in §4.4.
Lens Galaxy Environments
11
Fig. 4. -- Spatial distribution of the group member galaxies on the sky. North is up and east is to the left. The fields are centered on the
lens. Each panel has an angular size of 6.8′ × 6.8′, which is roughly equivalent to the typical size of nearby poor groups. The open circles
mark the group galaxies, an open square denotes the lens galaxy, and a four-pointed star indicates the group centroid and its 2σ errorbars
(see §4.3). All galaxies with measured velocities are marked with small solid points. The scale bar in the lower left corner of each panel
corresponds to 200 kpc at the lens redshift.
4.3.2. Brightest Group Galaxy vs. Group Potential
Key issues that bear on the evolution of groups and
their galaxies are whether the brightest group galaxy
(hereafter BGG) is kinematically and spatially distinct
from (1) the other group members or (2) from the cen-
ter of the group potential.
In nearby X-ray luminous
groups (σr & 300 km s−1), there is a bright, giant ellip-
tical galaxy that occupies the center of the potential (as
defined by the spatial and kinematic centroids) and that
lies on an orbit distinct from the other group members
(ZM98). Our survey now allows us to ask the same ques-
tions for intermediate-redshift groups, and to consider
what the answers imply about group evolution.
In each of our six groups the BGG appears to have an
early-type morphology. The identifications are as follows:
• MG0751: The BGG is the G1 galaxy, which lies
6′′ from the lens galaxy and is 2.2 mag brighter
in I. Leh´ar et al. (1997) fit elliptical profiles to
G1 and find an acceptable fit.
Furthermore,
Tonry & Kochanek (1999) find that G1 is much
brighter that the lens galaxy and has a pure ab-
sorption line spectrum.
• BRI0952: We identify the BGG by visual inspec-
tion of our photometry as an elliptical 2.3′ away
from the lens galaxy and 1.5 mag brighter in I.
• PG1115: The BGG is the giant early-type galaxy
labeled G1, located 12′′ away from the lens galaxy
and 0.7 mag brighter in I. Impey et al. (1998) com-
ment on its early type morphology.
• B1422: Kundi´c et al. (1997b) claim the BGG is the
galaxy G3 located 8′′ from the lens galaxy and 1.5
mag brighter in V. It is the brightest group galaxy
in our sample, has a central location, and exhibits
an early type spectrum.
• MG1654 and B2114: In both cases, the lens galax-
ies are the brightest galaxies in the groups and are
also ellipticals as classified by Kochanek (1995) and
Augusto et al. (2001), respectively.
The top panel of Figure 5 shows the projected spatial
x and kinematic y offsets from the group centroid for the
BGGs (filled squares) and for all other member galax-
ies (filled circles) in our sample groups. The y errorbars
(ǫy) represent the 1σ uncertainties based on adding the
galaxy velocity errors and the group mean velocity er-
rors in quadrature. The x errorbars (ǫx) represent the
1σ uncertainties based on the adding the centroid errors
and the individual galaxy position errors in quadrature
(the centroid errors dominate). To estimate the statisti-
cal errors of the centroid for a group of Ngrp members,
we carry out a statistical bootstrap analysis where we
draw 500 random samples of Ngrp galaxies without re-
placement from the B1422 group (the one with the most
members), and adopt the variance of the centroid posi-
tion as its error. For B1422, we use the smallest of the
12
Momcheva et al.
errors calculated for the other groups.
We define the quantity R2 = (x/δx)2 + (y/δy)2 as a
measure of the phase-space distance of a galaxy from the
group centroid. Here δx and δy are the rms deviations in
x and y for all galaxies plotted in the top panel of Figure
5. A galaxy will have a large value of R if it has a large
peculiar velocity and/or a position that is far from the
projected spatial centroid. Conversely, galaxies at rest
in the center of the group potential will have small R
values. The bottom panel of Figure 5 shows the distri-
butions of R values for the BGGs (shaded histogram) and
for all other group members (unshaded histogram). We
can now use these distributions to answer two questions
about the BGGs.
Are the BGGs distributed differently than the other
group galaxies? We compute the R distributions for
BGGs and for all other group galaxies (see the bottom
panel of Figure 5), and then compare them using three
statistical tests: the KS-test (to compare the overall dis-
tribution), the t-test (to compare the means), and the
F-test (to compare the variances). All three fail to dis-
tinguish between the two distributions. In other words,
there are no significant differences between the orbits of
BGGs and the orbits of other group members, at least
for these small samples. This result differs from observa-
tions of nearby groups (see ZM98). It is not clear from
our present sample whether our result indicates real evo-
lution in the group galaxy population between z ∼ 0.3
and z = 0, or is due simply to small number statistics.
It will be interesting to return to this question with the
larger sample of lensing-selected groups that we are ob-
taining.
Are the BGGs consistent with the group centroids? We
compare the distribution of R values for the BGGs with
a model distribution expected for a galaxy lying at the
bottom of the group potential. To incorporate measure-
ment errors, we treat the model distribution as a Gaus-
sian in x and y with rms deviations of ǫx/δx and ǫy/δy,
respectively, and make 1000 random draws using the ap-
propriate values of ǫx and ǫy for each group. The bottom
panel of Figure 5 shows the observed and model distribu-
tions. A KS-test gives 6.6 × 10−3 as the probability that
the two samples are drawn from the same distribution.
The probability that the means of the two distributions
are the same is 7 × 10−9, while the probability that the
variances are the same is 10−5. We conclude that the
BGGs do not occupy the center of the group potential.
This offset of the BGG from the kinematic and spatial
centroid of the group is not seen in nearby X-ray lumi-
nous groups. While our result could suggest group evo-
lution from z ∼ 0.3 to now, another possibility is that we
are not comparing apples to apples. For example, there
are groups in our sample with velocity dispersions lower
than what is typical for dynamically-evolved X-ray lu-
minous groups nearby. Among nearby groups, lower-σr
systems tend to be dynamically younger and are more
likely to have an offset BGG (e.g., the Local Group).
This may be true among our z ∼ 0.3 groups as well.
Support for this latter interpretation comes from a
closer look at the BGGs in the three high-σr groups in
our sample: B1422, PG1115, and MG0751. The BGGs
of MG0751 and PG1115 are the galaxies closest to the
projected spatial/kinematic centroid in their respective
groups, and, within the errors, are consistent with be-
Fig. 5. -- (Top) Projected spatial and kinematic offsets of the
brightest group galaxy (BGG; filled squares) and all other group
members (filled circles) from the group centroid for five of our
groups. The lens galaxies are marked with open circles.
(The
lens galaxy in BRI0952 is omitted because it does not have a mea-
sured redshift.) The velocity offset is normalized by the velocity
dispersion of the group to compensate for the differences among
the group potentials. The y errorbars represent the 68% confi-
dence level based on adding the errors for the lens galaxy velocity
and the mean group velocity in quadrature. The x error bars are
the 68% confidence level based on a statistical bootstrap test (see
text). (Bottom) Distribution of the phase-space offset R for the
BGGs (shaded histogram) and all other group galaxies (open his-
togram, normalized by the number of the brightest group galaxies).
Statistical tests fail to distinguish between these two distributions.
The heavy line (scaled down by a factor of 400 to fit the y-axis)
shows the model R distribution for a galaxy assumed to lie at the
bottom of the group potential (see text). A KS-test gives 7 × 10−3
as the probability that the BGG and model distributions are drawn
from the same parent distribution, and a t-test gives 7 × 10−9 as
the probability that the means of the BGG and model distributions
are the same. We conclude that the BGGs generally do not occupy
the center of the potential, and, that, within the large uncertain-
ties, their phase space distribution is consistent with that of the
other group members.
ing at the centroid. In contrast, the BGG in B1422 lies
at a large projected distance from the projected spatial
centroid (∼ 0.22 Mpc) and has a substantial peculiar ve-
locity (1270 km s−1). The BGG spatial and kinematic
offset, together with the clumpiness of B1422 (see §4.1),
suggest that this system is not yet relaxed. Therefore, in
two out of three cases, the high-σr groups in our sample
are comparable to those nearby groups with high velocity
dispersions and centrally located BGGs.
We also compare the projected spatial centroid with
the peak of emission from the diffuse, luminous X-ray
halos in PG1115 and B1422 (the two highest-σr groups)
seen by Grant et al. (2004).
In PG1115, the projected
spatial centroid and X-ray peak are consistent within 2σ.
By contrast, in B1422 the peak of the X-ray emission is
substantially offset (more than 3σ) from both the un-
weighted and luminosity-weighted group centroids.
Our ability to address the questions of evolution raised
in this section will improve when we finish obtaining
the larger sample of lensing-selected groups. Since lens-
ing is sensitive to groups spanning the redshift range
0.2 . z . 1, it will naturally provide the large red-
Lens Galaxy Environments
13
shift baseline needed to probe evolution. Working with a
self-consistently selected sample of groups will mitigate
selection effects.
4.4. Group Contributions to Lens Potentials
The understanding of environment-related biases and
uncertainties in lensing constraints on the masses and
shapes of galaxy dark matter halos, H0, and substruc-
ture requires detailed lens modeling (see KZ04). That is
beyond the scope of this paper and will be treated sepa-
rately. For our purposes here, a simple way to quantify
environmental effects is to determine the dimensionless
convergence κ and shear γ that the group contributes to
the lens potential. The convergence represents gravita-
tional focusing created by additional mass at the posi-
tion of the lens galaxy, while the shear represents tidal
distortions created by having an inhomogeneous distri-
bution of matter near the lens galaxy. Convergence can
never be constrained using lens models alone because of
the mass sheet degeneracy (Gorenstein et al. 1988; Saha
2000). As a result, it is often omitted, which can lead to
significant biases in the model results (KZ04). Shear can-
not be constrained in models of two-image lenses, which
leads to enormous model uncertainties. While shear can
be constrained in models of four-image lenses, one of the
puzzling results is that shear is required in nearly all four-
image lenses, and the required shear strengths cannot
easily be explained by traditional models of large-scale
structure (Keeton et al. 1997; Dalal & Watson 2004).
Models of four-image lenses lead to the rule of thumb
that a shear of γ ∼ 0.1 is common for groups, and γ ∼
0.3 for clusters. Generally, we expect κ ≥ γ because of
the way convergence and shear add when there is more
than one perturber (see the Appendix). Many lensing
conclusions scale as (1 − κ) to some power (see KZ04 and
the Appendix), which can help us estimate the biases.
For example, if a lens has convergence κ ∼ 0.1, then lens
models that omit the convergence will overestimate H0
by ∼10%. Hence, we consider convergences and shears
larger than ∼0.1 to be quite important, and values down
to ∼0.05 worth consideration.
In this section, we quantify the effect of the group envi-
ronments in our sample by determining the convergence
and shear due to the group surrounding each lens galaxy.
Because it is not clear how the mass is divided between
the individual group members and a common group dark
matter halo, we consider two extreme cases that bound
the range of possibilities. In the "group halo limit," we
assume that all of the mass is associated with a common
group halo, and we estimate the amount of mass from
the velocity dispersion. In the "group galaxies limit," we
assume that all of the mass is bound to the individual
member galaxies, and we calibrate the individual shears
and convergences with recent weak lensing observations
(Sheldon et al. 2005). Comparing results from the two
limits can indicate how much our conclusions depend on
how the mass is distributed within the group.
4.4.1. Group Halo Limit
The group halo limit is appropriate for considering re-
laxed, dynamically evolved groups in which the individ-
ual dark matter halos of member galaxies may have been
stripped by interactions, so that the dominant compo-
In this case, the
nent is a common dark matter halo.
luminous galaxies just trace the underlying mass distri-
bution, and the velocity dispersion of the galaxies is a
measure of the mass of the group as a whole. If we model
the group halo as a singular isothermal sphere (SIS), then
we can take the velocity dispersion σr of a group and its
centroid position (b, φ) in polar coordinates centered on
the lens galaxy, and compute the convergence κgrp and
shear γgrp using eq. (A20) in the Appendix. (Note that
κgrp = γgrp in the SIS approximation.) We then use
the uncertainties in the velocity dispersion and centroid
position to determine the uncertainties in the conver-
gence and shear. These derived uncertainties are highly
non-Gaussian, so we compute them using Monte Carlo
simulations.
The results are presented in Table 5.4 In three of the six
groups (MG0751, PG1115, and B1422), the convergence
and shear are significant.
In these systems the group
environment needs to be accounted for in lens models,
and our measurements of the group properties make that
feasible. In the other three groups (BRI0952, MG1654,
and B2114), it appears from our current data that the
convergence and shear are small.
The case of B1422 illustrates how our identifications of
additional group members affect estimates of the conver-
gence and shear. Based on six members, Kundi´c et al.
(1997b) estimated that the group velocity dispersion is
σr = 550 km s−1, and that the centroid lies a projected
distance of b ∼ 14′′ from the lens galaxy. That led them
to estimate a large "observed" shear of γobs = 0.23, a
value consistent with the shear γmod ∼ 0.2 required by
lens models (Hogg & Blandford 1994; Keeton et al. 1997;
Dobler & Keeton 2005). Now with 16 members, we find a
smaller velocity dispersion σr = 470 km s−1 and a larger
offset b = 43′′, which reduce the nominal observed shear
to a more moderate value of γobs = 0.058. Accounting for
the measurement uncertainties (which has not been done
before), we find allowed ranges of 0.034 ≤ γobs ≤ 0.095
at 1σ, and 0.020 ≤ γobs ≤ 0.170 at 2σ. It is clear that
omitting the group members can bias the estimates of
how the environment affects the lens potential. However,
even with an extensive catalog (16 members), uncertain-
ties are important when comparing observations of the
group with inferences from lens models. The large dif-
ference between the nominal observed shear γobs = 0.058
and the shear required by models γmod ∼ 0.2 might not
be significant at more than 2σ. To test consistency be-
tween the observations and models, it is best to make
new models with the environment constrained by our
observations (including the uncertainties, and possible
clumpiness within the group; see §4.1).
4.4.2. Group Galaxies Limit
In the group galaxies limit, we suppose that all the
mass in the group is bound to the individual member
galaxies. This is probably a better approximation for
dynamically younger groups, which may still be in the
process of collapse and whose galaxies still retain their
halos.
To calibrate the shear and convergence from each
galaxy, we turn to observations that are perfectly suited
4 As noted in the table caption, we assume zs = 2 for the un-
known source redshift in B2114, but the particular value has little
effect on our results.
14
Momcheva et al.
Convergence and Shear in the Group Halo Limit
TABLE 5
Lens
PG1115
B1422
MG0751
MG1654
BRI0952
B2114a
b
[′′]
23
43
23
65
61
49
σr
[km s−1]
κgrp = γgrp
θγ
[deg]
Number of
Images
440
470
320
200
170
110
0.089 +0.065
6 ± 37
−0.046
0.058 +0.037
−0.024 −65 ± 20
0.049 +0.071
−0.033 −27 ± 36
0.007 +0.011
−0.005 −35 ± 30
0.005 +0.013
−0.004 −81 ± 42
0.003 +0.014
−0.002 −12 ± 52
4
4
4+R
R
2
2
Note. -- The groups are sorted by decreasing values of the con-
vergence and shear. The magnitudes of the convergence and shear
are equal (κgrp = γgrp) in the SIS approximation. Values larger
than 0.05 are marked in boldface. The angle θγ defines the direction
of the shear (measured North through East). The convergence and
shear errorbars are 1σ uncertainties derived from the uncertaintines
in the group centroid position and velocity dispersion (from Table
4). Column 6 lists the image configuration for each lens: 2-image,
4-image, or ring (R).
aWe assume a source redshift zs = 2 for B2114, but the particular
value has little effect on our results. In particular, assuming zs = 3
leads to the same numerical values for the κgrp, γgrp, and θγ.
to our needs: weak lensing. The advent of large surveys,
such as the Sloan Digital Sky Survey, has made it possible
to measure shear as a function of distance from an aver-
age galaxy, on scales from 20 h−1 kpc to 7 h−1 Mpc (e.g.,
Sheldon et al. 2005).5 The projected offsets of galaxies
in our sample range over ∼ 20 -- 700 h−1 kpc, so we are
working in precisely the regime studied by Sheldon et al.
(2005). One possible concern with this calibration is
that the Sheldon analysis provides limited information
about how shear and convergence scale with luminosity
(or mass). Sheldon et al. (2005) measured the shear pro-
file in three luminosity bins, but most of our galaxies fall
into just one of the bins (−22 < Mi − 5 log h < −17).
Thus, we expect that our present analysis characterizes
the average properties of the group galaxies well, but it
will be worthwhile to redo the analysis when future weak
lensing data provide finer luminosity resolution.
The quantity that is measured in weak lensing studies
is shear, but the observed shear profile is consistent with
a power law which means that there is a simple relation
between the shear and convergence. The Appendix gives
this relation, and also provides more details about how
we use the weak lensing data to calibrate our analysis.
Once we have computed the convergence and shear
from each galaxy, we combine them in the manner of
eqs. (A12) -- (A16) in the Appendix. The net convergence
and shear for each group are presented in Table 6. We
also list statistical uncertainties derived (using Monte
Carlo simulations) from the errorbars on the weak lens-
ing parameters quoted by Sheldon et al. (2005). These
are generally small and are probably less important than
systematic effects due to incompleteness (see below). As
in the group halo limit, in the galaxies limit we find that
MG0751, PG1115, and B1422 all have significant group
contributions to the lens potential. The group contribu-
5 Sheldon et al. (2005) quote comoving distances, but we have
converted to angular diameter distances as those are more natural
for our analysis.
tions are fairly small in MG1654, BRI0952, and B2114.
Comparing Tables 5 and 6 leads to a crucial point: our
conclusions about which groups significantly affect the
lens potentials do not depend on assumptions about how
the mass is distributed within the groups. While the
detailed lensing implications do depend on the mass dis-
tribution -- which is actually useful (see §4.4.3) -- it is
reassuring to see that our qualitative conclusions about
which groups are important for lensing are robust.
In carrying out this analysis we do not distinguish be-
tween different morphological galaxy types. Given their
higher mass-to-light ratios, elliptical (red) galaxies would
produce more shear and convergence than spiral (blue)
galaxies of the same luminosity. To test the effects of
morphology on our results we consider the extreme as-
sumption that all group galaxies are red and use the
Sheldon et al. (2005) shear profiles for red galaxies. In al-
most all cases the total convergence and shear due to the
group increase by ∼70%. Thus, if most group members
are red galaxies then our current analysis may actually
underestimate the convergence and shear by as much as
70%.
As a simple sanity check, in Table 6 we also compute
the net convergence and shear assuming that each galaxy
can be treated as a singular isothermal sphere (SIS) with
velocity dispersion σ = 100 km s−1. We expect that
many of the galaxies have larger velocity dispersions, and
given that convergence and shear scale as σ2, this sim-
ple case should provide a conservative lower bound on
the net convergence from the group. (The shear is more
complicated, as discussed below.) Comparing the results
assuming σ = 100 km s−1 with those based on the weak
lensing calibration confirms this expectation, and gen-
erally suggests that the weak lensing calibration is rea-
sonable. We should note that the convergence from the
simple SIS model may not be a strict lower bound if ha-
los are significantly truncated, but in practice most of the
convergence (and shear) arise from galaxies close enough
Lens Galaxy Environments
15
Net Convergence and Shear in the Group Galaxies Limit
TABLE 6
Lens
κtot
[Sheldon]
γtot
θγ
[deg]
PG1115
B1422
MG0751
MG1654
BRI0952
B2114a
0.066 +0.012
−0.006
0.092 +0.019
−0.009
0.115 +0.040
−0.026
0.027 +0.006
−0.004
0.013 +0.002
−0.002
0.016 +0.004
−0.002
0.028 +0.009
64.2 ± 3.2
−0.003
0.010 +0.012
−0.004 −35.6+25.4
−13.5
0.091 +0.057
−0.034 −82.9 ± 1.8
0.006 +0.005
21.4 ± 13.2
−0.003
0.007 +0.001
−0.001 −64.6 ± 8.3
0.012 +0.003
−0.002 −15.9 ± 3.3
[σ = 100 km s−1]
κtot
θγ
γtot
[deg]
0.040
0.064
0.040
0.013
0.006
0.008
0.031
66
0.019 −34
0.029 −88
0.004
43
0.003 −67
0.007 −16
Note. -- The groups are sorted as in Table 5. Values larger than 0.05 are
again marked in boldface. Columns 2 -- 4 list the net convergence and shear
when we calibrate the individual galaxies using the weak lensing observations
by Sheldon et al. (2005). The errorbars represent 1σ statistical uncertainties
derived from the weak lensing errorbars quoted by Sheldon et al.; even more
important may be systematic effects due to spectroscopic incompleteness (see
text). For a simple comparison, Columns 5 -- 7 list the results when we treat each
galaxy as an isothermal sphere with velocity dispersion σ = 100 km s−1.
aAgain, we assume a source redshift zs = 2 for B2114, but the particular value
has little effect on our results.
to the lens that truncation would have little effect. For
example, if all halos had cut-off radii at 300 kpc, the SIS
model convergence would drop from κ = 0.040 to 0.036 in
PG1115, and from 0.064 to 0.057 in B1422. Even with
an extreme cut-off at 100 kpc, the convergences would
still be 0.027 for PG1115 and 0.051 for B1422. (In all
cases, the shears are basically unchanged.) We expect
that the conservative assumption of a small velocity dis-
persion more than compensates for the omission of a cut-
off radius, so that the SIS model results are indeed lower
bounds.
It is again interesting to consider how our efforts to in-
crease the group membership have affected conclusions
about the convergence and shear. The most instruc-
tive case is PG1115. With our catalog of 13 members,
we find a net convergence κ = 0.066+0.012
−0.006 and a net
shear γ = 0.028+0.009
−0.003 at position angle θγ = 64 ± 3. If
Kundi´c et al. (1997a) had done the same analysis with
their catalog of four members, they would have found
κ = 0.027+0.008
−0.003, and θγ = 69 ± 5.
(Tonry 1998 would have obtained similar results.)
In
other words, the previous catalogs missed at least half of
the sources of shear and convergence. The problem was
that they focused on the region within ∼ 30′′ of the lens,
but (in the galaxies approach) the galaxies have extended
dark matter halos, requiring the inclusion of group mem-
bers out to the full virial radius of the group in order to
fully characterize environmental effects in lens models.
−0.005, γ = 0.014+0.006
With this thought in mind, we must consider how spec-
troscopic incompleteness may affect the results in Table
6. We cannot account for galaxies that we missed, but
we can ask how our results would have differed had we
omitted a few of the galaxies that we did actually in-
clude. The effects of incompleteness on the net conver-
gence are simple: since convergences from different galax-
ies combine in a simple scalar sum (see eq. A12 in the
Appendix), omitting galaxies causes us to underestimate
the net convergence. Turning this around, we can say
that our estimates of κtot are strict lower bounds on the
convergence from the group. To be more quantitative,
in analyzing subsamples of our group catalogs we find
that the net convergence scales roughly with the num-
ber of group members; so if the true number of members
is, say, 50% larger than what we have observed then we
expect the true convergence to be ∼50% larger than our
estimate. For understanding biases in lens models, it is
very valuable to have a lower bound on the total conver-
gence, because that can be turned into lower bounds on
the biases.
The shear is more complicated, because multiple
contributions sum as tensors rather than scalars (see
eqs. A13 and A14); adding more contributions can ei-
ther increase or decrease the net shear, and modify the
position angle. The effects depend on the spatial dis-
tribution of member galaxies. In PG1115 the most im-
portant galaxies lie roughly in a line on the sky, which
means that the direction of the net shear is robust against
incompleteness, while the amplitude of the shear scales
roughly with the number of group members. In contrast,
in B1422 the galaxies are distributed more broadly, so in-
completeness may change the shear direction by tens of
degrees and the shear amplitude by tens of percent. (Of
course, the shear from the galaxies in B1422 is small, so
even large fractional uncertainties are not so important.)
Finally, in MG0751 the environmental effects are domi-
nated by the galaxy G1 that is massive and close to the
lens.
The bottom line is that incompleteness does not sig-
nificantly affect our qualitative conclusions. Since our
estimate of the net convergence is a lower bound, we
know that our conclusion that groups are important on
MG0751, PG1115, and B1422 is robust. While it is pos-
sible that our conclusion that the groups are not so im-
portant in MG1654, BRI0952, and B2114 could change
if we measure more galaxies and find that κ rises, the
observed correspondence between th group velocity dis-
persion and richness (see §4.2) suggests that adding more
members would not affect κ significantly. Incompleteness
16
Momcheva et al.
issues will need to be considered when making detailed
quantitative comparisons between our environment ob-
servations and lens models.
4.4.3. Discussion
It is important to understand the similarities and dif-
ferences between the results from the "group halo" and
"galaxies" limits. We have already noted that both limits
lead to the same conclusions about which groups are im-
portant for lensing. All three high-σr groups (MG0751,
PG1115, and B1422) produce a significant convergence
in both approaches, so these groups cannot be ignored in
lens models. At the same time, two of the low-σr groups
(BRI0952 and B2114) produce a small shear and conver-
gence in both approaches, suggesting that these groups
are not so important for lensing. While this latter conclu-
sion may not seem exciting, it is actually quite valuable.
Two-image lenses (including both BRI0952 and B2114)
suffer from a strong degeneracy between ellipticity and
shear, if both quantities are unknown. That degeneracy
can now be broken by ruling out models with large shear.
The situation is less clear for MG1654, because the group
halo analysis implies negligible shear, while the galaxies
analysis yields a small but non-negligible shear γ = 0.03
(and that could be an underestimate).
When we turn to a more quantitative comparison of the
two approaches, we notice some significant differences.
The differences suggest that it may be possible to dis-
tinguish between the group halo and galaxies limits, and
thus to learn about the distribution of dark matter within
the groups. For example, in B1422 the group halo analy-
sis leads to a moderate convergence and shear, while the
galaxies analysis leads to a larger convergence but a neg-
ligible shear. Lens models require a large shear γ ∼ 0.2
that is marginally consistent with the group halo ap-
proach (given our uncertainties) but grossly inconsistent
with the galaxies approach. In PG1115, as in B1422, lens
models require large external shear γ ∼ 0.1 (Impey et al.
1998). The group halo limit result (γ = 0.089+0.065
−0.046) is
within 1σ of the model requirement while the galaxies
limit gives a factor of three lower shear, inconsistent with
predictions. The results in both B1422 and PG1115 --
the two highest velocity dispersion groups -- suggest that
the mass is distributed in a common halo rather than be-
ing attached to the individual group members (as might
be expected if high velocity dispersion groups are more
dynamically evolved). This hypothesis needs to be ex-
amined more carefully with detailed lens models; rather
than just comparing the shear required by lens models
with that inferred from our observations, it is important
to build models that explicitly incorporate the environ-
ment (which may even consist of multiple subgroup halos;
see §4.1) in both the group halo and galaxies limits and
see whether either case can fit the lens data. Systems like
B1422 and PG1115 may provide an exciting opportunity
to determine the distribution of dark matter in a distant
group.
We also note that KZ04 used PG1115 as a fiducial
example with which to asses environment-related biases
in lens models. Dalal & Watson (2004) suggested that
PG1115 is a very atypical lens environment, and that
KZ04 therefore overestimated environmental effects. To
the contrary, we find that PG1115 is nothing if not typi-
cal: the group's kinematic properties and shear and con-
vergence are consistent with at least half of the groups
in our sample.
Another interesting system is MG0751. Here, the
galaxies analysis is dominated by the G1 galaxy,
ly-
ing just 6′′ from the lens. Even so, Leh´ar et al. (1997)
showed that lens models including only the lens galaxy,
G1, and up to three other nearby galaxies cannot fit the
lens data.
It will be interesting to use new lens mod-
els to test the hypothesis that both G1 and the common
group halo contain significant mass, and to see whether
we can constrain their relative masses. We must issue
two warnings, however. First, the projected offset of G1
from the lens galaxy is just 20 h−1 kpc, which lies at the
inner limit of the range studied by Sheldon et al. (2005);
thus, the reliability of the weak lensing calibration is not
clear. Second, we show in §4.5.2 that MG0751 also has
a significant shear from a group along the line of sight,
which must be included in lens models along with the
group at the lens redshift.
In summary, our results above show that: (1) Group
environments, whether the mass lies with individual
member halos (galaxies limit) or in a common group halo
(group halo limit), can contribute significantly (κ, γ ≥
0.05) to the lens potential. (2) If the members have ha-
los (galaxies limit), they can have a big effect, perhaps
even greater than that of a common group halo. (3) In
the galaxies limit, correcting for incompleteness is only
going to boost the convergence (but will move the shears
in either direction). (4) In the highest velocity dispersion
groups, the shears produced in the group halo limit are
more consistent with the observationally required values,
suggesting that we might be able to discriminate between
the models of the mass distribution.
Finally, it is remarkable that the three lenses with sig-
nificant environmental effects include both quad lenses
(PG1115 and B1422) and the one quad/ring (MG0751),
whereas the double/ring (MG1654) and the two double
lenses (BRI0952 and B2114) all have small convergence
and shear. While this result is limited by small number
statistics, it may suggest a correlation between image
configuration and environment. Conventional wisdom
(e.g., Rusin et al. 2001) holds that shear does not sig-
nificantly affect the relative numbers of quads and dou-
bles. However, KZ04 argue that treating environment
properly (including terms beyond a simple shear) does
change the quad/double ratio. Our new results provide
empirical evidence that there is a connection.
4.5. Lensing Effects of Line-of-Sight Structures
Since lensing is a projected phenomenon, we must
consider whether structures projected along the line
of sight significantly affect strong lens systems. Dif-
ferent theoretical approaches to studying the effects
of interlopers on strong lensing in a ΛCDM universe
have yielded contradictory results
(e.g., Bar-Kana
1996; Keeton et al. 1997; Premadi & Martel 2004;
so an empir-
Wambsganss, Bode & Ostriker 2004),
ical approach is clearly necessary.
To date, there
are only three lenses with confirmed line-of-sight
groups (B0712+472, B1608+656 and MG 1131+0456;
Fassnacht & Lubin
2000;
[Fassnacht et al. 2005) and several other candidates
(Faure et al. 2004; Morgan et al. 2005). Here we present
the first systematic survey for structures along the
2002; Tonry & Kochanek
Lens Galaxy Environments
17
line-of-sight to strong lens systems. The large redshift
baseline and wide field of view of our spectroscopic
survey make it ideally suited to address this issue.
4.5.1. The Zone of Influence
Our first task is to estimate the "zone of influence" for
a perturber along the line of sight to each lens in our
sample. Although it has not been done before, the cal-
culation is straightforward using the formalism presented
in the Appendix. Briefly, if we assume that a perturber
can be modeled as an isothermal sphere with some given
velocity dispersion, then we can use eqs. (A6), (A7), and
(A20) to compute the effective convergence and shear
(κeff and γeff ) as a function of the impact parameter b of
the perturber relative to the lens, and the redshifts of the
lens galaxy and perturber. (In the SIS approximation,
κeff = γeff for a single perturber.)
Figure 6 shows contours of κeff = γeff in the plane of b
and ∆z = zpert − zlens, for perturbers with velocity dis-
persions of 100, 300, or 500 km s−1. (Because we do not
see rich clusters along the lines of sight to these lenses,
this range of σr should span the observed range of struc-
tures.) The grayscale is explained in the figure caption.
We consider the zone of influence to be the region in
which κeff, γeff ≥ 0.05, i.e., the unshaded (white) region
in the figure. The shape of this region is not sensitive to
the fact that an SIS halo has an infinite extent; a cut-off
halo radius of 300 kpc does not change the zone of influ-
ence. Also interesting are regions in which κeff and γeff go
negative (shaded black), which represents a breakdown
of our formalism. This happens only when the intrinsic
convergence of the perturber is κ > 0.5, which means
that the offset between the lens and perturber is small
enough that the lens actually lies within the Einstein ra-
dius of the perturber (see the Appendix). In this case,
the "perturber" is no longer just a perturbation because
its caustics interact with those of the main lens galaxy,
and we would observe a strong lensing effect from the
second mass as well. This breakdown does not affect our
conclusions because we do not actually see any structures
lying within this region; besides, any objects that lie so
close to the line of sight to the lens would presumably be
known from previous observations.
As the velocity dispersion of the perturber increases,
the zone of its influence grows dramatically ∝ σ2
r . Con-
sequently, a more massive perturber can produce a large
shear and convergence even when offset from the lens; a
perturber with σr ∼ 500 km s−1 (i.e., a rich group) can
be offset by as much as 1′ and still produce γ ∼ 0.05.
Another striking feature of Figure 6 is the very wide red-
shift baseline in front of and behind the lens over which a
perturber can cause significant effects. This result shows
that it is crucial to catalog not just mass structures in
the immediate vicinity of the lens galaxy, but also else-
where along the line of sight, in order to model the lenses
accurately.
4.5.2. Prominent Interloping Structures
To quantify line-of-sight effects for the lenses in our
sample, we consider the most prominent structures
(those likely to be groups and clusters) identified from
our redshift catalog. This approach is analogous to the
group halo limit for groups around lens galaxies (§4.4.1),
For figure see file f6 sm.jpg
Fig. 6. -- Contours of κeff = γeff produced by a perturber
with a velocity dispersion of 100, 300 and 500 km s−1, lo-
cated at a redshift 0 < zpert < 1 and having an impact pa-
rameter b with respect to the lens, computed for seven of
the lenses in our sample.
(We exclude PMN2004 because
its lens redshift is unknown.) The horizontal axis represents
the redshift difference between the perturber and lens galaxy.
The vertical axis represents the projected distance of the per-
turber from the lens, out to 2′. Contours are drawn at κeff =
γeff = 0.001, 0.01, 0.05, 0.1 and 0.5, although not all of them
are clearly visible in all panels. The grayscale is as follows:
κeff , γeff ≥ 0.05 (white); 0.05 > κeff , γeff ≥ 0.01 (light gray);
0.01 > κeff , γeff ≥ 0.001 (medium gray) and κeff , γeff < 0.001
(dark gray).
Important regions are 0.5 > κeff , γeff > 0.05,
i.e., the areas in white. At small impact parameter, the per-
turbation approximation breaks down as we enter a regime in
which the perturber itself causes strong lensing (see text); in
our calculations, this leads to negative values of κeff and γeff
(shaded in black). Notice the strong dependence on σ: the
zone of influence scales as σ2. Massive perturbers can have
large effects even when they lie far from the lens. Another
striking feature is the very wide redshift baseline in front and
behind the lens over which the perturber can cause a signifi-
cant effect.
because it accounts for the dark matter in massive bound
structures. To identify potentially important line-of-
sight structures, we show the redshift histograms again in
Figure 7, now shading only those galaxies that lie within
1′ of the lens. We have over-plotted the curve of the nor-
malized shear strength κeff/κ = γeff/γ (eq. A11) to give
an indication of how the convergence and shear vary with
redshift. From the discussion in §4.5.1, we expect struc-
tures with large velocity dispersions, small projected off-
sets from the lens, and/or small redshift offsets from the
lens galaxy to contribute most to the lens potential. To
be conservative, we select only those peaks in Figure 7
that: (1) have at least four members; (2) lie within ∆z
such that κeff/κ and γeff /γ are & 0.5; and (3) have at
least one member projected within 1′ of the lens. For
every peak, we set pessimistic 3σ velocity limits and use
bi-weight estimators of location and scale to calculate the
mean velocity υi and the line-of-sight velocity dispersion
σr,i. Based on the membership, we then calculate the
projected spatial centroid of the structure, and its offset
from the lens galaxy bi and position angle φi. Finally,
we can use eqs. (A6), (A7), and (A20) to determine the
effective shear and convergence.
The results are presented in Table 7. We list the in-
dividual effects of all prominent structures. We choose
to include the groups at the lens redshifts (repeating re-
sults from Table 5), so that we can compare local versus
interloping structures. We then sum all structures (us-
ing eqs. A12 -- A16) to obtain the total convergence and
shear for each lens (in the group halo limit). We empha-
size that our results are conservative in the sense that we
have not tallied all the line-of sight objects that might
affect the lens models. We do not include galaxies out-
side prominent peaks and plan to address this question in
future work. We have not included all peaks in the veloc-
ity histograms (Fig. 7), because most are undersampled
or intrinsically poor, and thus have too few members for
us to interpret them as likely groups and to compute a
18
Momcheva et al.
Fig. 7. -- Same as the left panels in Figure 3, except that here we have shaded only those galaxies within 1′ of the lens. We have
also overplotted the normalized shear strength and normalized convergence (κeff /κ = γeff /γ; see text) to guide the eye regarding general
behavior of convergence and shear as the perturbing structure is moved away from the lens in redshift.
meaningful velocity dispersion.
In addition, the veloc-
ity dispersions we compute are probably underestimates
not only because of the narrow (conservative) choice of
initial velocity ranges but also because many of the veloc-
ity peaks are poorly sampled. Finally, our spectroscopic
target selection prioritizes galaxies thought to lie at the
lens redshift (see §3), and to some extent that limits our
ability to identify interloping structures.
The discussion of incompleteness in §4.4.2 applies here
as well, with one minor change. As before, we note that
convergence is a scalar and that the contribution from an
overdensity is always positive, and conclude that iden-
tifying additional prominent structures would only in-
crease the total convergence κtot. The change is that
an uncertainty in the convergence zeropoint prevents us
from declaring that we have obtained a strict lower bound
on the total convergence from the line of sight. The zero-
point uncertainty arises from the effects of underdensities
such as voids. Strong lensing calculations conventionally
assume that there are a few density peaks superposed on
top of a smooth background at the mean density of the
universe. Voids can then be thought of as regions where
the density is negative (relative to the mean), which con-
tribute negative convergence to the lens potential (e.g.,
Seljak 1994).
If there is a significant negative conver-
gence, it effectively changes the convergence zeropoint:
in summing the effects of overdense structures along the
line of sight, we should start not from zero but from the
appropriate negative convergence.
We cannot measure the convergence zeropoint directly
because that requires knowledge of the local density of
matter (dark and luminous) at every point along the line
of sight, which is impossible to obtain even with complete
redshift surveys. Nevertheless, we can make a useful es-
timate based on a simple model. We create Monte Carlo
simulations of random lines of sight in a universe in which
some of the mass is contained in halos while the rest is in
a smooth background at a level below the mean density.
In these simulations we are able to determine the total
convergence κtot from density fluctuations along the line
of sight, as well as the contribution κpeaks from promi-
nent structures (κpeaks is analogous to what we compute
above from our data). We can then interpret the differ-
ence, κzp ≡ κtot − κpeaks, as the convergence zeropoint.
(Details of the calculation, and further discussion of the
results, are given by Keeton et al. 2005b.)
Lens Galaxy Environments
19
Fig. 7. -- continued.
Considering many random lines of sight, we find that
κzp has a mean of zero and a dispersion of . 0.02, for typ-
ical lens and source redshifts and reasonable halo mass
functions. Because the zeropoint uncertainty is small,
we believe that our κtot values above are probably lower
limits on the total convergence. In other words, it is un-
likely that voids contribute enough negative convergence
to counter the positive κ from prominent peaks in a given
line of sight.
The main result is that four of the eight lenses
in our sample have significant interloping structures.
MG0751, PG1115, and B1422 each have one structure,
and HE2149 has three. At least one of those structures,
along the line of sight to MG0751, has a significant con-
tribution to the lens potential. This perturbing group lies
at a redshift of z = 0.56, which places it between the lens
galaxy and the source quasar, and has a velocity disper-
sion σr = 550 km s−1 based on the six members that we
have identified. With a centroid that lies at a small pro-
jected offset of 28′′ from the lens, the group contributes
a convergence and shear κlos = γlos = 0.066. This is
clearly an important contribution to the lens potential
-- in fact, it is slightly stronger than our estimate of the
contribution from the group at the lens redshift (in the
group halo limit).
It represents one more piece in the
interesting puzzle of fully understanding lensing in the
MG0751 system.
To our knowledge, our lens sample is not biased toward
having significant line-of-sight effects. Our survey meth-
ods are, if anything, somewhat biased against finding
line-of-sight structures (as discussed above). Therefore,
our discovery of a significant structure in 1/8 lenses sug-
gests that line-of-sight effects are important in at least
∼10% of all lenses. That estimate needs to be confirmed
with a larger sample, but it does indicate that lensing
effects from the line of sight deserve further attention.
5. CONCLUSIONS
We have presented the first results from our spectro-
scopic survey of the environments of strong gravitational
lenses. We have used multislit spectroscopy to measure
the redshifts of 355 galaxies in the fields of eight strong
gravitational lenses with lens galaxies at redshifts be-
tween 0.25 and 0.50. After adding 16 redshifts from the
literature, we have analyzed a total sample of 371 galax-
ies with redshifts.
The lens galaxy belongs to a poor group in six of the
eight systems in our sample. We discover three new
groups associated with the lens galaxy of BRI0952 (five
20
Momcheva et al.
TABLE 7
Convergence and Shear Due to Prominent Line-of-Sight Structures
Lens
ID
zpert
b
[′′]
σr
[km s−1]
κeff = γeff
MG0751
MG0751
MG0751
BRI0952
BRI0952
PG1115
PG1115
PG1115
B1422
B1422
B1422
MG1654
MG1654
HE2149
HE2149
HE2149
HE2149
B2114
B2114
1
2
total
1
total
1
2
total
1
2
total
1
total
2
3
4
total
1
total
0.35
0.56
0.42
0.31
0.49
0.34
0.28
0.25
0.27
0.45
0.60
0.31
23
28
61
23
44
43
80
65
51
70
56
49
320
550
170
440
300
470
400
200
400
180
150
110
0.049
0.066
0.005
0.089
0.010
0.058
0.020
0.007
0.017
0.004
0.003
0.003
θγ
[deg]
−27
−56
−82
7
−74
−65
−28
−35
−62
−80
88
−12
κtot
γtot
θγ,tot
[deg]
0.116
0.101
−44
0.005
0.005
−82
0.099
0.080
5
0.078
0.066
−57
0.007
0.007
−35
0.024
0.022
−68
0.003
0.003
−12
Note. -- "Prominent" structures are defined as having at least four members, at
least one of which is projected within 1′ of the lens galaxy and displaced by ∆z such
that κeff /κ = γeff /γ > 0.5. Column 2 labels all the prominent structures along the line
of sight to each lens; "0" refers to a group at the lens redshift. (HE2149 is the only lens
with more than one interloping structure.) Columns 3 -- 7 refer to individual structures,
while Columns 8 -- 10 give the final results after combining all the prominent structures
along the line of sight to each lens (including a group at the lens redshift, if there is
one). Values larger than 0.05 are marked in boldface.
members), MG1654 (seven members), and B2114 (five
members). We more than double the number of mem-
bers for another three previously known groups around
the lenses MG0751 (now 13 members), PG1115 (13 mem-
bers), and B1422 (16 members). These six groups add
to the still small number of all poor groups identified at
intermediate redshifts.
We determine the kinematics of the six groups, includ-
ing their mean velocities, velocity dispersions, and pro-
jected spatial centroids. For the newly discovered groups,
we quantify these properties for the first time. For the
other three groups, the increased membership allows us
to make more robust estimates of the kinematic proper-
ties of the groups than previously possible. The highest
velocity dispersions we measure (320 to 470 km s−1 for
MG0751, PG1115, and B1422) are consistent with those
of nearby dynamically-evolved, X-ray luminous groups
(MZ98), while the lower velocity dispersions (110 to 200
km s−1) are more typical of dynamically younger groups
at low redshift.
In the two cases where a diffuse X-
ray component has been measured (PG1115 and B1422;
Grant et al. 2004), the X-ray temperatures and our ve-
locity dispersions are consistent with the local σr-TX re-
lation (MZ98).
To understand the evolution of groups and their galax-
ies, it is important to determine the relation of the bright-
est group galaxy (BGG) to the group potential.
(In
four of the six groups, MG0751, BRI0952, PG1115, and
B1422, the lens galaxy is not the BGG.) We find that
the BGG generally lies off the center of the group po-
tential and occupies an orbit indistinguishable from the
other group members. This result is surprising in com-
parison with nearby, X-ray luminous groups, in which
the BGG is always a giant elliptical galaxy occupying
the center of the potential, with an orbit distinct from
the other group members (ZM98). However, most of the
effect we see comes from the three groups with lower ve-
locity dispersions. In two (MG0751 and PG1115) of the
three highest velocity dispersion groups, the BGGs lie
within the errors of the group centroid, suggesting that
at least these systems are comparable to dynamically-
evolved poor groups in the local universe.
We use our detailed observations of the groups to as-
sess how environments affect gravitational lens models.
A key ingredient is an accurate determination of any off-
set between the lens galaxy and the group centroid on
the sky. In MG0751, PG1115, B1422, and MG1654, the
lens galaxy is offset spatially from the group centroid.
Obtaining a larger sample (which is underway) to deter-
mine the full distribution of lens vs. group offsets will
be important for understanding how lens environments
affect statistical quantities such as the quad/double ratio
and lensing constraints on ΩΛ (see KZ04).
To quantify environmental contributions to lens poten-
tials in more detail, we estimate the convergence (gravi-
tational focusing) and shear (tidal distortions) from each
group. We consider two different models of the group
mass distribution that bound the extremes of dynami-
cal states of groups. The members of young groups are
likely to still have large dark matter halos and the group
mass may be dominated by the dark matter halos of the
individual member galaxies. In this approach, we cali-
brate the shear and convergence from each galaxy based
on observations of weak lensing in the Sloan Digital Sky
Survey (Sheldon et al. 2005), and then sum the contri-
butions appropriately. As the group evolves, these halos
Lens Galaxy Environments
21
may be truncated via interactions, so the mass will be
redistributed into a common group halo that may be the
dominant group mass component. In this approach, we
approximate that halo as an isothermal sphere and use
the measured group centroid and velocity dispersion to
compute the convergence and shear.
At least three of the lenses in our sample (MG0751,
PG1115, and B1422) have convergences and shears large
enough (κ, γ ≥ 0.05) to indicate that the environment
plays a significant role in the lens potential. For these
systems, our survey substantially improves the observa-
tional constraints that will be needed to make detailed
lens models that properly include environmental effects.
Remarkably, the high shear and convergence values occur
in the quad lens systems, while the environments of the
double lenses are relatively weak. This result suggests
that environment may affect the relative numbers of quad
and double lenses, a topic much debated in the literature
(see King et al. 1996; Kochanek 1996b; Keeton et al.
1997; Rusin & Tegmark 2001; Cohn & Kochanek 2004;
Keeton & Zabludoff 2004). For the other lenses, the con-
clusion that environment does not significantly affect the
lens potential is also valuable: constraining previously
unknown environmental terms to be near zero will still
improve lens models.
For the first time, we present a systematic assessment
of whether structures along the line of sight to lens sys-
tems are important for lensing. We show that interlop-
ing structures can in principle affect lens models over a
wide range of spatial and redshift offsets. Our pencil-
beam survey is ideally suited to identifying such struc-
tures if they are present. We find that at least four out of
eight lenses have prominent line-of-sight structures, i.e.,
groups whose spatial and redshift offsets place them in
the "zone of influence" of the lens. MG0751, PG1115,
and B1422 each have one substantial group along the
line of sight, while HE2149 has three groups at differ-
ent redshifts. Of these, the interloping group in MG0751
has a significant effect on the lens potential. Our sur-
vey is actually biased against interloping groups (and is
not complete), so finding that at least one of eight lenses
(∼ 10%) is affected by projected structures is intriguing
and worth further study.
We thank the staff of the Magellan Observatory for
their tireless efforts on behalf of this project. We appre-
ciate the advice and spectral templates given by Marc
Postman, as well as the helpful comments provided by
Chris Impey. We thank the anonymous referee for care-
ful and constructive comments. This work was sup-
ported by NSF grant #AST-0206084 and NASA LTSA
award #NAG5-11108.
IM acknowledges the support
of the Martin F. McCarthy Scholarship in Astrophysics
awarded by the Vatican Observatory.
Allington-Smith, J. R., Breare, J. M., Ellis, R. S., Parry, I. R., &
Gorenstein, M. V., Shapiro, I. I., & Falco, E. E. 1988, ApJ, 327,
Shaw, G. D. 1990, Proc. SPIE, 1235, 691
693
Angonin-Willaime, M.-C., Hammer, F., & Rigaut, F. 1993,
Grant, C. E., Bautz, M. W., Chartas, G., & Garmire, G. P. 2004,
Gravitational Lenses in the Universe, 85
ApJ, 610, 686
REFERENCES
Augusto, P., et al. 2001, MNRAS, 326, 1007
Bar-Kana, R. 1996, ApJ, 468, 17
Bar-Kana, R. 1997, ApJ, 489, 21
Beers, T. C., Flynn, K., & Gebhardt, K. 1990, AJ, 100, 32
Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393
Blandford, R., Surpi, G., & Kundi'c, T. 2001, in Gravitational
Lensing: Recent Progress and Future Goals (ASP Conference
Proceedings, vol. 237), ed. T. G. Brainerd & C. S. Kochanek, p.
65
Browne, I. W. A., Wilkinson, P. N., Patnaik, A. R., & Wrobel, J.
M. 1998, MNRAS, 293, 257
Bruzual, G., & Charlot, S. 1993, ApJ, 405, 538
Burud, I., et al. 2002, A&A, 383, 71
Carlberg, R. G., Yee, H. K. C., Morris, S. L., Lin, H., Hall, P. B.,
Patton, D. R., Sawicki, M., & Shepherd, C. W. 2001, ApJ, 552,
427
Chae, K.-H., Mao, S., & Augusto, P. 2001, MNRAS, 326, 1015
Chae, K. 2003, MNRAS, 346, 746
Chiba, M. 2002, ApJ, 565, 17
Cohn, J. D., Kochanek, C. S., McLeod, B. A., & Keeton, C. R.
2001, ApJ, 554, 1216
Cohn, J. D., & Kochanek, C. S. 2004, ApJ, 608, 25
Dalal, N., & Kochanek, C. S. 2002, ApJ, 572, 25
Dalal, N. & Watson, C. R. 2004, astro-ph/0409438
Danese, L., de Zotti, G., & di Tullio, G. 1980, A&A, 82, 322
Dobler, G., & Keeton, C. R. 2005, astro-ph/0502436
Fassnacht, C. D., & Lubin, L. M. 2002, AJ, 123, 627
Fassnacht, C. D., et al. 2004, Proceedings of IAU Symposium
Impact of Gravitational Lensing on Cosmology (also
225:
astro-ph/0409086)
Fassnacht, C. D., et al., astro-ph/0510728
Faure, C., Alloin, D., Kneib, J. P., & Courbin, F. 2004, A&A, 428,
741
Ferreras, I., Saha, P., & Williams, L. L. R. 2005, astro-ph/0503168
Fischer, P., Schade, D., & Barrientos, L. F. 1998, ApJ, 503, L127
Fukugita, M., Shimasaku, K., & Ichikawa, T. 1995, PASP, 107, 945
Gerke, B. F., et al. 2005, ApJ, 625, 6
Hege, E. K., Hubbard, E. N., Strittmatter, P. A., & Worden, S. P.
1981, ApJ, 248, L1
Henry, J. P., & Heasley, J. N. 1986, Nature, 321, 139
Hogg, D. W., & Blandford, R. D. 1994, MNRAS, 268, 889
Holder, G. P., & Schechter, P. L. 2003, ApJ, 589, 688
Impey, C. D., Foltz, C. B., Petry, C. E., Browne, I. W. A., &
Patnaik, A. R. 1996, ApJ, 462, L53
Impey, C. D., Falco, E. E., Kochanek, C. S., Leh´ar, J., McLeod,
B. A., Rix, H.-W., Peng, C. Y., & Keeton, C. R. 1998, ApJ, 509,
551
Keeton, C. R., Kochanek, C. S., & Seljak, U. 1997, ApJ, 482, 604
Keeton, C. R., & Kochanek, C. S. 1997, ApJ, 487, 42
Keeton, C. R., Kochanek, C. S., & Falco, E. E. 1998, ApJ, 509, 561
Keeton, C. R., Christlein, D., & Zabludoff, A. I. 2000, ApJ, 545,
129
Keeton, C. R. 2003, ApJ, 584, 664
Keeton, C. R., & Zabludoff, A. I. 2004, ApJ, 612, 660
Keeton, C. R., Gaudi, B. S., & Petters, A. O. 2005a,
astro-ph/0503452
Keeton, C. R., et al., 2005b, in preparation
King, L. J., Browne, I. W. A., & Wilkinson, P. N. 1996, IAU
Symp. 173: Astrophysical Applications of Gravitational Lensing,
173, 191
King, L. J., Browne, I. W. A., Marlow, D. R., Patnaik, A. R., &
Wilkinson, P. N. 1999, MNRAS, 307, 225
Kneib, J., Cohen, J. G., & Hjorth, J. 2000, ApJ, 544, L35
Kochanek, C. S. 1991, ApJ, 373, 354
Kochanek, C. S. 1995, ApJ, 445, 559
Kochanek, C. S. 1996a, ApJ, 466, 638
Kochanek, C. S. 1996b, ApJ, 473, 595
Kochanek, C. S., et al. 2000, ApJ, 543, 131
Kochanek, C. S. & Schechter, P. L. 2003,
in Measuring and
Modeling the Universes (Carnegie Observatories Astrophysics
Series, vol. 2), ed. W. L. Freedman (also astro-ph/0306040)
Kundi´c, T., Cohen, J. G., Blandford, R. D., & Lubin, L. M. 1997,
AJ, 114, 507
22
Momcheva et al.
Kundi´c, T., Hogg, D. W., Blandford, R. D., Cohen, J. G., Lubin,
L. M., & Larkin, J. E. 1997, AJ, 114, 2276
Kurtz, M. J., & Mink, D. J. 1998, PASP, 110, 934
Langston, G. I., et al. 1988, BAAS, 20, 1001
Langston, G. I., et al. 1989, AJ, 97, 1283
Leh´ar, J., McMahon, R. G., Irwin, M. 1993, AAS, 183, 3307L
Leh´ar, J., et al. 1997, AJ, 114, 48L
Linder, E. V. 2004, Phys. Rev. D, 70, 043534
Lopez, S., Wucknitz, O., & Wisotzki, L. 1998, A&A, 339, L13
Mao, S., & Schneider, P. 1998, MNRAS, 295, 587
McMahon, R. & Irwin, M. 1992, Gemini, 36, 1M
Metcalf, R. B., & Madau, P.. 2001, ApJ, 563, 9
Mitchell, J. L., Keeton, C. R., Frieman, J. A, & Sheth, R. K. 2005,
ApJ, 622, 81
Seljak, U. 1994, ApJ, 436, 509
Sheldon, E., et al. 2004, AJ, 127, 2544
Soucail, G., Kneib, J.-P., Jaunsen, A. O., Hjorth, J., Hattori, M.,
& Yamada, T. 2001, A&A, 367, 741
Tonry, J. L. 1998, AJ, 115, 1
Tonry, J. L., & Kochanek, C. S. 1999, AJ, 117, 2034
Tonry, J. L., & Kochanek, C. S. 2000, AJ, 119, 1078
Treu, T., & Koopmans, L. V. E. 2004, ApJ, 611, 739
Turner, E. L. 1990, ApJ, 365, L43
van Waerbeke, L., & Mellier, Y. 2003, astro-ph/0305089
Wambsganss,
J., Bode, P., & Ostriker,
J. P.,
2004,
astro-ph/0405147
Williams, K., Momcheva, I., Keeton, C. R., & Zabludoff, A. I. 2005,
in preparation
Morgan, N. D., Kochanek, C. S., Pevunova, O. & Schechter, P. L.
2005, AJ, 129, 2531
Mulchaey, J. S. & Zabludoff, A. I. 1998, ApJ, 496, 73
Omont, A., McMahon, R. G., Cox, P., Kreysa, E., Bergeron, J.,
Pajot, F., & Storrie-Lombardi, L. J. 1996, A&A, 315, 1
Wilman, D. J., et al. 2005, astro-ph/0501183
Winn, J. N., Hall, P. B., & Schechter, P. L. 2003, ApJ, 597, 672
Winn, J. N., Hewitt, J. N., Patnaik, A. R., Schechter, P. L.,
Schommer, R. A., L´opez, S., Maza, J., & Wachter, S. 2001, ApJ,
121, 122
Patnaik, A. R., Browne, I. W. A., Wilkinson, P. N., & Wrobel, J.
Wisotzki, L., Kohler, T., Lopez, S., & Reimers, D. 1996, A&A, 315,
M. 1992a, MNRAS, 254, 655
L405
Patnaik, A. R., Browne, I. W. A., Walsh, D., Chaffee, F. H., &
Foltz, C. B. 1992, MNRAS, 259, 1P
Weymann, R. J., et al. 1980, Nature, 285, 641
Wilkinson, P. N., Browne, I. W. A., Patnaik, A. R., Wrobel, J. M.,
Postman, M., Lauer, T. R., Oegerle, W., & Donahue, M. 2002,
& Sorathia, B. 1998, MNRAS, 300, 790
ApJ, 579, 93
Premadi, P., & Martel, H. 2004, ApJ, 611, 1
Refsdal, S. 1964, MNRAS, 128, 307
Rusin, D., et al. 2001, ApJ, 557, 594
Rusin, D., & Tegmark, M. 2001, ApJ, 553, 709
Rusin, D., et al. 2003, ApJ, 587, 143
Rusin, D., & Kochanek, C. S. 2005, ApJ, 623, 666
Saha, P. 2000, AJ, 120, 1654
Schechter, P. L., et al. 1997, ApJ, 475, L85
Schneider, P., Ehlers, J., & Falco, E. E. 1992, Gravitational Lenses
(Berlin: Springer)
Yahil, A. & Vidal, N. V. 1977, ApJ, 214, 347
Young, P., Deverill, R. S., Gunn, J. E., Westphal, J. A., & Kristian,
J. 1981, ApJ, 244, 723
Young, P., Gunn, J. E., Oke, J. B., Westphal, J. A., & Kristian, J.
1981, ApJ, 244, 736
Zabludoff, A. I., & Mulchaey, J. S. 1998, ApJ, 496, 39
Zabludoff, A. I., & Mulchaey, J. S. 1998, ApJ, 498, L5
FORMALISM FOR CONVERGENCE AND SHEAR
APPENDIX
Keeton (2003) presents a formalism for deriving a simple analytic estimates of the effective convergence κeff and
shear γeff produced by a perturber somewhere along the line of sight to a lens. The same formalism can be used for
all the situations considered in the main text (see §§4.4 and 4.5). The only assumption is that the convergence and
shear are small, so that we can work at first order in both quantities. We will show below that there is a context in
which this assumption breaks down, but with no effect on our conclusions.
If the perturber were at the same redshift as the lens galaxy, it would produce a convergence and shear given by
κ =
γc =
γs =
1
2 (cid:18) ∂2ϕpert
∂x2 +
2 (cid:18) ∂2ϕpert
∂x2
−
1
∂2ϕpert
∂y2 (cid:19) ,
∂2ϕpert
∂y2 (cid:19) ,
∂2ϕpert
∂x ∂y
,
γ =pγ2
θγ =
s ,
c + γ2
tan−1(cid:18) γs
γc(cid:19) ,
1
2
(A1)
(A2)
(A3)
(A4)
(A5)
where ϕpert is the lens potential of the perturber. Here γ is the shear strength, and θγ the shear direction (which we
measure North through East). If the perturber lies at a different redshift zpert 6= zl, then the convergence and shear
are modified to the effective values (Keeton 2003)
κeff =
γeff =
(1 − βκ)2 − (βγ)2
(1 − β)(cid:2)κ − β(κ2 − γ2)(cid:3)
(1 − βκ)2 − (βγ)2 ,
(1 − β)γ
where
β =
D(z1, z2)
D(0, z2)
D(0, zs)
D(z1, zs)
,
z1 = min(zl, zpert) ,
z2 = max(zl, zpert) ,
,
(A6)
(A7)
(A8)
(A9)
(A10)
Lens Galaxy Environments
23
where D(z1, z2) is the angular diameter distance between redshifts z1 and z2. Note that when zpert = zl, we have
β = 0 and so we recover κeff = κ and γeff = γ. Moving the perturber away from the lens redshift increases β and
decreases κeff and γeff .
In certain circumstances, the effective convergence and shear can apparently go negative, which seems puzzling.
The sign flip occurs only when the denominator in eqs. (A6) -- (A7) goes negative. This can happen only when κ and
γ are sufficiently large -- in particular, only when the line of sight passes through the strong lensing regime of the
perturber. (For an isothermal sphere [below], this corresponds to κ, γ ≥ 0.5.) In this case, the lensing critical curves
of the perturber would merge with those of the main lens galaxy, which would completely change the configuration
of lensed images. In other words, our formalism breaks down when the "perturbation" is sufficiently strong, but if it
were that strong it would (presumably) be known already.
If κ and γ are small, then we can expand eqs. (A6) -- (A7) to first order write
κeff
κ
≈
γeff
γ
≈ 1 − β .
(A11)
We can think of κeff/κ and γeff /γ as the "normalized" convergence and shear -- the actual perturbation strength,
normalized by the value that would apply if the perturber were at the same redshift as the lens galaxy. We use this
quantity in the text as a simple way to characterize the redshift dependence of the perturbation strength.
We now specify how to determine the net convergence and shear when multiple perturbers are present. This is
relevant when we assess the effects of a group around a lens by considering the member galaxies (§4.4.2), and also
when we consider the effects of structures along the line of sight (§4.5.2). We see from eqs. (A1) -- (A3) that the
quantities κ, γc, and γs are linear in the perturber potential. The shear is a combination of γc and γs that actually
corresponds to a rank-2 traceless tensor, or a headless vector (headless because it is invariant under rotation by 180◦).
(See Schneider et al. 1992 for more discussion.) Thus, if we know the effective convergence κeff,i and shear γeff,i, as
well as the shear position angle θγ,i (North through East), for a set of perturbers, then the proper way to compute
the net effects is as follows:
κeff,i .
γeff,i cos 2θγ,i ,
γeff,i sin 2θγ,i ,
κtot =Xi
γc,tot =Xi
γs,tot =Xi
γtot =qγ2
1
2
c,tot + γ2
s,tot ,
θγ,tot =
tan−1(γs,tot/γc,tot) .
(A12)
(A13)
(A14)
(A15)
(A16)
Note that κeff,i ≥ 0 for any real perturber, so the terms in the convergence sum all go in the same direction. By
contrast, the cosine and sine factors mean that terms in the shear sums may add or cancel. Hence, incompleteness in
our sample can only cause us to underestimate κtot, but it may cause us to over- or underestimate γtot.
It is worthwhile to recall how convergence and shear affect lens models. Convergence is largely responsible for
systematic biases in lens models (KZ04), through the mass sheet degeneracy (Gorenstein et al. 1988; Saha 2000). The
biases can be thought of as simple rescalings of model parameters, such as:
β ∝ (1 − κtot) ,
h ∝ (1 − κtot) ,
µ ∝ (1 − κtot)−2 ,
(A17)
(A18)
(A19)
where β is a mass parameter related to the lens velocity dispersion, h is the Hubble parameter, and µ is the image
magnification (see KZ04 and references therein). Neglect of convergence is the main source of biases in lensing results
for quad lenses. Double lenses, by contrast, are so under-constrained that poor knowledge of convergence and shear
leads to lens models that are just plain wrong. In both cases, detailed observations of lens environments are necessary
to derive the constraints necessary to make lens models reliable.
While the formalism presented so far is fully general, it is valuable to discuss two particular perturber models. First,
if we can approximate a perturber as an isothermal sphere then we can easily relate measurable quantities to κ and
γ. Specifically, for an isothermal sphere with velocity dispersion σ and impact parameter b, we have
κ = γ = 1.44 × 10−5(cid:18) 1′′
b (cid:19)(cid:18) σ
km/s(cid:19)2 D(zpert, zs)
D(0, zs)
.
(A20)
We can measure the line-of-sight velocity dispersion σr of the group and the offset b between the lens galaxy and group
centroid, use them to determine κ and γ, and finally fold in the redshift difference as above to determine κeff and γeff .
The second specific model we consider is a power law density profile calibrated by weak lensing. In a large and
detailed analysis of galaxy -- galaxy weak lensing in the Sloan Digital Sky Survey, Sheldon et al. (2005) present the
24
Momcheva et al.
average shear as a function of radius for a sample of 127,001 lens galaxies. They find that the shear profile is consistent
with a simple power law
γ(R) =
R−α ,
(A21)
A
Σcrit
where Σcrit = (c2Dos)/(4πGDolDls) is the critical surface density for lensing, which carries all the dependence on the
lens and source redshifts through the angular diameter distances Dol, Dos, and Dls between the observer, lens, and
source. Sheldon et al. tabulate the values of α and A for galaxies in three luminosity bins.6 For any given galaxy that
we observe, we convert from our measured I-band apparent magnitude to absolute magnitude in the SDSS i-band,
using the color and K-corrections computed with Bruzual & Charlot (1993) spectral synthesis models. (The luminosity
bins used by Sheldon et al. are wide enough that small systematic uncertainties in the magnitudes do not shift galaxies
between bins.) We then look up the values of A and α for each luminosity from Table 2 of Sheldon et al. We use the
lens and source redshifts to compute Σcrit and hence γ(R). The last thing we need is the convergence. For a power
law γ(R) ∝ R−α, there is a very simple relation between shear and convergence:
Thus, it is straightforward to compute κ and γ for each galaxy. We can then factor in the redshift distance relative to
the main lens galaxy as described above.
κ(R) =
2 − α
α
γ(R) .
(A22)
6 Sheldon et al. (2005) actually tabulate power law parameters for the galaxy -- mass correlation function rather than the shear directly,
but their formalism makes it straightforward to convert back to A and α. At any rate, those parameters are more fundamental in terms of
what they measure. One additional technical point is that Sheldon et al. quote lengths using comoving distances, but we convert those to
angular diameter distances in our analysis.
This figure "f6_sm.jpg" is available in "jpg"(cid:10) format from:
http://arxiv.org/ps/astro-ph/0511594v1
|
astro-ph/0409231 | 1 | 0409 | 2004-09-09T18:39:54 | Magnetically Driven Accretion Flows in the Kerr Metric IV: Dynamical Properties of the Inner Disk | [
"astro-ph"
] | This paper continues the analysis of a set of general relativistic 3D MHD simulations of accreting tori in the Kerr metric with different black hole spins. We focus on bound matter inside the initial pressure maximum, where the time-averaged motion of gas is inward and an accretion disk forms. We use the flows of mass, angular momentum, and energy in order to understand dynamics in this region. The sharp reduction in accretion rate with increasing black hole spin reported in Paper I of this series is explained by a strongly spin-dependent outward flux of angular momentum conveyed electromagnetically; when a/M > 0.9, this flux can be comparable to the inward angular momentum flux carried by the matter. In all cases, there is outward electromagnetic angular momentum flux throughout the flow; in other words, contrary to the assertions of traditional accretion disk theory, there is in general no "stress edge", no surface within which the stress is zero. The retardation of accretion in the inner disk by electromagnetic torques also alters the radial distribution of surface density, an effect that may have consequences for observable properties such as Compton reflection. The net accreted angular momentum is sufficiently depressed by electromagnetic effects that in the most rapidly-spinning black holes mass growth can lead to spindown. Spinning black holes also lose energy by Poynting flux; this rate is also a strongly increasing function of black hole spin, rising to 10% or more of the rest-mass accretion rate at very high spin. As the black hole spins faster, the path of the Poynting flux changes from being predominantly within the accretion disk to predominantly within the funnel outflow. | astro-ph | astro-ph |
Magnetically Driven Accretion Flows in the Kerr Metric IV:
Dynamical Properties of the Inner Disk
Julian H. Krolik
Physics and Astronomy Department
Johns Hopkins University
Baltimore, MD 21218
and
John F. Hawley
Astronomy Department
University of Virginia
P.O. Box 3818, University Station
Charlottesville, VA 22903-0818
and
Shigenobu Hirose
Physics and Astronomy Department
Johns Hopkins University
Baltimore, MD 21218
[email protected]; [email protected]; [email protected]
ABSTRACT
This paper continues the analysis of a set of general relativistic 3D MHD
simulations of accreting tori in the Kerr metric with different black hole spins.
We focus on bound matter inside the initial pressure maximum, where the time-
averaged motion of gas is inward and an accretion disk forms. We use the flows
of mass, angular momentum, and energy in order to understand dynamics in this
region. The sharp reduction in accretion rate with increasing black hole spin
reported in Paper I of this series is explained by a strongly spin-dependent out-
ward flux of angular momentum conveyed electromagnetically; when a/M ≥ 0.9,
this flux can be comparable to the inward angular momentum flux carried by the
matter. In all cases, there is outward electromagnetic angular momentum flux
-- 2 --
throughout the flow; in other words, contrary to the assertions of traditional ac-
cretion disk theory, there is in general no "stress edge," no surface within which
the stress is zero. The retardation of accretion in the inner disk by electromag-
netic torques also alters the radial distribution of surface density, an effect that
may have consequences for observable properties such as Compton reflection.
The net accreted angular momentum is sufficiently depressed by electromagnetic
effects that in the most rapidly-spinning black holes mass growth can lead to
spindown. Spinning black holes also lose energy by Poynting flux; this rate is
also a strongly increasing function of black hole spin, rising to & 10% of the
rest-mass accretion rate at very high spin. As the black hole spins faster, the
path of the Poynting flux changes from being predominantly within the accretion
disk to predominantly within the funnel outflow.
Subject headings: Black holes - magnetohydrodynamics - instabilities - stars:accretion
1.
Introduction
In this paper, we continue the analysis of a series of simulations of accretion disks in
the Kerr metric introduced in De Villiers, Hawley, & Krolik (2003; hereafter Paper I). The
emphasis in this paper will be on the properties of the accretion flow inside the initial pressure
maximum, where the net time-averaged motion is towards the black hole. This region of
the accretion flow, in the terminology introduced in Paper I, comprises the inner part of
the main disk body, the inner torus, the plunging region, and the corona. It is arguably
the most important region of all in an accretion disk because it is here that most of the
energy is released that ultimately powers the observed radiation and outflows. Our goal is to
understand the peculiarities of dynamics in this region; our principal tool will be to analyze
the flows of three conserved quantities, mass, energy, and angular momentum. By so doing,
we aim to shed light on the accretion process, how it is moderated by black hole spin, and
how energy and angular momentum can be extracted from the spinning black hole.
Accretion onto black holes differs from accretion onto ordinary objects in several ways,
but two stand out: the absence of stable circular orbits inside of a critical radius, and the
presence of frame-dragging due to black hole spin. The absence of stable circular orbits
between the radius of marginal stability rms and the event horizon means that in this region
matter flows in rapidly and the mass surface density is much smaller than in the main disk
body where matter orbits stably. We may therefore talk about the disk having an "inner
edge" in the vicinity of rms. However, as emphasized by Krolik & Hawley (2002), where one
places this inner edge depends on precisely which disk property is under consideration. For
-- 3 --
example, two distinct edges can be defined by dynamical properties: the turbulence edge,
where a transition takes place in the character of magnetic field evolution -- from turbulence-
driven evolution in the disk proper to evolution by magnetohydrodynamic (MHD) flux-
freezing, and the stress edge, within which the stress in the accreting fluid goes to zero so
that there is no outward transport of angular momentum. Another two are more closely
related to photon emission properties: the reflection edge, where production of Compton
reflection features ceases, and the radiation edge, where intrinsic luminosity detected by
distant observers goes to zero due to reduced radiative efficiency or relativistic effects. Fully
relativistic simulations can explore the relation between these observable disk edges, and the
detailed dynamics of the accretion flow. Defining the location and nature of these edges is
central to linking disk dynamical models to disk observables because these edges determine,
among other things, the global radiative efficiency of accretion.
The other salient feature of relativistic accretion is the presence of frame-dragging when
the black hole spins. Frame-dragging introduces a new dynamical effect into the accretion
flow, and the spin of the black hole represents a potential source of energy output beyond
that of the accretion flow itself. In a previous paper of this series (De Villiers et al. 2004) we
presented evidence from the simulations that black hole spin energy helps to power axial jets.
Here we focus on the flow of angular momentum and energy through the disk and corona,
the influence of the black hole spin on the main accretion flow, and how the character of
that flow depends on that spin. We find that frame-dragging and its associated effects can
have a dramatic impact on the rate of accretion, the structure of the inner disk, and the
fundamental flow of energy and angular momentum.
This paper is organized as follows: After a quick summary of the simulations and the
numerical diagnostics we used to study them (§2), we devote one section each to the flow
through the disk of the three conserved quantities, mass (§3), angular momentum (§4), and
energy (§5). In each case, we try to highlight how much goes where, how it is split between
matter and electromagnetic contributions, and how the situation changes with changing black
hole spin. Section 6 summarizes our findings and comments on astrophysical implications of
the numerical results.
2. Preliminaries
2.1. Simulation definition
We solve the equations of ideal MHD in the metric of a rotating black hole. The specific
form of the equations we solve, and the numerical algorithm incorporated into the GRMHD
-- 4 --
code are described in detail in De Villiers & Hawley (2003). For reference we reiterate here
the key terms and the definitions of the primary code variables.
We work in the Kerr metric, expressed in Boyer-Lindquist coordinates, (t, r, θ, φ), for
which the line element has the form, ds2 = gtt dt2 + 2 gtφ dt dφ + grr dr2 + gθθ dθ2 + gφφ dφ2.
We use the metric signature (−, +, +, +). The determinant of the 4-metric is g, and √−g =
α√γ, where α is the lapse function, α = 1/√−gtt, and γ is the determinant of the spatial
3-metric. We follow the usual convention of using Greek characters to denote full space-time
indices and Roman characters for purely spatial indices. We use geometrodynamic units
where G = c = 1; time and distance are in units of the black hole mass, M.
The state of the relativistic test fluid at each point in the spacetime is described by
its density, ρ, specific internal energy, ǫ, 4-velocity, U µ, and isotropic pressure, P . The
relativistic enthalpy is h = 1 + ǫ + P/ρ. The pressure is related to ρ and ǫ through the
equation of state of an ideal gas, P = ρ ǫ (Γ − 1), where Γ is the adiabatic exponent. For
these simulations we take Γ = 5/3. The magnetic field of the fluid is described by two
sets of variables, the constrained transport magnetic field, Fjk = [ijk]Bi, and magnetic field
4-vector √4π bµ = ∗F µνUν. The ideal MHD condition requires U νFµν = 0. The magnetic
field bµ is included in the definition of the total four momentum, Sµ = (ρ h + kbk2) W Uµ,
where W is the Lorentz factor. We define the transport velocity as V i = U i/U t.
In Paper I we presented results of a series of high- and low-resolution simulations, the
KD (Keplerian Disk) set of disk models. These models have an initial condition consisting
of an isolated gas torus orbiting near the black hole, with a pressure maximum at r ≈ 25M,
and a slightly sub-Keplerian initial distribution of angular momentum throughout. The
initial magnetic field consists of loops of weak poloidal field lying along isodensity surfaces
within the torus. In this paper the emphasis will be on the high-resolution models designated
KD0, KDI, KDP, and KDE, which differ in the spin of the black hole around which they
orbit, with a/M = 0, 0.5, 0.9 and 0.998 respectively. These models used 192 × 192 × 64
(r, θ, φ) grid zones. The radial grid is set using a hyperbolic cosine function to maximize
the resolution near the inner boundary, which is at rin = 2.05 M, 1.90 M, 1.45 M, and
1.175 M for models KD0, KDI, KDP, and KDE, respectively. The outer radial boundary
is set to rout = 120M in all cases. The θ-grid ranges over 0.045 π ≤ θ ≤ 0.955 π, with
an exponential grid spacing function that concentrates zones near the equator; reflecting
boundary conditions are enforced in the θ-direction. The φ-grid spans the quarter plane,
0 ≤ φ ≤ π/2.
-- 5 --
2.2. Simulation diagnostics
Three-dimensional numerical simulations generate an enormous amount of data, only a
representative sample of which can be examined. Our analysis is based on a specific set of
volume- and shell-averaged history data, taken every M in time, and complete data snapshots
taken every 80M in time. Although the details of the history calculations are given in Paper
I, we provide here a brief summary to clarify the calculations of mass, energy, and angular
momentum transport in the inner torus and plunging region.
The normalized shell-average of a quantity Q at radius r is calculated using
hQiA(r, t) =
1
A Z dθ dφ √−g Q,
(1)
where the bounds of integration range over the full θ and φ computational domains and A
is the area of the shell. The density-weighted shell-average of a quantity will be denoted
hQiρ ≡ hρQi/hρi. The flux F of a given quantity through a shell at radius r is computed
using a similar integral, but the result is not normalized to the shell area. For example, the
accretion rate is defined
M (r) = hρU ri = −Z dθ dφ √−g ρU r
In addition to the accretion rate (mass flux) we consider the total energy flux,
hT r
ti = hρ h U r Uti +
1
4 π hF r
α Ft
αi
and the total angular momentum flux,
hT r
φi = hρ h U r Uφi +
1
4 π hF r
α Fφ
αi
(2)
(3)
(4)
The first term in each of these sums represents the contribution to the flux from the fluid. The
second term represents the contribution due to the magnetic field, where the electromagnetic
field strength tensor, F µν, reduces to combinations of transport velocities and constrained
transport (CT) variables, as discussed in Paper I.
We denote the time-average of a shell-averaged quantity (or flux) as
hhQii(r) =
1
tmax − tmin Z tmax
tmin
dthQi (r, t).
(5)
M ) is denoted by a single bracket. The
The time-average of an integrated quantity (e.g.,
bounds of integration are chosen to analyze the late-time state of the accretion flow after the
-- 6 --
initial transient described in Paper I has passed. Unless otherwise stated, tmin = 2000 M,
which corresponds to about 2.5 orbits at the initial pressure maximum, and tmax = 8100 M,
the end-time of the simulations. Such time-averages are important in establishing persistent
features in various diagnostics; by their nature, these averages suppress short timescale
variations, so we will also be interested in rms fluctuations about the time averages.
The time-history data contains the total contribution on a shell, but does not distinguish
between contributions from bound material (i.e., in the disk and corona) and unbound (in
the axial funnel and funnel-wall jet). For the purposes of this paper, where we are focused
on the accretion flow, this is an important distinction. Thus, for many quantities, we will
compute time-averages using the full data dumps, with integrations restricted to zones where
the specific total energy satisfies −h Ut < 1, i.e., including only the contributions from bound
material.
3. Mass Accretion
3.1.
Inflow equilibrium and fluctuations about equilibrium
The initial conditions in these simulations consist of an isolated torus of gas orbiting
around the black hole. A relevant question is the degree to which the resulting accretion
flow from such an initial state replicates a time-steady accretion disk. Outside the initial
torus pressure maximum (r = 25 M), the gas must move outward as it accepts angular
momentum from matter at smaller radius. This portion of the flow certainly cannot represent
the behavior of a statistically time-steady accretion disk fed from large distances. Inside the
initial pressure maximum, however, there is inflow throughout the simulation. Because the
inner edge of the initial torus is at r = 15 M, well outside the marginally stable orbit, a
significant amount of time is required before the flow reaches the black hole and a Keplerian
disk forms. After that point in time we can investigate whether the accretion flow has
reached an approximate statistical equilibrium.
We first examine the rate at which gas is fed into the region r < 15 M, namely the
region inside of the inner boundary of the initial torus. Some sense of the time history can
be seen in the spacetime diagrams of density (Fig. 5 of Paper I) which show both that density
in the inner region is a highly dynamical quantity and that higher densities are found near
the black hole in the high-spin models. Another view is given in Figure 1, which presents
the total mass inside of r = 15 M as a function of time for each of the four models. In all
cases significant matter infall begins by t = 500 M and the mass begins to build up. After
t = 2000 M in the a/M = 0, 0.5, and 0.9 models, the disk in this region appears to settle
-- 7 --
into a rough steady state of constant total mass. The high-spin model (a/M = 0.998) stands
out:
in that model, the mass inside r = 15 M continues to grow throughout most of the
simulation.
More detailed information can be obtained by examining the time-averaged accretion
rate as a function of radius h M (r)i. We begin the time average after t = 2000 M, i.e., after
the first 25% of the total evolution time. Figure 2 plots h M (r)i for each of the four models.
If the disk is in a statistically steady state, the time-averaged accretion rate should be nearly
constant as a function of radius. This is clearly true inside of r = 15 M in the slowly-rotating
simulations (a/M = 0, 0.5), and nearly so for the a/M = 0.9 model, but is not the case for
the a/M = 0.998 simulation. Combining this result with Figure 1, we conclude that a state
of local accretion equilibrium was attained for r . 10 -- 15 M in three of the four simulations.
Although the a/M = 0.998 simulation ran for the same amount of time, namely 10 orbits at
r = 25M and hundreds of orbits in the inner disk, it did not reach a time-steady state with
regard to mass flow.
Despite the constancy with radius of the time-averaged accretion rate, the instantaneous
accretion rate varies considerably with radius and with time. For example, Fig. 14 of Paper
I shows that the accretion rate into the black hole varies substantially as a function of time,
with significant power across a wide range of timescales. The dashed curves in Figure 2 show
the rms deviation about the local mean accretion rate as a function of radius. The relative
magnitude of accretion rate fluctuations (as compared to the mean) increases toward larger
radii. Some of this effect may be attributed to the initial conditions and the absence of a
mean accretion rate for r & 20M. However, a part of this effect is also likely to be real. It
is significant, for example, that in each case the fluctuation level begins to rise well within
the region where the accretion rate is nearly constant as a function of radius.
There is also a trend for the relative size of accretion rate fluctuations at a fixed radius
Mrms/h Mi ≃ 0.5 at r ≃ 10M when
to increase with increasing black hole spin. Whereas
a/M ≤ 0.5, it rises to ≃ 0.7 when a/M ≥ 0.9; the trend is similar at other radii well inside
the initial pressure maximum (r = 25M). Given that the mean accretion rate itself is poorly-
defined when a/M = 0.998, it is not surprising that the rms fluctuation in the accretion rate
in this simulation is comparable to the mean at all radii . 20M.
Figure 2 also illustrates a point made in Paper I: the fraction of the initial torus mass
accreted in a fixed time declines with increasing black hole spin. While 14% was accreted
over the duration of the a/M = 0 simulation, only 3.5% was accreted when a/M = 0.998.
-- 8 --
3.2.
Inflow time
The mass accretion rate describes the bulk flow of matter toward the black hole; the
inflow time is, on the other hand, a property of individual particles that make up that matter.
As with many quantities in fluid dynamics, the inflow time may be studied from either the
Lagrangian or Eulerian points of view. In this context, the "Lagrangian" inflow time is the
(mean) time for fluid elements to move from radius r to the inner boundary; the "Eulerian"
inflow time is the characteristic time for fluid elements to move past a fixed radius r. The
former is of most interest for answering questions such as, "Does radiation have time to
escape before a fluid cell passes the event horizon?"; the latter is of most interest for gauging
the time for the surface density at some radius to equilibrate with accretion fluctuations.
We compute both inflow times from the mean inflow velocity obtained from a time-
averaged density-weighted shell-integral. The quantity hhU riρi is shown in Figure 3a for all
four simulations. The inflow velocity is smaller with greater spin a/M at all radii less than
about r = 15 M. The slope as a function of r also decreases with increasing black hole spin.
Note too that, except for the highest-spin model, hhU riρi is very nearly a pure power-law in
radius that shows no feature whatsoever at rms.
Because it is primarily used for comparisons to equilibration rates, it is most convenient
to define the Eulerian inflow time in terms of the mean inflow rate,
in i = hhV riρi/r = (cid:28)−R dθ dφ√−gρU rα/W
rR dθ dφ√−gρ (cid:29) .
ht−1
(6)
Figure 3b shows the radial dependence of this inflow time relative to the orbital period. In
all cases but the most rapidly spinning black hole, this ratio rises gradually from ≃ 2 near
r = rms to ≃ 10 -- 15 at large radii. The a/M = 0 and 0.5 models are nearly identical; the
a/M = 0.9 model shows a slightly flatter curve and a larger value inside rms. The behavior
of this ratio for the a/M = 0.998 model sharply contrasts with that of all the others: it is
10 -- 20 at all radii.
For a Lagrangian interpretation, one can define the proper (i.e., fluid frame) radial infall
time from a given radius to the inner radial boundary by integrating
τin = Z r
rh kdsk
(7)
with ds2 = gtt dt2 + grr dr2 using the time-averaged inflow velocity. Doing so produces curves
that are very similar to those in Figure 3b, except that the inflow time is reduced by a factor
∼ 3 -- 4. This contrast between the Eulerian and Lagrangian definitions is due to the sharp
fall in the ratio of inflow time to orbital period at smaller radii. The inflow velocity scales
-- 9 --
roughly ∝ r−3 for the a/M = 0 case, and is only slightly less steep when a/M = 0.5 or 0.9.
Because of this rapid increase in velocity near the hole, most of the Lagrangian inflow time
is spent at large radius.
The Lagrangian inflow time provides one estimate for the location of the turbulence edge,
the point where the magnetic field dynamics switch from being controlled by turbulence to
being controlled by flux freezing in a plunging infall. Krolik & Hawley (2002) examined the
ratio of infall to orbital time near the location of the turbulence edge, and found that it lies
near the point where the (Lagrangian) infall time becomes shorter than the dynamical time.
This occurs at ≃ 1.3 -- 1.4rms for all three models with a/M ≤ 0.9. When the black hole spins
very rapidly, however, this ratio is almost constant as a function of radius, not betraying a
clear location for the turbulence edge at all.
Indeed, a striking aspect of Figure 3b is the contrast between the three more slowly-
spinning cases and the most rapidly-spinning one. The slow-spin cases behave as one would
expect: well outside the marginally stable orbit, inflow takes multiple orbits because it is
limited by the slow loss of angular momentum due to magnetic torques. As one approaches
the location of the marginally stable orbit, however, there is less to prevent inward radial
motion at constant angular momentum because the effective potential is very nearly flat.
There, the ratio of infall time to orbital period drops toward one.
The a/M = 0.998 case is very different. This contrast presents a puzzle because the
effective potential seen by its material is not qualitatively different from the others. As was
discussed in Paper I (see Fig. 9), the specific angular momentum distribution in the accretion
disk is close to Keplerian in all four simulations.
It is therefore difficult on the basis of
gravitational orbital dynamics alone to understand why the ratio of inflow to orbital time
remains large even at r = rms when a/M = 0.998, and why the transition from turbulence
to plunging inflow is inhibited. We believe that this result has implications for the impact
of electromagnetic stresses on the angular momentum budget. This will be discussed in § 4.
3.3. Radial distribution of surface density
One definition of an accretion disk's inner edge is a rapid decline in the disk's surface
density distribution. Figure 4 is the time-averaged rest-mass surface density distribution
as a function of radius for all four simulations. The mean surface density for each of the
simulations is normalized to the initial torus mass for that simulation. In addition, in order
to avoid creating a misleading visual impression, the radius in this plot is proper radius,
not Boyer-Lindquist coordinate radius; the relation between this radial variable and Boyer-
-- 10 --
Lindquist radius is defined as
R = rin +Z r
rin
dr′pgrr(r′),
(8)
where the metric coefficient is evaluated in the equatorial plane. We derive the surface den-
sity from the history data, which provides the total mass in each radial shell for every M
in time. To obtain a surface density we divide the shell mass by an equatorial unit area
and average over time for t > 2000 M. We restrict radial coverage to regions within 15M
in order to consider only those places in approximate inflow equilibrium; for this reason,
when plotted relative to Rms, the curves extend for different distances. All these curves
may be compared with one illustrating a conventional disk model (triple dot-dashed line).
This curve was constructed combining the run of stress predicted by the (zero-stress bound-
ary condition) Novikov-Thorne model and the fixed ratio of integrated stress to integrated
pressure suggested by Shakura & Sunyaev (1973).
Several points stand out in Figure 4. All four simulations show quite similar behavior,
with Σ rising steadily with increasing radius, steeply within the plunging region and more
gradually in the disk proper. As was emphasized by Krolik & Hawley (2002), there is no
sharp break in surface density, not even at the marginally stable radius. Normalized to the
initial mass, the surface density in the marginally stable region declines by about a factor
of two as a/M increases from 0 to 0.9, but then rises again when a/M = 0.998. In sharp
contrast to this behavior, the conventional model predicts a well-defined maximum in Σ
at 2 -- 3Rms, with a drastic drop-off at and inside Rms and a more gradual decline at larger
radius. The reason for the very sharp break in the standard model is that the net radial
accretion velocity is much smaller than either the orbital velocity or sound speed outside
Rms, but is assumed to abruptly rise to the free-fall velocity inside Rms. The result is an
abrupt drop in surface density. As can be seen in Figure 3a, the accretion velocity in our
simulations increases smoothly and continuously from outside Rms inward toward the hole.
The surface density distribution is important in locating the reflection edge of the disk,
the point within which the disk can no longer efficiently Compton scatter X-rays (Krolik
& Hawley 2002). Compton reflection is a process whereby dense, cool matter in the disk
reflects hard X-rays generated somewhere outside the disk, imprinting upon their spectrum
the signature of absorption by heavy element K-edges. Efficient reflection requires a disk
column density large enough to make it thick to Compton scattering. Because the column
density must decrease sharply with diminishing radius in the plunging region, the reflection
edge is generally assumed to be located at or near rms.
Although the simulations have no absolute density scale, the results may be phrased
in terms of an absolute density once the accretion rate in Eddington units is specified: the
-- 11 --
relation is Σ = 38 m g cm−2, where the accretion rate relative to Eddington m is defined
assuming unit radiative efficiency in rest-mass units ( MEdd = 1.7 × 1017 mM/M⊙ g s−1).
A maximum value for m in these units would be ≃ 20 (if the efficiency is as low as 0.05
and the luminosity is exactly Eddington), and in most instances one would expect it be
a good deal less. The surface density in code units for the three slower-spinning simula-
tions near rms is ≃ 0.01, while it is ≃ 0.06 for the a/M = 0.998 simulation. Translating
to physical units, we predict Compton optical depths τT ≃ 1.3( m/10)(Σ/0.01 code units).
Thus, rapidly-accreting systems may indeed find their reflection edges near rms, but in more
slowly-accreting systems it may be significantly farther out. And, of course, in all cases,
the location of this edge can vary significantly in time. Because disk thermodynamics can
influence the radial distribution of surface density, it may be too early to say definitively
how the effects we are examining alter classical expectations, but it is clear that they can
strongly affect this basic property of disks.
To summarize the results of this section, we find that the low spin models with a/M ≤
0.9 develop accretion disks that are in a quasi-steady state, but the highest spin model,
a/M = 0.998 does not. For fixed surface density at large radius, the accretion rate into the
hole decreases with increasing a/M, although the accretion rates feeding mass into the region
inside the initial torus are comparable. This means that the mass of the inner torus tends
to be greater for larger black hole spin. Accretion infall times, too, are longer with larger
spin. All these results suggest that the spin of the black hole affects the disk in ways that go
beyond simply determining the location of the marginally stable orbit. The mechanism by
which the black hole does so is elucidated by the next topic, the angular momentum fluxes
in the disk.
4. Angular Momentum Transport
The stress tensor
T µ
ν = ρhU µUν − bµbν + b2U µUν + (p + b2/2)gµ
ν
(9)
describes the flux of four-momentum. Conservation of four-momentum means that ∇µT µ
ν =
0, where the ∇ indicates covariant differentiation. The four terms contributing to the stress
tensor each have simple interpretations: the first is the flux directly associated with matter;
the second is the purely electromagnetic part due to correlations in the magnetic field; the
third is the part proportional to the magnetic energy advected with the flow; the fourth is
the momentum flux associated with pressure. In the context of accretion disks, the most
interesting component of this tensor is T r
φ, the angular momentum flux in the radial direc-
tion. The second term then represents the magnetic torque that moves angular momentum
-- 12 --
outward, while the first and third represent the angular momentum moving inward with the
accretion flow, and the fourth is zero. Thus, the magnetic torque contribution −brbφ has
particular importance for accretion disks because it is responsible for transferring angular
momentum outward from one fluid element to another. The net flux (the sum of all the
terms) is the conserved angular momentum flux through the disk.
4.1. Shell-integrated angular momentum flux
To understand better the flow of angular momentum in the disk, it is helpful to look at
the spherical shell-integrated time-average of each of the three terms in equation 9 separately
as a function of radius (Fig. 5). Each of the three terms is integrated over the angular
coordinates, but we restrict the integration to the region where the flow is bound, i.e.,
−hUt < 1. This removes the contributions from the magnetic fields in the funnel which, as
discussed in De Villiers et al. (2004), are not insignificant when the black hole is rotating.
The computation is done using complete data sets spaced every 80M in time; these are then
averaged over the last 75% of the simulation.
Because ∇µT µ
ν , is constant in a steady-state
disk. Separating the terms of the r-φ component of the stress tensor and employing Gauss's
Theorem transforms this statement into
ν = 0, the integral over a radial shell, R dΩ T r
Mh(h + b2/ρ)Uφi −Z dΩ brbφ = const.
(10)
In conventional disk models it is assumed that the stress term in equation (10) is zero inside
some radius that lies well outside the event horizon. Generally this "stress edge" is assumed
to lie at the marginally stable orbit, or, in more detailed hydrodynamical models, it is taken
to be the sonic point, which is generally a short way inside rms (e.g., Abramowicz et al. 1988).
However, as Page & Thorne (1974) remarked, and Krolik & Hawley (2002) discussed in the
context of pseudo-Newtonian models, when magnetic fields are important, the stress edge
need not coincide with rms. We will shortly examine where it may be found in our general
relativistic simulations.
Time-steadiness also implies that the accretion rate is constant as a function of radius,
so equation (10) requires the shell-integrated stress to increase outward to match the angular
momentum per unit mass, which is ∝ r1/2 in the Newtonian limit. Physically, this means that
for matter to move inward it must lose angular momentum by transferring the appropriate
amount of angular momentum outward. A steady state at a fixed position is maintained
when that transfer is balanced by matter flowing in carrying angular momentum.
-- 13 --
We can directly compare this picture to the simulation results. In Figure 5 we see that
in the low spin cases, a/M = 0.0 and a/M = 0.5, some of this picture is reproduced in a
time-averaged sense. The shell-integrated stress rises outside the marginally stable orbit,
accompanied by a compensating gradient in the matter flux. However, rather than going to
zero inside rms, the shell-integrated stress, while smaller than in the main disk body, stays
at a constant level.
In the two higher spin models the departures from conventional expectations are even
greater. Just as in the low-spin cases, stress continues throughout the plunging region, but
its magnitude (there and in the disk proper) is everywhere proportionally larger. In rough
terms, its magnitude is comparable to the angular momentum flux carried with the matter.
We conclude, therefore, that, as viewed in the coordinate frame, there is no stress edge;
stress continues all the way from the disk body to the event horizon.
A second, even more dramatic, contrast with conventional views is that in the inner
regions the gradient in the shell-integrated stress has the wrong sign: the magnetic angular
momentum flow decreases outward. Where this is the case, rotationally-supported matter
does not on average lose angular momentum by magnetic torques. This change in sign of the
gradient of the angular momentum flow appears to be the reason why the rate of accretion
onto the black hole is smaller in the two simulations with rapidly-spinning black holes, and
likewise why the mass of the inner torus becomes large in these two cases.
4.2. Mean specific angular momentum
Additional information can be obtained through an analysis of the mean accreted angu-
lar momentum per unit rest-mass, j. Again with the integrals restricted to the bound matter,
φii/h Mi. We plot the function j(r) in Figure 6. Here the solid
this is defined as j(r) = hhT r
line shows the specific angular momentum associated with the total angular momentum flux,
while the dashed line is that of the matter flux alone. The a/M = 0 and a/M = 0.5 simu-
lations behave very much in line with expectations; j is very close to constant with radius
through the disk, indicating that the angular momentum flux has equilibrated to an approx-
imately time-steady state throughout these regions. Although some angular momentum is
carried in the jet when a/M = 0.5, it is too little to have much effect on the total. In each
model the dotted line indicates the value j(rms) associated with the marginally stable orbit,
and it is clear that jin < j(rms), where jin is the value carried into the black hole. This
demonstrates that torques inside the marginally stable orbit do remove some angular mo-
mentum from material flowing through that region, but the effect is not dramatically large:
it ranges from 6% to 8%.
-- 14 --
In the two simulations with the more rapidly-spinning black holes, the radial slope in
j(r) provides evidence that they are not in steady state, particularly the a/M = 0.998
model. Again the specific angular momentum carried by the mass does not differ much
from the value j(r). However, in these cases there is a significant difference between the net
angular momentum carried by the bound material and the total. The angular momentum
flux associated with the jet is ∼ 10% of that carried in bound matter when a/M = 0.9, but
is comparable to the flux through the bound matter when a/M = 0.998. When all is said
and done, j is 12% below j(rms) for the a/M = 0.9 model, and 42% below for a/M = 0.998.
If the conditions of accretion remain constant long enough for the accreted mass to
dominate the original mass in the black hole, j determines a/M. As noted in Paper I, with
jin = 0.8 in the a/M = 0.998 simulation, in the long run the spin of the black hole can not
be maintained (see also Gammie, Shapiro, & McKinney 2004).
4.3. Angular distribution of angular momentum flux
We can see more precisely where the angular momentum goes by studying the angular
distribution of the magnetic torque contribution to the stress, T r
In Figure 7 the
dashed lines are the (inbound) matter component, plotted at 2rms and rms, and the solid
lines are the (outbound) total electromagnetic component, plotted at the same locations as
the matter component and at the inner boundary as well.
φ (r, θ).
The plot for the Schwarzschild model is straightforward. The matter angular momen-
tum flux narrows going from 2rms to rms, while the magnetic stress peaks through the disk
and drops off through the corona. In the corona there is a region where the outward elec-
tromagnetic stress exceeds the inbound matter flux. There is negligible electromagnetic
angular momentum flux at the inner boundary, and what there is is ingoing, represented by
a dot-dashed line.
In contrast, all the non-zero spin simulations have a significant outward angular mo-
mentum flux carried by magnetic fields throughout the region between the marginally stable
orbit and the inner boundary. Moreover, as was also shown in Figure 5, the magnitude of
this flux relative to the stress associated directly with matter (ρhU rUφ, the dashed curve
in the figure) increases steadily with increasing a/M. As a/M increases, the angular width
through which the magnetic angular momentum flux travels both spreads and develops a
double-maximum structure. That is, when the black hole spins slowly, the magnetic angular
momentum flux is largely confined to the region where the matter flux is greatest, a span of
perhaps ±30◦ on either side of the equatorial plane. Moving from the disk to the corona,
-- 15 --
there is a point where the magnitude of the magnetic stress exceeds that of the matter stress.
When the black hole spins rapidly, the magnetic angular momentum flux is relatively con-
stant with polar angle through the disk region, but rises to a maximum at the funnel wall
and slowly declines in the funnel interior toward the axis.
In the Schwarzschild simulation the magnetic stress peaks within the disk, and falls as
the flow approaches the hole, becoming inward-directed at the inner boundary. In the a/M =
0.5 simulation the magnetic stress can be described in terms of two pieces: a contribution
spanning a broad angular range that increases slowly with increasing radius, and another
contribution whose angular extent roughly matches that of the disk matter and grows toward
larger radii. As the black hole spins faster, the relative importance of the broad angular range
component grows, and a double-peak structure develops, with the greatest stress occurring
near the funnel wall.
We can understand these trends in terms of a changing balance between effects arising
from the magnetorotational instability in the disk and effects stemming from magnetic cou-
pling to the spin of the black hole itself. In the disk, the stress results from MHD turbulence
generated by the MRI (Balbus & Hawley 1998), and the magnetic stress is tied directly to
the presence of the matter. The narrow disk component can be identified with this process;
M Uφ, which increases with
it increases with radius because its gradient must match that of
radius. The broader component we associate with frame-dragging acting on the magnetic
field in the corona and outflow. Close to the black hole, the magnetic field intensity varies
little with angle, so there can be stress across a wide range of polar angles. Its magnitude
changes little with radius because the matter density is so small in the corona and outflow
that there can be little angular momentum exchange between fields and matter.
In rough terms, it is possible to separate the electromagnetic angular momentum flux
going into the disk from that going into the corona by evaluating this flux on the shell at
r = rms and distinguishing disk from corona in terms of polar angle. Applying this criterion
quantifies some of the remarks of the previous paragraphs: as a/M increases, the ratio of the
electromagnetic angular momentum flux entering the corona to that entering the disk rises
from tiny (< 10−3 in the Schwarzschild case) to ≃ 0.3 (a/M = 0.5), ≃ 1 (a/M = 0.9) and
≃ 3 (a/M = 0.998). Because angular momentum can readily move from matter to fields and
back again, and because polar angle distribution at a fixed radius does not fully distinguish
the ultimate end of the flux, these numbers are not entirely well-defined; nonetheless, they
do express the qualitative point that the corona receives a growing proportion as the black
hole spins faster.
When the black hole spins rapidly, the overall level of electromagnetic angular momen-
tum flux is so high that, despite the growing share devoted to the corona, the portion going
-- 16 --
into the disk is large enough that it fundamentally alters the angular momentum budget of
the accretion flow. It appears that in both the a/M = 0.9 and a/M = 0.998 simulations, the
time-average magnetic torque adds to the matter's angular momentum over a substantial
radial range from near the hole to rms and beyond (Fig. 5). As a result of this process, the
accretion flow is retarded. In Paper I, we observed that in each simulation an inner torus is
created in the region immediately outside rms, and that the prominence of this inner torus
increases with increasing a/M. The electromagnetic stress driven by the spin of the black
hole is the likely origin of this dependence on a/M. In the case of the a/M = 0.998 sim-
ulation, magnetic torques are so strong that the mean of the local angular momentum can
exceed the value associated with a circular orbit. For example, the density-weighted ratio of
local angular momentum to the angular momentum of a circular orbit at that radius rises
from ≃ 0.95 to ≃ 1.05 in the inner torus region (see also Fig. 13 of Paper I). The effective
potential for the flow is then not as flat as expected in the marginally stable region, and
inflow cannot occur as easily (see also Fig. 3).
4.4. Fluid frame stresses
It is also useful to look at the stresses in the fluid frame. The MHD turbulence is
primarily due to local instabilities, so its physics is best analyzed in the local fluid frame. In
addition, Novikov & Thorne (1973) showed that, if the disk is assumed to be time-steady,
axisymmetric, and geometrically thin, the vertically-integrated stress in the fluid frame can
be written in the physically-enlightening form
M Ω
2π
RT (r/M, a/M),
φ (r) =
(11)
W r
where the function RT encapsulates both relativistic corrections and the net angular momen-
tum flux (the notation here follows Krolik 1999). The assumption that stress goes to zero
at the marginally stable orbit means that RT (r) = 0 at all radii r ≤ rms. In the Newtonian
limit of large r, RT → 1. In this relation, "stress" means the part of the stress tensor other
φ − M Uφ(r), where Uφ(r) is the
than that directly associated with accretion, i.e., W r
circular orbit angular momentum at r.
φ = T r
To compare our results with the Novikov-Thorne prediction, we compute a closely-
related quantity,
(φ)(r) ≡ R dθ dφ √−g [b2U µUν − bµbν] e(r)
W (r)
R dφp−g(θ = π/2)/gθθ(θ = π/2)
µ eν
(φ)
,
(12)
where the eµ
(ν) are an orthonormal tetrad designed to represent the directions labeled by (ν) in
the fluid frame (see Appendix for details about their construction). We limit the integration
-- 17 --
(φ)
to the bound portion of the flow, where −hUt < 1. We divide the shell-integrated fluid-frame
stress by the area per unit radial coordinate in the equatorial plane in order to give this W (r)
the same units as W r
φ.
We compute W (r)
(φ)(r) at t = 8000 M for all four simulations; this is shown in Figure 8.
To facilitate comparison of simulations with different black hole spins, the value of W (r)
(φ) dis-
played in each panel is normalized to the initial maximum pressure found in that simulation.
There is rough agreement with the Novikov-Thorne prediction (assuming the time-average
accretion rate at the inner boundary) for 1.5rms . r . 40M. At larger radii, one would not
expect agreement because that is the region in which the matter in our simulations must
move outward to absorb the angular momentum flux, whereas in the Novikov-Thorne model
that region is supplied by accreting matter from farther away. Inside rms, however, where
the Novikov-Thorne model predicts zero stress, our data are dramatically different. We find
that the stress continues quite smoothly, and, in fact, continues to increase in amplitude
from rms to the inner boundary whenever a/M > 0. Only in the Schwarzschild case does
the fluid-frame stress eventually go to zero, albeit only very close to the event horizon.
In other words, there is considerable outward angular momentum flux in the fluid frame
virtually throughout the flow, including the plunging region. Moreover, whenever a/M > 0,
there is always at least some outward angular momentum flux in the fluid frame even at the
smallest radii outside the event horizon. Further, even with the overall normalization, the
general level of the fluid-frame magnetic stress rises substantially with increasing a/M. This
is due to the increasing amplitude of the magnetic fields. Thus, while there is a fluid-frame
stress edge in the Schwarzschild case just outside the event horizon, there is no fluid-frame
stress edge when the black hole spins.
The standard Novikov-Thorne model assumes a cold, unmagnetized disk, with h =
b2 = 0. The origin of the stress is unspecified, although it is presumed due to turbulent
motions. In our simulations neither h nor b2 is zero. In the calculation of Figure 8 we
included only the magnetic stress and neglected the turbulent Reynolds stress, i.e., ρhU rδUφ,
where the perturbation is with respect to a local circular orbit. Local simulations (Hawley,
Balbus & Winters 1999) show that the Reynolds stress is generally a factor of several less
than those due to the magnetic fields. Obviously the presence of magnetic fields explains the
qualitatively different behavior of W (r)
φ from the marginally stable region inward.
(φ) and W r
Another potential influence on the stress edge is the temperature in the disk and the
associated vertical scale height. Purely hydrodynamical models of accretion disks more de-
tailed than the Novikov-Thorne model (e.g., Muchotrzeb & Paczy´nski 1982; Matsumoto et
al. 1984) relax the assumption of h = 0 and consider the effects of finite geometric disk thick-
ness. These studies indicate that, although the inner boundary for stress may not coincide
-- 18 --
exactly with the marginally stable orbit, in thin disks it should lie very close. Even in "slim
disks" (Abramowicz et al. 1988), the stress does not extend all the way across the plunging
region: in the pseudo-Newtonian approximation, designed to mock up Schwarzschild metric
dynamics, the innermost radius for the end of stress stretches inward from 6M to 4M as the
disk thickens. But it is clear that the purely hydrodynamic picture is inadequate; magnetic
stresses operate in ways that Reynolds stresses and simple viscosity cannot. Further, in
the MHD simulations there is significant magnetic field in the surrounding disk corona; the
magnetic scale height is greater than the gas pressure scale height of the disk. In fact, the
magnetic scale height near the black hole is ∼ r more or less independent of the matter scale
height in the disk. Thus, when magnetic fields are included, the picture changes qualitatively
from the purely hydrodynamic predictions.
Earlier we found that, when stress is measured in the coordinate frame, there is no stress
edge anywhere. When it is evaluated in the fluid frame, the situation changes only slightly:
when a/M > 0, there is similarly no stress edge; when a/M = 0, one exists, but it is deep
in the plunging region, immediately outside the event horizon.
5. Energy transport
5.1. Energy fluxes
The conventional approach to defining the radiative efficiency of accretion onto black
holes was set out by Novikov & Thorne (1973). In the standard picture, thermal energy
in the disk is assumed negligible compared to orbital energy, in part because the thermal
loss time is assumed to be very short compared to the inflow time everywhere outside rms.
With those assumptions, the radiation rate in each ring is simply the difference between the
energy brought into that ring by mechanical work (torques acting on the ring by neighboring
zones) and energy lost from that ring by accretion. Both contributions amount, of course,
to the radial divergence of an energy flux which does not vanish; the energy deposited by
the flux is the disk's luminosity. As the fluid drops toward the hole, it becomes increasingly
bound, with a binding energy equal to that of the circular orbit at its radius. The energy
of the last circular orbit at rms sets the canonical efficiency of standard thin disk accretion,
η = 1 + Ut(rms). For a Schwarzschild hole this is 0.057.
In the simulations, things are less straightforward. As in the standard picture, energy
is extracted from orbital motion as the gas accretes, but that energy is not radiated. It first
goes into magnetic fields and velocity fluctuations within the turbulent disk. Some magnetic
field rises out of the disk into the corona, and a general coronal outflow removes some energy
-- 19 --
from the disk. As the turbulence in the disk is dissipated at the grid scale, some of the
energy is converted into heat by artificial viscosity; some is lost. Enough of this energy is
retained to keep the disk hot.
We can investigate how energy is transported in these simulations by considering the
t component of the stress tensor. Just as for
magnitude of the energy flux, given by the T r
the angular momentum flux, it has both a fluid part
and an electromagnetic part
(T r
t )fluid = ρhU rUt
(T r
t )EM = b2U rUt − brbt
= F µrFµt
= grr(cid:0)gφφE φBθ − gθθE θBφ − gtφE rE φ(cid:1) ,
(13)
(14)
where E r = V φBθ − V θBφ, E θ = V rBφ − V φBr, and E φ = V θBr − V rBθ. The last form for
the electromagnetic energy flux shows that it is the radial component of the Poynting flux,
modified by a frame-dragging correction.
In a steady-state disk, T r
t integrated over spherical shells should be constant with radius.
However, because its largest contribution is from the flow of rest-mass, its radial behavior
is dominated by that flux, a conserved quantity to which we have already devoted much
attention (§3.1). In order to focus attention on the non-rest-mass portion, we instead study
t − M = hhρ(hUt + 1)U r + b2U rUt − brbtii. In a time-
shell-integrals of time-averaged T r
M with radius means that this "adjusted" energy flux should
steady disk, the constancy of
likewise be constant as a function of r. In an effort to make still closer contact with ordinary
t ii/h Mi, the binding energy per unit rest-mass accreted,
intuition, in Figure 9 we plot 1−hhT r
but restricting the integral to cells in which the matter is bound. In this figure, the dashed
line is the specific energy carried by the fluid term alone (eq. 13), and the solid line is
obtained from both the fluid and electromagnetic components. As references, there are two
dotted lines: the straight one shows the binding energy of the marginally stable orbit, the
curved one the binding energy of a circular orbit in the equatorial plane at that radius.
If all the simulations achieved a steady state within the region shown, and if there were
no energy exchange between bound matter and anything else, the solid line would be flat.
As is immediately apparent, it is not. At large radii, the fluid binding energy flux (dashed
line) closely tracks the orbital energy curve. This is because the numerical energy loss time
is shorter than the inflow time, crudely mimicking the effects of radiation losses in a real
accretion disk. At smaller radii, the fluid binding energy is smaller than the orbital binding
energy for two reasons: artificial viscosity adds heat to the gas, and the inflow time becomes
-- 20 --
shorter than the numerical loss time. Some energy is lost to the outflow, especially when
a/M ≥ 0.9, but not enough to strongly cool the disk. Here the mimicry is closer in spirit to
a non-radiative disk.
5.2. Poynting flux
At all radii there is a substantial offset between the total energy flux curve and the
fluid-only curve. This offset is due to the outward energy flux conveyed electromagnetically
(Fig. 10). With the exception of the inner boundary in the zero-spin simulation (where it is
also very small in magnitude), the Poynting flux is always directed outward. Independent
of the black hole spin, there is always Poynting flux associated with the work done by
torques in the disk. In the outer portions of the accreting disk (r ≃ 15M), where our initial
conditions make all four simulations very similar, its magnitude varies little with black hole
spin. However, in the inner disk, where the mean magnetic field strength grows substantially
with more rapid rotation, so does the Poynting flux. At the equator on the inner boundary,
it increases by about a factor of two from a/M = 0.5 to a/M = 0.9 and by almost an order
of magnitude from a/M = 0.9 to a/M = 0.998.
Poynting flux in the jet outflow varies far more dramatically with black hole spin. It is
negligible in the Schwarzschild simulation, while in the a/M = 0.998 simulation the Poynting
flux in the funnel wall completely dominates the Poynting flux in the disk. Another way to
make this comparison is to consider the ratio of the total (outward) shell-integrated Poynting
flux to the (inward) fluid energy flux (nearly all rest-mass energy). This ratio is plotted in
Figure 11. Here we see that relative to the rate at which fluid energy is accreted, the Poynting
flux integrated over the inner boundary rises from 0.25% when a/M = 0.5 to ∼ 1% when
a/M = 0.9 to ∼ 10% when a/M = 0.998. These ratios increase steadily outward for the
a/M = 0 and 0.5 simulations, and have a local maximum around r = 3 -- 5 M for the high spin
models. In the simulation with the most rapid spin at the radius where this ratio reaches its
maximum, Poynting flux carries outward ∼ 60% of the ingoing fluid energy flux.
Another way to look at these results is from the point of view of the Penrose process.
In its simplest form, this process depends on the fact that inside ergospheres there can
exist negative energy orbits, so that black holes can give up energy (i.e., lose mass) by
absorbing particles with negative energy. An analogous process exists for magnetic fields.
The electromagnetic "energy density at infinity" may be defined (Koide 2003) by
EM/α = −T t
e∞
t (EM) = − (1/2)gtt(cid:2)grr(E r)2 + gθθ(E θ)2 + gφφ(E φ)2(cid:3) + (1/2)(cid:0)gtφE φ(cid:1)2
+(cid:2)gθθgφφ(Br)2 + grrgφφ(Bθ)2 + grrgθθ(Bφ)2(cid:3) .
(15)
-- 21 --
(16)
¿From this form we see that, if all the spatial elements of the inverse metric were positive
EM could never be negative. However, inside the ergosphere, gφφ < 0. Thus,
definite, e∞
negative electromagnetic energy-at-infinity is associated with regions inside the ergosphere
where the components of the field tensor associated with poloidal magnetic field (Br and Bθ)
and toroidal electric field (E φ) are large in magnitude compared to the other field components.
Moreover, because E φ = V θBr − V rBθ, a large poloidal magnetic field also helps to make E φ
large, particularly when the poloidal transport speed is substantial. That an electromagnetic
version of the Penrose process is at work is suggested by the fact that in the a/M = 0.9 and
EM < 0 in the region within 40◦ -- 60◦ of the equatorial
0.998 simulations, the time-averaged e∞
plane and ≃ 0.3M of the inner boundary.
One of the key questions of black hole accretion is whether or not energy can be extracted
from the region inside rms. Although it is difficult to say how much radiation will escape
from inflowing gas within the plunging region (discussed below), it is clear that the Poynting
flux provides an avenue for the outward transport of energy from the plunging region and
from the black hole itself.
5.3. Radiation
Although the present simulations do not include radiative losses, it is interesting never-
theless to consider whether the standard assumptions of thin disk theory would apply. The
issues involve where and how rapidly energy is dissipated into heat, and whether there is
sufficient time for the heat to be transformed into photons and radiated prior to the gas
entering the black hole. Well outside the marginally stable orbit, the time required for these
processes to occur should be significantly shorter than the inflow time. Where that is true,
the conservation law approach to computing the radiation rate should work well. In the sim-
ulations the turbulent energy released by dissipation is captured only partially by artificial
viscosity; numerical losses thus mimic radiation losses, and in most of the main disk body,
the numerical loss time is appropriately shorter than tin. The only difference between the
simulations and the Novikov-Thorne model in this regard is that we do not force the stress
to go to zero at rms. A formalism for tracing the consequences of that changed boundary
condition in time-steady disks was developed in Agol & Krolik (2000). Applying it to the
simulation results indicates an enhanced radiative efficiency from this part of the disk that
is relatively modest for the cases with a/M = 0, 0.5, and 0.9. Because the a/M = 0.998
simulation did not achieve an inflow steady-state, the assumptions of this formalism may
not be valid, but the data suggest that the enhancement to the radiative efficiency may be
-- 22 --
considerably larger.
Inside ≃ 2rms, the inflow time falls rapidly (see Fig. 3), becoming roughly an orbital
period near and inside rms. It then becomes questionable whether the several processes that
stand between inter-ring work and radiation have enough time to run to completion. If not,
the gas advects its energy inward, much as envisioned by Rees et al. (1982), Narayan & Yi
(1995), and Abramowicz et al. (1988), albeit for different reasons. However, particularly
in the plunging region proper, new processes can lead to radiation. Machida & Matsumoto
(2003) suggest that field reconnection in the plunging region could lead to X-ray flares.
Rather than magnetic field being dissipated on very small lengthscales at the end of a
turbulent cascade, it can instead by dissipated in large-scale current sheets. In Hirose et al.
(2004) we drew attention to the prevalence of high current density regions in the inner disk
and plunging region where the current density is so high that rapid field dissipation might
be expected. If the local temperature in these current sheets can be raised high enough that
Compton scattering is the dominant cooling mechanism, radiation can release energy very
rapidly: the Compton cooling time is
tC =
3
4
me
µe (cid:18) L
LE(cid:19)−1
(cid:16) r
M(cid:17)3
M,
(17)
where µe ∼ mp is the mass per electron and L is the luminosity of seed photons injected
within radius r. Some of these photons will be captured by the black hole, and some will
strike the disk farther out, but some portion will also reach distant observers. To compute
the actual radiative efficiency of flows of this sort will require calculations that explicitly
balance the various competing processes and timescales.
6. Discussion and Summary
In Paper I of this series we reported that the fraction of the initial mass accreted in
a fixed time diminishes sharply with increasing black hole spin. We also noted that the
angular momentum per unit rest-mass that is accreted likewise drops as the black hole spins
more rapidly, and by significantly more than the associated decrease in the specific angular
momentum of the innermost stable circular orbit. On the basis of the more detailed analysis
presented here, we can now explain both these effects. Not surprisingly, they are closely
related.
An examination of the angular momentum fluxes in the accretion flow reveals how
significant magnetic fields are to the evolution of the inner region of a black hole accretion
disk. The low a/M models attain a quasi-steady state that behaves much as one would
-- 23 --
expect from simple accretion disk theory. The net flux of angular momentum through the
disk is essentially equal to the value carried into the black hole. Within the flow itself
outward-going magnetic angular momentum flux balances the excess in the inward-going
component carried by the matter flow.
Contrary to standard accretion theory, however, there are always magnetic stresses
exerted across the location of the marginally stable orbit and in the plunging region interior
to this point. In fact, when the black hole rotates, the r -- φ component of the stress tensor
in the fluid frame rises more steeply with decreasing radius in the plunging region than in
the disk itself. Thus, we have shown that the concerns voiced quietly in Page & Thorne
(1974) and Thorne (1974) have been borne out: the presence of magnetic fields invalidates
the "zero-stress" inner boundary condition.
Hirose et al. (2004) showed that the relative magnitude of the magnetic field in the inner
disk grows steadily with increasing black hole spin. This is a by-product of the creation of
the dense inner torus: higher pressure gas can support stronger magnetic field when the
field's origin lies in the magneto-rotational instability. Thus, as the spin increases, the
magnetic stresses grow rapidly stronger. There is more going on than can be attributed
merely to stronger field strengths, however. The rotating hole itself represents a source of
angular momentum and energy that can be carried outward by the magnetic stress. When
a/M & 0.9, the magnetic stresses carry out of the ergosphere an angular momentum flux that
is comparable to what is brought inward with accreting matter. A striking suppression of
the accretion rate is the result. In the most extreme case (a/M = 0.998), accretion through
the inner disk is retarded to such a degree that even in a simulation time of 8100 M no
accretion equilibrium is achieved; rather, the mass of the inner torus builds steadily through
the simulation as matter accretes from outside, but is held back not far outside rms. This
result is reminiscent of the suggestion by van Putten and Ostriker (2001) that magnetic
torques driven by rotating black holes might lead to a state of suspended accretion, but
the field structure we see (turbulent field, largely confined to the accreting matter) is quite
different from the one posited by them (organized poloidal field with loops spanning all the
way from the disk to the event horizon).
We should note that although the accretion rate over the course of our simulation is
reduced by a factor of 4 contrasting the simulation with a/M = 0.998 to the one with
a/M = 0, the evident lack of equilibrium in the former case prevents us from saying what
will happen in the long run. It is possible that when the inner torus mass grows large enough,
the time-averaged accretion rate will rise. In that case, the effect of black hole spin would
be better described as altering the relation between accretion and inner disk mass rather
than as a simple suppression of accretion. On the other hand, without doing more extended
-- 24 --
simulations, we cannot say whether such an equilibrium is ever achieved -- it is possible that
the structure of the inner disk traverses some complicated limit cycle when the black hole
spins very rapidly. In any case, it is clear that a magnetized spinning black hole can affect
the rate at which matter accretes into it in ways that go beyond the changes in the effective
potential usually considered.
Another consequence of the substantial magnetic torques near the black hole is a re-
duction in the net angular momentum brought to the hole as a result of accretion. As
mentioned briefly in Paper I and at greater length in Gammie et al. (2004), the accreted
angular momentum per unit rest-mass decreases so much with increasing a/M that the black
hole can spin down (in the sense of decreasing a/M) even as M increases. Our results are in
approximate agreement with those of Gammie et al. in indicating that there is near balance
when a/M ≈ 0.9, but likely net spin-down for greater a/M. This result suggests that black
holes cannot be spun-up by accretion to a/M & 0.9; only a black hole merger might create
a more rapidly spinning black hole.
We also examined energy transport in the vicinity of the black hole. As with angular
momentum transport, the combination of magnetic fields and rotating holes leads to phe-
nomena not normally included in standard disk theory. Magnetic fields can extract both
energy and angular momentum from the black hole in a fashion resembling the Blandford
& Znajek (1977) mechanism, but with several notable differences. First, the magnetic field
is far from force-free. Second, the relevant field lines connect the plunging region to the
disk and the outflow rather than only connecting the event horizon to infinity (cf. similar
results found by McKinney & Gammie 2004). Energy is drawn from a rotating black hole in
a manner that, far from being independent of accretion as in the original Blandford-Znajek
picture, is inherently dependent upon accretion. In fact, the most convenient way to describe
its magnitude is relative to the rate at which rest-mass is accreted: for modest spins, the
electromagnetic energy release per unit mass is . 1%, while for very rapid spin, it can be
≃ 10% or more. As the spin increases, the angular distribution of the Poynting flux also
changes, with most going into the disk at low spin but the greater share entering the outflow
when the black hole spins very rapidly.
The simulations have demonstrated several new dynamical effects for magnetically
driven accretion flows in a Kerr metric. At this point we can give only qualitative assessment
of the observational consequences of these effects. For example, strong outward magnetic
angular momentum flux from a spinning black hole leads to a substantial alteration in the
mass surface density distribution, creating a sizable inner torus in the region just outside
rms. Completely unexpected in conventional disk models, this large mass concentration
could significantly alter the expected X-ray reflectivity of this region deep in the relativis-
-- 25 --
tic potential. Because magnetic stresses continue throughout the inner disk and plunging
region, the efficiency with which orbital shear energy can be dissipated into heat and then
escaping photons could be greater than conventionally envisioned, but quantitative results
require an improved treatment of thermodynamics. For rapidly spinning holes, the Poynting
flux represents a very significant part of the energy flow near the black hole. Although the
impact of these effects on disk thermal emission remains to be worked out, their presence
will surely lead to significant effects not found in traditional models.
We would like to thank Charles Gammie, Shinji Koide, Masaaki Takahashi, and Ethan
Vishniac for helpful conversations. JHK would also like to thank the Institute of Astronomy,
Cambridge for their hospitality while this work was completed, and the Raymond and Beverly
Sackler Fund for support during his visit there. This work was supported by NSF grants
AST-0070979 and PHY-0205155, and NASA grant NNG04-GK77G (JFH), and by NSF
grants AST-0205806 and AST-0313031 (JHK and SH). The simulations were carried out by
Jean-Pierre De Villiers on the facilities of the San Diego Supercomputer Center, supported
by the NSF.
Abramowicz, M., Czerny, B., Lasota, J.-P. & Szuszkiewicz, E. 1988, ApJ, 332, 646
REFERENCES
Agol, E., & Krolik, J. H. 2000, ApJ, 507, 304
Balbus, S. A., & Hawley, J. F. 1998, Rev. Mod. Phys., 70, 1
Blandford, R. D. & Znajek, R. 1977, MNRAS 179, 433
De Villiers, J. P. & Hawley, J. F. 2003, ApJ, 589, 458
De Villiers, J. P., Hawley, J. F. & Krolik, J. H. 2003, ApJ, 599, 1238 (Paper I)
De Villiers, J. P., Hawley, J. F. Krolik, J. H. & Hirose, S., 2004, astro-ph/0407092, ApJ,
submitted
Gammie, C. F., Shapiro, S. L. & McKinney, J. C. 2004, ApJ 602, 312
Hawley, J. F., Balbus, S. A., & Winters, W. F. 1999, ApJ, 518, 394
Hirose, S., Krolik, J. H., Hawley, J. F., & De Villiers, J.-P. 2004, ApJ, 606, 1083
Koide, S. 2003, Phys. Rev. D., 67, 104010
-- 26 --
Krolik, J. H. 1999, Active Galactic Nuclei (Princeton: Princeton Univ. Press)
Krolik, J. H. & Hawley, J. F. 2002, ApJ, 573, 754
Machida, M., & Matsumoto, R. 2003, ApJ, 585, 429
Matsumoto, R., Kato, S., Fukue, J. & Okazaki, A.T. 1984, PASJ, 36, 71
McKinney, J. C. & Gammie, C. F. 2004, ApJ, in press
Muchotrzeb, B. & Paczy´nski, B. 1982, Acta Astron., 32, 1
Narayan, R. & Yi, I. 1995, ApJ 444, 231
Novikov, I.D. & Thorne, K.S. 1973, in Black Holes: Les Astres Occlus, eds. C. de Witt & B.
de Witt (New York: Gordon & Breach), 344
Page, D. N. & Thorne, K. S. 1974, ApJ, 191, 499
Rees, M.J., Phinney, E.S., Begelman, M.C. & Blandford, R.D. 1982, Nature 295, 17
Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337
Thorne, K. S. 1974, ApJ, 191, 507
van Putten, M. & Ostriker, E. C. 2001, ApJL, 552, 31L
A. Defining the Fluid Frame Tetrads
In order to compute quantities measured by a physical observer, one must define a set
of basis vectors (commonly called "tetrads") appropriate to that observer, which we write
here as eµ
In this notation, (ν) identifies the direction in the
observer's frame.
(ν) for the contravariant set.
Ideally, these should be directed in four-space so as to be most useful for physical
interpretation. For example, in most instances it is important to be able to identify the
direction of proper time in the observer's frame. This is simply done by setting eµ
(t) = U µ,
the observer's four-velocity. In the observer's own frame, the only non-zero component of the
This preprint was prepared with the AAS LATEX macros v5.2.
-- 27 --
four-vector is the time component, so it automatically satisfies the requirement of appropriate
directionality. Because U µUµ ≡ −1, it is also automatically normalized to unit magnitude.
Identifying other directions is, in general, more ambiguous. One can always construct
the remaining three tetrads by a Gram-Schmidt procedure, requiring the newly-constructed
tetrad element to be orthogonal to all the previously-defined ones and normalized to unit
magnitude. However, there are many different ways in which the tetrad elements can be
ordered for this procedure and the results depend on the ordering. In fact, the results depend
in general on whether one first seeks contravariant or covariant versions. This mathematical
ambiguity emphasizes the physical ambiguity in identifying spatial directions in one frame
with those in another.
For our purposes, we wish to construct a tetrad basis set that both reduces to the
conventional directions in the Newtonian limit and may be most directly compared with
the set chosen by Novikov & Thorne (1973). The choice made by Novikov and Thorne
is a special case that is unambiguous because they were interested only in the limit that
U r = U z = 0 in the coordinate frame. When that is so, tetrad elements that are purely
radial or purely vertical in the coordinate frame are automatically orthogonal to the four-
velocity. Consequently, the R and z tetrad elements can be taken as purely radial and
vertical in the coordinate frame, while the t and φ elements couple only to each other. This
so simplifies the problem that no ambiguities remain.
Although the components of the fluid four-velocities found in our simulations are in
general non-zero everywhere, in much of the simulation region, U φ ≫ U r, U θ. In an effort to
ensure that the direction of greatest difference between the fluid frame and the coordinate
frame is aligned as closely as possible, after setting eµ
(t) = U µ, we begin the Gram-Schmidt
process with the φ element of the tetrad and set its r- and θ-components in the coordinate
frame to zero everywhere. The r-element can then be constructed with a zero θ-component,
but it is also necessary to insist that er
(r) > 0 to preserve the right orientation of the tetrad
set. Finally, only the θ-element remains to be computed; it has non-zero components in all
four coordinates. The result is:
eµ
(t) = (cid:2)U t, U r, U φ, U θ(cid:3)
eµ
(r) = (s/N2)(cid:2)−gttC1 − gtφC2, GrrC0,−gφφC2 − gtφC1, 0(cid:3)
eµ
(φ) = −(1/N1) [l, 0, 1, 0]
eµ
(θ) = (1/N3)(cid:2)1, (C1/C0)(gtt − l2)/grr,−l,
C−/C3 − (U r/U θ)(C1/C0)(gtt − l2)/grr(cid:3) ,
(A1)
(A2)
(A3)
(A4)
-- 28 --
where
l ≡ Uφ/Ut
C0 ≡ U tUt + U φUφ
C1 ≡ U rUt
C2 ≡ U rUφ
C3 ≡ U θUt
C− ≡ U φUφ − U tUt
N1 ≡ (cid:0)gttj2 − 2gtφj + gφφ(cid:1)1/2
2 + 2gtφC1C2(cid:1)1/2
N2 ≡ (cid:0)gttC 2
0 + gφφC 2
N3 ≡ (cid:2)gtt − 2gtφj + (C1/C0)2(gtt − j2)2/grr
s ≡ C0/C0.
1 + grrC 2
+ gφφj2 + gθθ(C−/C3 − (U r/U θ)(C1/C0)(gtt − j2)/grr)2(cid:3)1/2
-- 29 --
Fig. 1. -- The total mass found inside r = 15 M as a function of time, given as fraction of
the initial torus mass. The solid line corresponds to the a/M = 0.0 run, the dotted line to
a/M = 0.5, the dashed line to a/M = 0.9 and the dot-dashed line to a/M = 0.998. In the
three lower spin models the mass in the inner accretion disk remains roughly constant after
t/M ≈ 2000, but the mass continues to grow throughout the simulation for the high spin
model.
-- 30 --
Fig. 2. -- Time-averaged, shell-integrated accretion rate as a function of radius (solid curve),
and rms accretion rate fluctuation about the mean at each radius (dashed curve). The units
of accretion rate are fraction of the initial torus mass accreted per M of time (multiplied by
104). Plots for all four black hole spin models are shown.
Fig. 3. -- (a) Time-averaged, density-weighted inflow velocity hhU riρi as a function of r. The
solid curve is the a/M = 0 simulation, the dotted curve is a/M = 0.5, the dashed curve is
a/M = 0.9, and the dash-dot curve is a/M = 0.998. (b) Ratio of inflow time (as defined
by eqn. 6) to orbital period in the equatorial plane as a function of r/rms. The curves are
identified as in (a).
-- 31 --
Fig. 4. -- Time-averaged surface density Σ as a function of proper radius, R, normalized to
the marginally stable value, out to r = 15M in each of the four simulations. Each curve is
also normalized to the initial torus mass of the simulation. The solid curve is a/M = 0, the
dotted is a/M = 0.5, the dashed is a/M = 0.9, and the dot-dashed is a/M = 0.998. For
comparison, the triple dot-dashed curve is the surface density predicted by a conventional
Novikov-Thorne disk model for a/M = 0.9, assuming a constant ratio of stress to pressure.
The amplitude of the curve is arbitrary; only the shape is relevant.
-- 32 --
Fig. 5. -- Absolute values of the shell-integrated and time-averaged angular momentum
fluxes in the bound portion of the flow for the four simulations. The solid curve is the flux
associated with the matter, the dashed curve is the magnetic torque, the dot-dash curve
is the part due to advected magnetic energy. The sign convention chosen is such that the
matter flux and advected magnetic energy carry prograde angular momentum inward, while
the magnetic torque carries positive angular momentum outward.
-- 33 --
φii/h Mi
Fig. 6. -- The time-average flux of angular momentum per accreted rest-mass, hhT r
for each of the four runs. The data are time-averaged over the last 75% of the simulation.
The dashed line is the angular momentum carried by the matter flux; the solid line is the
total angular momentum flux. For both these curves the space integrals are restricted to
the bound material, i.e., cells with −hUt < 1. The dot-dashed line shows the total angular
momentum flux over the whole domain, including the jet. The values are contrasted with
the angular momentum per rest-mass of the marginally stable orbit j(rms) (dotted line).
-- 34 --
10-1
10-2
s
s
e
r
t
S
a/M = 0.0
rin
rms
10-1
10-2
s
s
e
r
t
S
a/M = 0.5
rin
rms
φ
r
T
10-3
10-4
10-1
10-2
s
s
e
r
t
S
φ
r
T
10-3
10-4
T
rms
rin
2rms
φ
r
10-3
10-4
0
1
2
θ
3
0
1
2rms
rms
rin
θ
2
3
a/M = 0.9
rin
rms
2rms
rms
rin
θ
0
1
10-1
10-2
s
s
e
r
t
S
φ
r
T
10-3
10-4
a/M = 0.998
rms
2rms
rms
rin
rin
1
2
3
θ
2
3
0
Fig. 7. -- The azimuthally-averaged T r
φ stress as a function of polar angle θ, time-averaged
over the last 75% of the simulation. The solid curves are the magnetic contribution to the
outward-directed stress at r = 2rms, r = rms and the inner boundary.
In the a/M = 0
model the magnetic stress is inward-directed (dot-dashed line). The dashed lines are the
mass contribution to the stress, (i.e., ρhU rUφ), at r = rms and the inner boundary.
-- 35 --
Fig. 8. -- The shell-integrated magnetic r-φ stress in the fluid frame per unit area in the
equatorial plane is plotted against radius for each of the four simulations at a selected late
time. The dashed curves show the prediction of the Novikov-Thorne model for the time-
average accretion rate at rin for each of these simulations.
-- 36 --
t ii/h Mi for each of the four simulations.
Fig. 9. -- Specific energy flux, defined as 1 − hhT r
The dashed line is derived using the fluid energy flux only, while the solid line has both the
fluid and electromagnetic energy flux. The dotted line corresponds to the specific energy
of a circular orbit at each radius; the horizontal dotted line is the specific energy of the
marginally stable circular orbit at at rms.
-- 37 --
Fig. 10. -- Time- and azimuthally-averaged outgoing Poynting flux in the four simulations,
plotted on the same linear scale. The values are normalized by the initial energy and mul-
tiplied by 104. As the black hole spins more rapidly, the Poynting flux associated with the
funnel-wall outflow rapidly grows in importance.
-- 38 --
Fig. 11. -- Ratio of time-averaged electromagnetic energy flux, hhT r
t iiEM to fluid energy
flux, hhT r
t iif luid, as a function of radius. The solid line corresponds to the a/M = 0.0 run,
the dotted line to a/M = 0.5, the dashed line to a/M = 0.9 and the dot-dashed line to
a/M = 0.998. As the spin of the black hole increases, an increasing fraction of the energy
flux is in the form of Poynting flux.
|
astro-ph/9804256 | 1 | 9804 | 1998-04-24T08:26:04 | ASCA observations of two steep soft X-ray quasars | [
"astro-ph"
] | Steep soft X-ray (0.1-2 keV) quasars share several unusual properties: narrow Balmer lines, strong FeII emission, large and fast X-ray variability, rather steep 2-10 keV spectrum. These intriguing objects have been suggested to be the analogs of Galactic black hole candidates in the high, soft state. We present here results from ASCA observations for two of these quasars: NAB0205+024 and PG1244+026. Both objects show similar variations (factor of about 2 in 10 ks), despite a factor of about ten difference in the 0.5-10 keV luminosity (7.3E43 erg/s for PG1244+026 and 6.4E44 erg/s for NAB0205+024, assuming isotropic emission, H_0 = 50.0 and q_0 = 0.0). The X-ray continuum of the two quasars flattens by 0.5-1 going from the 0.1-2 keV band toward higher energies, strengthening recent results on another half dozen steep soft X-ray AGN. PG1244+026 shows a significant feature in the `1 keV' region, which can be described by either as a broad emission line centered at 0.95 keV (quasar frame) or as edge or line absorption at 1.17 (1.22) keV. The line emission could be due to reflection from an highly ionized accretion disk, in line with the view that steep soft X-ray quasars are emitting close to the Eddington luminosity. Photoelectric edge absorption or resonant line absorption could be produced by gas outflowing at a large velocity (0.3-0.6 c). | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 14 May 2018
(MN LATEX style file v1.4)
ASCA observations of two steep soft X-ray quasars
F. Fiore1,2,3, G. Matt4, M. Cappi5,6, M. Elvis3, K. M. Leighly6,7, F. Nicastro1,8,
L. Piro8, A. Siemiginowska3, B. J. Wilkes3
1 Osservatorio Astronomico di Roma, Via dell'Osservatorio, I -- 00044 Monteporzio Catone, Italy
2 SAX Science Data Center, Via Corcolle 19, I -- 00131 Roma, Italy
3 Harvard-Smithsonian Center of Astrophysics, 60 Garden Street, Cambridge MA 02138 USA
4 Dipartimento di Fisica, Universit`a degli Studi "Roma Tre", Via della Vasca Navale 84, I -- 00146 Roma, Italy
5 Istituto per le Tecnologie e Studio Radiazioni Extraterrestri (ITESRE), CNR, via Gobetti 101, I -- 40129 Bologna, Italy
6 Cosmic Radiation Laboratory, RIKEN, Hirosawa 2-1, Wako-shi, Saitama, 351, Japan
7 Columbia Astrophysics Laboratory, Columbia University, 538 West 120th Street, New York, NY 10027, USA
8 Istituto di Astrofisica Spaziale (IAS), CNR, Via E. Fermi 21, I00044 Frascati, Italy
14 May 2018
ABSTRACT
Steep soft X-ray (0.1-2 keV) quasars share several unusual properties: narrow
Balmer lines, strong FeII emission, large and fast X-ray variability, rather steep 2-
10 keV spectrum. These intriguing objects have been suggested to be the analogs of
Galactic black hole candidates in the high, soft state. We present here results from
ASCA observations for two of these quasars: NAB0205+024 and PG1244+026.
Both objects show similar variations (factor of ∼ 2 in 10 ks), despite a factor of
about ten difference in the 0.5-10 keV luminosity (7.3 × 1043 erg s−1 for PG1244+026
and 6.4 × 1044 erg s−1 for NAB0205+024, assuming isotropic emission, H0 = 50.0 and
q0 = 0.0).
The X-ray continuum of the two quasars flattens by 0.5-1 going from the 0.1-2
keV band toward higher energies, strengthening recent results on another half dozen
steep soft X-ray AGN.
PG1244+026 shows a significant feature in the '1 keV' region, which can be de-
scribed by either as a broad emission line centered at 0.95 keV (quasar frame) or as
edge or line absorption at 1.17 (1.22) keV. The line emission could be due to reflec-
tion from an highly ionized accretion disk, in line with the view that steep soft X-ray
quasars are emitting close to the Eddington luminosity. Photoelectric edge absorption
or resonant line absorption could be produced by gas outflowing at a large velocity
(0.3-0.6 c).
Key words: Galaxies: Seyfert -- Galaxies: individual: PG1244+026, NAB0205+024
-- X -- rays: galaxies -- Line: formation -- Line: identification
8
9
9
1
r
p
A
4
2
1
v
6
5
2
4
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1
INTRODUCTION
The ROSAT PSPC has found a large spread in the energy
spectral indices of low-z quasars⋆ : 0.5 < α0.1−2keV < 3.5. In
about 10% of cases α0.1−2keV
∼ 2 (e.g. Laor et al. 1994, 1997,
Walter & Fink 1993, Fiore et al. 1994). The large spread in
α0.1−2keV favoured the discovery of its correlation with other
>
properties. In fact, the steep soft X-ray quasars have then
been realized to share a cluster of unusual properties:
• narrow Balmer lines † ( Laor et al 1994, 1997, Boller et
al. 1995);
• strong FeII emission (Laor et al 1994, 1997, Lawrence
et al. 1997)
• Rapid, large amplitude variability (factor of 2-50 on
⋆ We use "quasars" to describe broad line emission objects, re-
gardless of luminosity.
† the permitted lines have FWHM <
clearly broader than the forbidden lines.
∼ 2000 km s−1, yet still are
c(cid:13) 0000 RAS
2
F. Fiore et al.
timescales from minutes to months, Boller et al., 1995,
Brandt et al., 1995, Otani 1995, Boller et al. 1997)
• Somewhat steep hard X-ray spectra (2> α2−10keV >
0.6, Pounds et al. 1995, Brandt et al., 1997);
Pounds et al. (1995), suggest the latter to be a close
physical analogy with the X-ray power-law produced by
Comptonization in a hot accretion disk corona in Galac-
tic black hole candidates (BHC) in their 'soft-high' state.
This is not the only analogy between BHC and steep X-
ray spectrum quasars. Laor et al. (1994, 1997) explained
the correlation with Hβ FWHM as due to the larger size
of a virialized broad emission line region for an AGN in a
high L/LEdd state. Ebisawa (1991) found that while the soft
component of 6 BHC observed by Ginga is roughly stable
on time scales of 1 day or less, the hard component exhibits
large variations down to msec time scales. These timescales
translates to 104 years and 0.1 day for quasars, if they scale
with the mass of the compact object. The soft component of
BHC extends up to ∼ 10 keV in BHC in 'soft-high' states,
and it is often associated with optically thick emission from
an accretion disk. If this is the case, the temperature should
scale with the mass of the compact object as M −1/4
BH , and
the above energy translates to 0.1-0.4 keV for quasars. The
rapid large amplitude variability shown by a few narrow line
Seyfert 1 galaxies (NLSy1) at about 1 keV on timescales of
hours to days (Otani 1995, Brandt et al. 1995, Boller et al.
1997) can then be analogous to the above BHC hard com-
ponent flickering.
A steep X-ray spectrum quasar with 10-100 times the
luminosity of NLSy1s, should be larger and so should vary no
more rapidly than several days. Instead Fiore et al. (1998a)
find that steep X-ray spectrum PG quasars commonly vary
by a factor 2 in 1 day. Variability seems therefore correlated
with X-ray spectral slope and Balmer line width (and there-
fore possibly with the accretion rate) rather than with the
luminosity.
Evidence for spectral features in the '1 keV' region in
many steep soft X-ray quasars is building up (Turner et al.,
1991, Brandt et al., 1994, Otani et al., 1995, Comastri et al.,
1995, Leighly et al., 1997, 1998a,b). Instead, 'normal' Seyfert
1 galaxies (having broad Balmer lines and flatter soft X-
ray spectra) usually have their strongest absorption features
at lower energies (in the 0.6-0.9 keV 'oxygen' band). An
intriguing possibility is that the appearance of these features
at different energies also depend on L/LEdd.
Detailed high energy X-ray spectra of luminous quasars
with steep soft X-ray spectra are essential to understand
the 'narrow-broad line' phenomenon in AGN, in particular
whether the peculiar X-ray properties depend on optical lu-
minosity, optical-to-X-ray ratio (αOX ), or on their Edding-
ton ratio. To this end we selected two bright quasars with
α0.1−2keV >2.0 (Fiore et al., 1994) at the extreme values of
optical luminosity, both with low Galactic NH (Table 1) of
1.9×1020 cm−2 for PG 1244+026, and of 3.0×1020 cm−2 for
NAB0205+024, Elvis et al., 1989) and observed them with
ASCA. We report the results in this paper.
2 OBSERVATION AND DATA REDUCTION
Table 1 gives the redshift, MV , the 0.2 -- 2 keV luminosity,
αOX, the average PSPC count rate and spectral index and
the Galactic NH for the two quasars. Table 2 gives the ASCA
observation log, the SIS and GIS exposure times and count
rates.
Both observations were performed in two CCD mode
with the source at the '1CCD mode' position. Data reduc-
tion was performed using ftools 3.6. We used "bright"
mode SIS data, combining LOW, MEDIUM and HIGH bit
rates. Conservative cleaning criteria were applied (minimum
Earth occultation = 7 degrees, minimum magnetic rigid-
ity = 6 GeV/c, minimum bright Earth angle = 20 degrees,
and excluding data collected in the first 32 seconds after
the satellite passage in the SAA and through the day-night
terminator). Counts, light curves and spectra from the two
quasars were accumulated in circular regions of 3 and 4 ar-
cmin radius for SIS and GIS respectively.
We are interested in the high energy spectrum of these
sources and since they are rather faint, and possibly very
steep, background subtraction plays a crucial role. Back-
ground counts were accumulated from regions surrounding
the sources and compared with counts accumulated from the
same regions from 'blanksky' observations. The 'local' and
'blanksky' background counts were always within 10 % of
each other for the four ASCA instruments. To obtain the
best possible signal to noise in the background subtracted
spectra we therefore used the 'blanksky' background in our
spectral analysis. We extracted background spectra from
'blanksky' event files using the same regions as for source
extraction. The count rates of the two sources become that
of the background at about 7 keV (observer frame). After
background subtraction PG1244+026 is observed in the GIS
up to 10 keV and NAB0205+024 up to 8 keV (9.3 keV quasar
frame), both at the > 3 σ level.
Spectral fits were made separately to the spectra from
the four ASCA instruments and to the spectra obtained
combining together the data from the two SIS and GIS de-
tectors. The results were consistent with each other. In the
following we present the results obtained following the sec-
ond approach. In some cases χ2 are smaller than 1. This
is due to the prescription adopted in adding the spectra,
for the propagation of the errors (the Gehrels, 1986, algo-
rithm: error = 1.0 + SQRT(N + 0.75)). Spectra were always
rebinned following 2 criteria: a) to sample the energy reso-
lution of the detectors with four channels at all energies
where possible, and b) to obtain at least 20 counts per en-
ergy channel. In the spectral fits we limited ourselves to the
0.6-10 keV energy band, to minimize the systematic effect
due to the uncertainty in the SIS calibration below 0.6 keV.
NH was always constrained to be greater than or equal to
the Galactic value along the line of sight.
In all cases the quoted errors represent the 90 % confi-
dence interval for 1 interesting parameter.
3 VARIABILITY
ASCA observed the two quasars for a total elapsed time of
105 ks (PG1244+026) and 119 ks (NAB0205+024). We can
therefore study the variability of these sources on time scales
from a few hundred seconds to about 1.5 days.
Figure 1 and 2 show the SIS light curves of the two
sources. Following Ptak et al. (1994) the values were com-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
ASCA observations of two steep soft X-ray quasars
3
Table 1. Steep soft X-ray quasars
name
z
Ma
V
L0.2−2keV
1045 erg s−1
αOX
PSPC α0.1−2keV
cts/s
NH(Gal)
1020 cm−2
PG1244+026
NAB0205+024
0.048
0.155
-21.1
-25.0
0.14
1.8
1.4
1.6
1.0
0.7
2.3 ± 0.1
2.3 ± 0.1
1.9b
3.0b
a H0 = 50, q0 = 0; b Elvis et al. 1989
Table 2. ASCA observations
name
Dates
Exposure SIS-GIS
2 SIS count rate
2 GIS count rate
PG1244+026
NAB0205+024
1-3 Jul 1996
18-20 Jan 1996
ks
37-39
50-54
cts/s
cts/s
0.442+/−0.004
0.150+/−0.001
0.241+/−0.004
0.153+/−0.002
A drop of a factor of about 2 is present at the beginning of
the light curve of NAB0205+024 on a ∼ 15 ks timescale. In
this paper we limit ourselves to pointing out that roughly
similar variability is observed in two sources which differs
in luminosity by a factor ∼ 10. Significant, although rather
small, spectral variability is also present in the ASCA ob-
servations of the two sources. A systematic analysis of the
variability in different energy bands and of the spectral vari-
ability is in progress and will be presented, together with a
similar analysis on a sample of about 20 Seyfert 1 galaxies
and quasars observed by ASCA and BeppoSAX, in a paper
in preparation. In the following sections we present the aver-
age properties of the spectra. We anticipate that the spectral
variability will not modify the results presented here.
Figure 1. The SIS 0.5-7 keV light curve of PG1244+026.
4 SPECTRAL ANALYSIS
Figures 3 and 4 show the SIS+GIS spectra of PG1244+026
and NAB0205+024 fitted with a simple power law absorbed
at low energy by a column of cold gas equal to or higher
than the Galactic column along the line of sight. Table 3
gives the best fit parameters and the χ2. It is clear that
this simple model is inadequate to describe the 0.6-10 keV
spectrum of both quasars. A hard tail, larger than the 5-10
% systematic uncertainties at these energies (Gendreau &
Yaqoob 1997), is evident in both cases. In the spectrum of
PG1244+026 there is also a significant excess with respect
to the model about 1 keV. The feature is visible in both SIS
and GIS detectors. We discuss these two findings in turn in
the next two sections.
4.1 0.6-10 keV continuum
The ASCA SIS and GIS responses are peaked at 1-2 keV
and decrease sharply at higher energies. Therefore the ASCA
'2-10 keV' slopes are strongly biased toward the lowest en-
ergy boundary. This means that some caution should be
used when comparing ASCA '2-10 keV' slopes with those
of experiments whose responses peak around 6 keV such
as EXOSAT, GINGA, BeppoSAX and XTE. To address
this, we fitted the SIS+GIS spectra of PG1244+026 and
NAB0205+024 with a simple power law in the observed 0.6-
Figure 2. The SIS 0.5-7 keV light curve of NAB0205+024
puted using variable bin sizes corresponding to good time
intervals longer than 200 seconds.
Significant variations are evident in both light curves. A
factor of two variability, both down and up, in about 10 ks is
present at the beginning of the light curve of PG1244+026.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4
F. Fiore et al.
Table 3. PG1244+026 & NAB0205+024 spectral fits
model
N a
H
αE or Tb
αH or Ω/2π Eb
break or Ac
χ2 (dof)
PG1244+026
PL 0.6-10 keV
PL 2-10 keV
PL 3-10 keV
PL 4-10 keV
Broken PL 0.6-10 keV
PL+Raym 0.6-10 keV
PL+BB 0.6-10 keV
PL+Comp.Refl 0.6-10 keV
NAB0205+024
PL 0.6-10 keV
PL 2-10 keV
PL 3-10 keV
PL 4-10 keV
Broken PL 0.6-10 keV
PL+Raym 0.6-10 keV
PL+BB 0.6-10 keV
PL+Comp.Refl 0.6-10 keV
1.9+0.6
1.67±0.04
1.9FIXED 1.35±0.12
1.9FIXED 1.03±0.30
1.9FIXED 0.67±0.55
1.80+1.1
−0.2
0.88+0.2
−0.05
0.16±0.03
2.78±0.06
3.6+2.6
−1.7
1.9+0.7
5.3+8.0
−3.4
2.6+1.3
−0.7
3.0+0.1
1.38±0.02
3.0FIXED 1.09±0.10
3.0FIXED 0.94±0.20
3.0FIXED 0.62±0.45
1.50+0.3
−0.1
0.49±0.05
0.16±0.02
1.44±0.04
3.0+1.4
3.0+7.0
3.0+0.2
3.0+0.2
--
--
--
--
1.06+0.24
−0.38
1.54±0.06
1.40±0.15
> 8
--
--
--
--
--
--
--
--
2.8±0.6
> 0.5
--
--
--
--
--
--
0.98±0.10
1.02±0.25
1.15±0.07
> 4.2
2.4±0.5
< 0.01
--
--
177.9 (148)
48.4 (82)
29.1 (58)
14.4 (38)
149.1 (145)
135.3 (145)
139.2 (145)
150.5 (146)
126.2 (147)
42.4 (86)
23.3 (58)
9.5 (38)
85.2 (145)
82.8 (144)
86.2 (145)
95.2 (146)
a in 1020 cm−2; b in keV; c metal abundances
Figure 3. The SIS+GIS spectra of PG1244+026 fitted with a
simple absorbed power law model. The lower panel shows the
ratio between the data and the best fit model.
Figure 4. The SIS+GIS spectra of NAB0205+024 fitted with
a simple absorbed power law model. The lower panel shows the
ratio between the data and the best fit model.
10, 2-10, 3-10 and 4-10 keV ranges. The results are given in
Table 3 and show in Figure 5.
In the last three series of fits the NH was fixed to the
Galactic value. The best fit slopes flatten by 0.5 -- 1 going
from the low energy dominated to the higher energy range.
(We note that for NAB0205+024 the quasar redshift implies
that these slopes refer to slightly harder energy ranges: 10
keV in the observer frame corresponds to 11.6 keV in the
quasar frame).
The PSPC spectral index for both quasars is 2.3(±0.1,
Fiore et al., 1994), steeper than any of the values in table
3. For NAB0205+024 the ASCA low energy index is flatter
than the PSPC by 0.4-0.8, and the ASCA best fit NH coin-
cides with the Galactic value, suggesting that the spectrum
continues to steepen below ASCA X-ray energies.
To parameterize a curved spectrum we have also fitted
the 0.6-10 keV spectra of the two quasars with: a broken
power law model; a power law + optically thin plasma emis-
sion (Raymond & Smith 1977) model; a power law + black
body model; a power law + Compton reflection model. The
results are again in Table 3 and confirm the presence of sig-
nificant curvature in the spectra of these quasars.
In PG1244+026 the power law + Raymond-Smith
model gives a χ2 significantly better than the broken power
law model because it fits also the ∼ 1 keV feature (see next
section). However, the best fit energy index is still very steep
and positive residuals are evident above 4 keV (Figure 6).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
ASCA observations of two steep soft X-ray quasars
5
Figure 6. The SIS+GIS spectra of PG1244+026 fitted with a
power law + Raymond-Smith model. The lower panel shows the
ratio between the data and the model
Figure 7. The SIS+GIS spectra of PG1244+026 fitted in the
0.6-4 keV range. The lower panel shows the residuals when the
spectrum is fitted with a simple absorbed power law
4.2 Low energy features
We have already noted that a '1 keV' feature seems to be
present only in PG1244+026, the quasar with the lower lu-
minosity. Figure 7 shows the results of a simple power law
model fit to the 0.6-4 keV spectrum of PG1244+026: the '1
keV' feature is clearly visible.
It has been recently realized that Residual Dark Distri-
bution (RDD, the error in ASCA's onboard correction for
CCD dark current) can affect the low energy ASCA SIS
spectra. The symptom of the RDD problem is a sudden de-
crease in the SIS effective area towards lower energies from
about 1 keV. We are however confident that this has little
effect on the '1 keV' feature for the following reasons: a) SIS
data of PG1244+026 and NAB0205+024 have similar RDD
value but the '1 keV' feature is visible in the spectrum of
Figure 5. The PG1244+026 (filled circles) and NAB0205+024
(open triangles) best fit power law index in the observed 0.6-10, 2-
10, 3-10 and 4-10 keV ranges as a function of the range minimum
energy. The NAB0205+024 points have been shifted from real
energy for a sake of clarity. The PSPC 0.1-2 keV points are also
showed for comparison.
In NAB0205+024 the power law + Raymond-Smith model
gives an acceptable χ2 only for low metal abundances, given
the absence of significant features in the spectrum, similar
to the ROSAT results of Fiore et al. (1994). Above 4 keV the
best fit is similar to that in the broken power law model. Fits
with a power law + black body models give acceptable χ2
in both cases. The power law indices in this case are slightly
steeper than in the broken power law model. Inspection of
the residuals shows again a slight excess of counts at high
energy. In PG1244+026 the power law + black body model
again gives a χ2 significantly better than the broken power
law model because it partly fits also the feature around ∼ 1
keV (see next section). Fits with a Compton reflection model
(plrefl in xspec) give χ2 significantly higher than the pre-
vious models and push the parameter Ω/2π to high and
implausible values.
We do not see any significant line emission at the en-
ergies of the iron Kα lines. The 90 % upper limits to the
equivalent width of a narrow line at 6.4 (and 6.7 keV), rest
frame, in PG1244+026 and NAB0205+024 are 400 eV and
230 eV (640 eV, 314 eV) respectively.
The χ2 for the fits to the PG1244+026 spectra are much
higher than those for the similar fits to the NAB0205+024
spectra because of the presence of the '1 keV' feature in the
former quasar. We discuss this feature next.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
6
F. Fiore et al.
PG1244+026 only, the 90 % upper limit to any unresolved
gaussian line emission at about 1 keV in NAB0205+024 is
of 20 eV; b) the feature is present also in the GIS spectrum;
the feature is present at the same level in SIS0 and SIS1,
while the RDD tends to be greater for SIS1 than for SIS0.
To investigate the physical origin of the '1 keV' feature
and to quantify its strength we performed a series of spectral
fits to the SIS and GIS spectra of PG1244+026 using only
the 0.6-4 keV range, to avoid the complications of the hard
tail (see previous section). The results are reported in Table
4, along with those of the other fits in this section.
4.2.1 Emission Line model
The fit with a power law model plus a gaussian line gives a
small χ2 (90.1 for 101 degrees of freedom). The equivalent
width of the 1 keV feature, in the power law plus gaussian
line fit, is similar in the SIS and GIS detectors. In Table 4 we
report the SIS determination of 64±15 eV. The line width
is well constrained and cleanly resolved in the SIS spectrum
to σ = 0.11+0.01
−0.03 keV.
A fit with a power law plus a thermal plasma model
(Raymond-Smith 1977) gives an acceptable χ2 (93.1, 100
dof). The best fit temperature (1 keV) implies that the '1
keV' emission is dominated by a blend of iron-L and neon
emission lines.
An emission line feature can be mimicked by fitting a
spectrum with a strong absorption feature at slightly higher
energies. We have then fitted the SIS and GIS spectra with
models including absorption structures.
4.2.2 Absorption Edge Fits
Fits with one absorption edges gives χ2 significantly higher
than the previous case (102.5, 102 dof). The edge energy
(1.17 keV, quasar frame) is consistent with that of NeIX
and/or iron L FeXVI and FeXVII. The best fit neutral NH
is significantly higher than the Galactic value. This is reason-
able, since if there is highly ionized Ne and Fe L absorption
it is likely to have also highly ionized oxygen absorption at
0.74-0.87 keV. We then refitted the SIS and GIS spectra with
a model including three absorption edges, at the energies of
the most abundant ions in highly ionized gas with high NeIX
abundance: OVII, OVIII and NeIX-FeXVI-FeXVII (Nicas-
tro et al. 1998), fixing the cold NH to the galactic value. The
results were not satisfactory. The depth of the oxygen edges
is zero with small upper limits and the χ2 is significantly
higher than in the previous case: 114.7. Leaving NH free
improves the χ2, but the oxygen edge depths are still zero
and the fit resembles completely the single edge fit. Fixing
NH to the Galactic value but leaving free the energies of
two edges produces again a good fit (χ2 = 94.8, 101 dof).
The best fit energy of one edge is again 1.12 keV, but that
of the other edge is < 0.64 keV (observer frame), close to
the lower boundary of the observed range. So, if the cold
absorption is fixed to the Galactic value, then there must
be additional absorption edge(s) at energies lower than the
observed range, corresponding to oxygen less ionized than
OVI. We note however that this conclusion is weakened by
the unknown contribution of the SIS RDD, which pushes
low energy events below the detection threshold.
A 1.17 keV absorption feature can also be interpreted in
terms of blueshifted oxygen absorption (Leighly et al. 1997).
In this case, assuming that the absorption is mostly due to
OVIII, the shift from the quasar frame would be equivalent
to z= -- 0.38. A more complex continuum has little effects on
the best fit parameters of absorption edges.
4.2.3 Ionized absorber models Fits
We fitted the data with a detailed ionized absorber model
(not including resonant scattering absorption lines). We first
generated a grid of photoionization equilibrium models us-
ing Cloudy (Ferland 1996), and fitted the spectrum in-
terpolating by this grid, using the method of Fiore et al.
(1993). To calculate the models we have assumed the ob-
served spectral energy distribution (Fiore et al. 1995, Elvis
et al. 1994). This is important, since the soft X-ray spectrum
of this source strongly differs from that of 'normal' Seyfert 1
galaxies, where warm absorbers are usually found (Reynolds
1997). A steep soft X-ray spectrum can completely ionize
oxygen and neon but not iron, and so can produce edges in
the 1-2 keV (Fe-L) and 7-9 keV (Fe-K) ranges, but not in the
'oxygen' 0.6-0.9 keV band. A fit with this model produces
an acceptable χ2 (see Table 4, ionized absorber model 1).
In Figure 8 we show the best fit steep SED model (thick
line), and a photoionization model obtained using a stan-
dard, much flatter AGN SED (a power law of α = 1.2 from
UV to X-rays, thin line), which, above 1 keV, gives a compa-
rably good fit to the data. While iron in the flat SED model
has a ionization structure similar to that of the model ob-
tained using the right SED, oxygen is much less ionized: note
the deep OVIII edge present in the flat SED model.
In this fit the redshift of the absorber is significantly
higher than that of the quasar, because the main feature in
the transmitted spectrum is the FeXVIII edge at 1.36 keV,
while the deepest edge in the quasar spectrum is at 1.17 keV
(quasar frame). However, a good fit can be also obtained for
a different absorber redshift (z= -- 0.33, Table 4, ionized ab-
sorber model 2). In this case the 1.17 feature is interpreted
in terms of OVII and OVIII absorption. We cannot discrim-
inate between these two solutions on statistical grounds.
We have also tried fits with a collisional equilibrium
model (Nicastro et al. 1998). The results were very similar
to those obtained in the case of photoionization equilibrium
(see Table 4, ionized absorber model 3).
In all fits with detailed warm absorber models the col-
umn of cold gas is significantly higher than the Galactic,
value, similar to the values found using single edges to pa-
rameterize the absorber (Table 4).
4.2.4 Absorption Line Fits
Fits with a single gaussian absorption line do not give ac-
ceptable χ2. Fits with two or more absorption gaussian lines
can produce χ2 of 96 or smaller. These models are indistin-
guishable for the SIS from models with a broad absorption
notch, which we discuss in the following. Physical models
including resonant absorption lines, as well as absorption
edges from ionized plasma will be discussed in a paper in
preparation (Nicastro et al. 1998b).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Table 4. PG1244+026: spectral fits in the band 0.6-4 keV
ASCA observations of two steep soft X-ray quasars
7
Continuum models
model
αE
PL
PL + BB
PL + Raym
emission line
1.70±0.02
1.50±0.11
1.61±0.03
N a
H
1.9+1.0
9.5±5.0
1.9+0.3
Tb
--
0.14 ± 0.02
1.00±0.06
Ac
--
--
>0.5
line Eb
line EWd
PL + Gauss
1.60±0.04
1.9+0.5
0.91±0.03
64±15
absorption edge models
PL + 1 edge
PL + 2 edges
1.88±0.11
1.88±0.11
8.2±3.0
1.9FIXED
ionized absorber models
edge Eb
1.12±0.03
1.12±0.10
< 0.64
τ
0.25±0.07
0.26±0.08
0.51±0.14
--
--
--
--
--
--
1 -- Phot. ion.
2 -- Phot. ion.
3 -- Coll. ion.
1.92±0.10
1.94±0.10
1.90±0.10
11.1±3.5
11.0±3.5
11.3±3.5
22.59±0.15
21.56±0.10
22.62±0.14
2.00±0.12
1.20±0.10
6.92±0.07
0.25±0.07
-0.33±0.10
0.24±0.06
logNH (ionized)
U or logT
z abs.
χ2 (dof)
131.8 (104)
109.5 (101)
93.1 (100)
χ2 (dof)
90.1 (101)
χ2 (dof)
102.5 (102)
94.8 (101)
χ2 (dof)
93.2 (101)
95.7 (101)
90.7 (101)
absorption notch models
PL + notch
PL+BB+notch
1.93±0.08
1.48±0.20
9.3±3.3
7.8±4.4
Tb, notch Eb
notch widthd
Cov. frac.
χ2 (dof)
1.44±0.05
680±70
0.14±0.04
T = 0.16 ± 0.03
1.16±0.03
14±7
1FIXED
97.5 (100)
97.1 (100)
a in 1020 cm−2; b in keV; c metal abundances; d in eV
4.2.5 Absorption Notch Fits
Fits with an absorption notch give χ2 higher than those
with an emission lines by ∆χ2 ≈ 7. While the power law +
notch fit is formally acceptable, the best fit value of the notch
width is implausibly large (almost 1 keV), forced by the very
low value of the covering fraction required by the fit. How-
ever, the notch best fit parameters are strongly dependent on
the proper modeling of the continuum. For example, using
a power law + a black body to parameterize the continuum
gives an acceptable fit fixing the notch covering fraction to
1, which in turn results in a much more reasonable value for
the notch width of 14±7 eV. The ASCA band width in not
wide enough and its spectral resolution is not good enough
to constrain adequately both a complex continuum and the
notch parameters.
4.2.6 Comparison with the PSPC results
PG1244+026 was observed with the PSPC in December
1991 and the results of this observation have been reported
by Fiore et al. (1994). The 0.6-2 keV flux level during the
PSPC observation was about 50 % lower than the mean flux
in the ASCA observation. The fit of an absorbed power law
to the PSPC spectrum gives an acceptable χ2 (22.8 for 25
dof), α0.1−2keV = 2.3 ± 0.1 and NH = 2.9 ± 0.3 × 1020 cm−2,
slightly higher than the Galactic value. Most of the PSPC
counts were detected below 0.3 keV, in the 'Carbon' band.
The quality of the spectrum between 0.3 and 2 keV is not
very high and emission or absorption features fainter than
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
∼ 20% cannot be excluded in this energy band. No evidence
of spectral variability is present in the ROSAT data despite
a factor of 2 flux variability.
The PSPC data strongly constrain the level of any cold
or warm absorption affecting the 'Carbon' band. Best fitting
models to the ASCA data including absorption features in
the 1-2 keV band (Table 4) require a rather large absorp-
tion in addition to the Galactic one below 1 keV. Therefore,
it is important to study whether the ASCA best fit models
are consistent with the PSPC ones. Rather than performing
joint fits to the ASCA and PSPC data, which are compli-
cated by the large uncertainty in the relative PSPC/ASCA
SIS calibration, and by the detailed shape of the continuum
over the broad 0.1-4 keV band, we fitted the PSPC data
with a power law model including the emission or absorp-
tion features found in the previous section (see Table 4). The
results are in Table 5, where we also report (in brackets) the
99 % parameter upper limits, or confidence intervals, when
appropriate.
We see that the presence of an emission line at 0.91 keV
is not required by the PSPC spectrum, but an equivalent
width of 65 eV is not excluded (10 % probability). However,
the presence of a cold absorber of thickness 8.2 ± 3.0 × 1020
cm−2 is inconsistent with the PSPC spectrum (probability
< 1%), while the presence of an edge at 0.62 keV with τ =
0.51 ± 0.14 is only marginally consistent with the PSPC
result.
8
F. Fiore et al.
Table 5. PG1244+026: spectral fits to the PSPC
model
PL
PL + gauss
PL + 1 edge
PL + 2 edges
N a
H
2.9±0.3
2.9±0.3
2.7±0.4(0.6)
2.7±0.4
αE
line or edge Eb
line EWc or τ
χ2 (dof)
2.3±0.1
2.3±0.1
2.2±0.2
2.2±0.2
0.91FIXED
1.12FIXED
1.12FIXED
0.62FIXED
0+68 (+100)
0.3+0.4
−0.3
0.3+0.4
−0.3
0+0.34(+0.62)
22.8 (25)
22.8 (23)
20.9 (24)
20.9 (23)
a in 1020 cm−2; b in keV; c in eV
components would be quite different from 'normal' quasars.
Laor et al. (1997) suggested a fainter hard component rela-
tive to the optical in the majority of low redshift PG quasars
(assuming a two component model). However, Grupe (1996)
found evidence for a stronger soft excess in a sample of soft
X-ray selected Seyferts dominated by narrow-line objects.
A large relative intensity of the soft component has im-
portant consequences on various competing models for the
soft component. Disc reprocessing models (Matt et al. 1993,
Fiore et al. 1997) would require highly anisotropic emission
to account for the discrepancy between the observed soft
and hard fluxes. Optically thin free-free emission (e.g. Bar-
vainis 1993) is ruled out by these observations because the
best fit power law slope is still too steep to fit the spec-
trum above 4 keV, because variability rules out optically
thin plasma (Elvis et al. 1991) and because of the implausi-
bly low metal abundances (< 1% solar, see Table 3) required
in NAB0205+024 (see Sect. 3.2.1). The most likely origin for
the steep component is Comptonized disc emission (e.g. Cz-
erny & Elvis 1987, Fiore et al. 1995, Pounds et al. 1995).
The high energy spectral index of the two quasars αH ∼
1.0 (see Table 3) is consistent with that of 'normal' Seyfert
1 galaxies (e.g. Nandra & Pounds 1994, Nandra et al. 1997).
The error on αH is however large and so no strong conclusion
can be drawn on the origin of the hard emission. An answer
to this question must await the large area and high energy
sensitivity of AXAF, XMM and Spectrum X-gamma.
5.2 Origin of the ∼ 1 keV feature in PG1244+026
The ∼ 1 keV feature in PG1244+026 could be explained
in terms of either a broad (σ = 0.1 keV) emission line at
0.91 keV (0.95 keV quasar frame) of about 60 eV equivalent
width or a τ = 0.25 absorption edge at 1.17 keV (or an
absorption notch at 1.22 keV). These possibilities cannot be
discriminated between on statistical grounds.
An absorption interpretation requires additional low en-
ergy absorption, either cold (with a column density higher
than Galactic by ∼ 7 × 1020 cm−2 (see Tables 4 and 5), in
contrast with the PSPC results, or a warm absorber with a
peculiar ion abundance distribution.
The observed spectrum can be interpreted in terms of
either an inflowing (v/c=0.25) or an outflowing (v/c= -- 0.33)
absorber. In the first case the ion contributing most to the
absorption is FeXVIII, in the second case it is OVIII. The
two cases cannot be discriminated on statistical grounds.
While an outflowing absorber has been suggested in several
other cases (see e.g. Mathur et al. 1994), this would be the
first case for an inflowing highly ionized absorber.
A similar situation is found in IRAS 13224-3809 by
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 8. The thick line shows the best fit photoionized absorber
model for PG1244+026 using the observed steep spectrum (note
that the absorber has z=0.25, while the quasar has z=0.048). The
thin line shows a photoionization model obtained using a much
flatter SED (see text). Both models produce Fe-L absorption fea-
tures around 1 keV. The flat SED also produces deep OVIII edge,
which is not observed
5 DISCUSSION
5.1 Continuum
The ASCA observations of PG1244+026 and NAB0205+024
have shown that the X-ray continuum of these two quasars
flattens by 0.5-1 passing from the 0.1-2 keV (PSPC) to the
2-10 keV band. Similar results were obtained by Brandt,
Mathur & Elvis (1997); and by Comastri et al. (1998) and
Leighly et al. (1998a) on TONS180, Pounds et al. (1995) and
Fiore et al (1998b) on REJ1034+390, Leighly et al. (1998b)
on AKN564. It appears that the X-ray spectrum of a sizeable
number of steep PSPC and narrow Balmer line quasars has
significant curvature, being flatter at higher energies.
This could be due to different components influencing
the spectrum at different energies, as might happen in 'nor-
mal' broad lines Seyfert 1 galaxies and quasars, where a soft
component is often present. The relative intensity of the two
ASCA observations of two steep soft X-ray quasars
9
Otani et al. (1995), in AKN564 by Brandt et al. (1994) and
by Leighly et al. (1998a). Otani et al (1995) and Leighly et
al. (1997) interpret the features in the 1 -- 2 keV band in terms
of blueshifted absorption from relativistically (v=0.2-0.6 c)
outflowing material. If the 1.17 keV feature in PG1244+026
is due to a blueshifted OVIII photoelectric absorption, then
the absorption seen below 0.64 keV may be due to CVI pho-
toelectric absorption from the same gas. This distribution is
far from an equilibrium distribution (see eg. Nicastro et al.
1998), not an impossible situation considering the large vari-
ability observed in this source.
The absorption features seen between 1 and 2 keV may
also be interpreted in terms of resonant lines (e.g. Leighly
et al. 1997). If the ion producing the absorption is oxygen
OVIII (resonant line at 0.65 keV), the best fit notch energy
of 1.22 keV (quasar frame) implies a very high gas velocity:
-- 0.56c. Fe XVIII (E=0.87 keV) or Fe XVII (E=0.81 keV)
can also contribute to the absorption, because of their high
oscillator strengths, 1.7, 0.6 respectively (Kato et al. 1976),
and abundances. If the 1.22 keV absorption notch is due to
these ions then the velocity of the outflowing gas will be
smaller, -- 0.33c.
In any case, in the blueshifted absorption scenario the
gas is outflowing at velocities which are a sizeable fraction
of c, reminiscent of blobs of gas in jets. It is interesting to
note that similar absorption features have been observed in
Blazars (e.g. PKS2155-304, Canizares & Kruper, 1884, other
BlLacs, Madejski et al. 1991, 3C273, Grandi et al. 1997), but
usually below 1 keV. Somewhat surprisingly this implies less
extreme conditions in these radio-loud objects than in our
radio-quiet quasars. High redshift radio loud quasars may
have similar jet-related absorption too (Elvis et al., 1997).
An alternative interpretation of the 1 keV feature is in
terms of an emission line due to highly ionized oxygen, neon
(NeIX) and/or to Iron L. There are two possible origin for
this line: recombination in an optically thin thermal plasma,
or "reflection" in photoionized matter.
5.2.1 Thermal plasma
A thermal plasma is highly implausible on physical grounds
(Elvis et al. 1991, Fiore et al. 1995). Emission measure is
∼ 2.5×1065 cm−3, For a spherical source (with radius R)
and constant electron density ne, ne = 2.43 × 1032R−
cm−3. R < 1014 cm from the observed X-ray variability,
ne < 1.5×1012 cm−3, This implies an electron scattering
∼ 30. With such values a thermal plasma
optical depth τT
is no longer optically thin. The situation is even worse if
the matter is clumpy, as the density of each cloud must be
greater.
3
2
>
the matter would be thick. Since both the "reflector" and
the primary emission (which provides most of the contin-
uum) are observed, the optically thick case gives the highest
values of the equivalent width. However, even in the opti-
cally thick case the expected EWs can barely account for the
observed values (see below); hence we neglect the optically
thin case altogether.
The observed line (which is significantly broad) may be
a blend of several lines (see for instance Zycki et al. 1994,
Netzer 1997). The observed energy suggests the 0.92 keV Ne
ix and 1.02 keV Ne x recombination lines, the 0.87 keV Oxy-
gen viii recombination (to ground state) line and the iron
L (around 0.8 keV) lines being the most important. None
of these lines alone can account for the observed EW: for
instance, the maximum value (i.e. for a face -- on disk with a
intervening ion fraction ∼ 0.6) for the Ne ix line is about 10
eV, while that of the oxygen recombination line is about 15
eV (note that a O viii Kα recombination line at 0.65 keV
with a similar EW should also be present; the 90% upper
limit on such a line is 30 eV). These values have been cal-
culated using the formulae of Basko (1978), and assuming a
reasonable ionization structure. (In the disc hypothesis an
iron Kα line at 6.5-6.9 keV is also expected, but the up-
per limit of 300-400 eV does not exclude the presence of
such a line.) Allowing for a possible factor of 2 neon and/or
iron overabundance (an oxygen overabundance would de-
crease the Ne line while not increasing the O line) and/or
anisotropy of the illuminating radiation, the observed equiv-
alent width could be explained (note that in many Seyfert
1 galaxies the iron Kα line is also stronger than expected,
suggesting iron overabundance or anisotropic illumination).
In this scenario the '1 keV' feature may arise from an
highly ionized accretion disk. In the Matt et al. (1993) mod-
els high ionization is mainly due to a high accretion rates
(the ionization parameter depends on m3). The detection
of these emission lines in steep soft X-ray quasars would
then be further evidence of high L/LEdd. We note that since
recombination can occur only in highly ionized atoms, we
would not expect features of this kind in 'normal' quasars,
as their disc should be much less ionized, as in fact observed.
The strong dependence of the ionization parameter on
m allows for large differences in the ionization structure
against small differences in m and therefore that the '1 keV'
feature may not be ubiquitous in NLSy1s. Indeed, a simi-
lar feature is not present in NAB0205+024. A line with the
same equivalent width as in PG1244+026 (60 eV) would
have been detected in the NAB0205+024 SIS spectrum (the
90 % upper limit is only 20 eV). In the accretion disc sce-
nario this would imply a different ionization state (higher or
lower) of the matter or a significant metal underabundance
in NAB0205+024.
5.2.2 Photoionized matter
5.3 Variability
The second possibility is that the '1 keV' emitting matter
is photoionized by the central nucleus. We can assume that
the gas is not covering the source because there is not signif-
icant OVII and/or OVIII absorption. We therefore assume
that we are not observing a "warm absorber", but rather
a "warm reflector". This could either be a warm absorber
viewed from its side, in which case the matter would be opti-
cally thin to Thomson scattering; or the accretion disc, and
The mean 0.5-10 keV luminosity measured by ASCA in
PG1244+026 and NAB0205+024 differs by an order of mag-
nitude: 7.3×1043 and 6.4×1044 erg s−1 respectively (assum-
ing isotropic emission, H0 = 50.0 and q0 = 0.0). The opti-
cal (3000 A) monochromatic luminosities differs even more:
4 × 1043 and 9.4 × 1044 erg s−1 respectively. The variations
seen in the NAB0205+024 light curve imply an efficiency in
the conversion of matter into radiation greater than 1.6 %
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
10
F. Fiore et al.
(e.g. Fabian 1984). Taken at face value, this excludes ther-
monuclear reactions as the origin of the observed X-ray lu-
minosity for which the upper limit on the efficiency is in this
case about 0.7 %.
The similar variability observed in the two quasars
agrees with the Fiore et al. (1998a) finding that the variabil-
ity properties of (PG) quasars are correlated with the shape
of the soft X-ray spectrum and the width of the Balmer lines,
and so possibly then with the accretion rate, in the scheme
of Pounds et al. (1995), and Laor et al. (1994), (1997).
6 CONCLUSIONS
ASCA observations of two steep soft X-ray quasars have
shown that:
(i) The X-ray continuum of the two quasars flattens by
∆α = 0.5 − 1 going toward high energies. Similar results
were obtained by authors on some half dozen steep soft X-
ray quasars.
(ii) PG1244+026 shows a significant feature in the '1 keV'
region. Similar features were again reported in other steep
soft X-ray quasars. The data are not good enough to dis-
criminate between a broad emission line centered at 0.95
keV (quasar frame) or an absorption edge at 1.17 keV, or
an absorption notch at 1.22 keV.
Line emission could be due to reflection from an highly
ionized accretion disk, in line with the view that steep soft
X-ray quasars are emitting close to the Eddington luminos-
ity. Photoelectric edge absorption or resonant line absorp-
tion could be produced by gas outflowing at a large veloc-
ity (0.3-0.6 c). In these absorption models significant cold
(i.e. oxygen less ionized than OVI) absorption in excess of
the Galactic is required. This would imply an increase by
a factor 2-3 of the cold column with respect to a previous
PSPC observation or a peculiar ionization structure. In nei-
ther the emission or absorption cases the SIS resolution is
good enough to identify unambiguously the ions responsible
for the feature. The high resolution and high throughput
of the low energy gratings and spectrometers of AXAF and
XMM are clearly needed to shed light on this puzzling case.
(iii) The two quasars show similar variability properties
(flux variations up to a factor of 2 in 10 ks) despite a fac-
tor of ten difference in the X-ray observed luminosity. This
agrees with the Fiore et al (1998a) finding that the variabil-
ity properties of radio-quiet quasars are correlated with the
shape of the X-ray spectrum, the width of the Balmer lines
and so possibly with the accretion rate.
F.F acknowledges support from NASA grants NAG 5-
2476 and NAG 5-3039, B.J.W. acknowledges support from
ASC contract NAS8-39073.
REFERENCES
Barvainis R., 1993, ApJ, 412, 513
Basko M.M., 1978, ApJ, 223, 268
Boller Th., Brandt W.N., Fink H., 1996, A&A, 305, 53
Boller Th., Brandt W.N., Fabian A.C., Fink H., 1997, MNRAS,
289, 393
Brandt W.N., et al., 1994, MNRAS, 271, 958
Brandt W.N, Pounds K.A., Fink H.H., 1995, MNRAS, 273, 47
Brandt W.N., Mathur S., Elvis M., 1997, MNRAS, 285, L25
Canizares C.R., Kruper J., 1984, ApJL, 278, 99
Comastri A., Molendi S., Ulrich M.-H., 1995, Proc. "X-ray imag-
ing and Spectroscopy of Cosmic Hot Plasmas" ed. F. Makino
and K. Mitsuda (Tokyo University Academy Press) p. 279
Comastri A., et al., 1998, A&A, in press
Czerny B., Elvis M., 1987, ApJ, 321, 305
Ebisawa K., 1991, PhD thesis, ISAS Research Note 483
Elvis M., Lockman F.J., Wilkes B.J., 1989, AJ, 97, 777
Elvis M., Giommi P., McDowell J., Wilkes B.J. 1991, ApJ, 378,
537
Elvis M., Wilkes B.J., McDowell J.C., Green R., Bechtold J.,
Willner S.P., Oey M.S., Polomski E., Cutri R., 1994, ApJS,
95, 1
Elvis M., Fiore F., Giommi P., Padovani P., 1997, ApJ, ApJ, 492,
91
Fabian A.C., 1984, Physica Scripta, T7, 129
Ferland G.J. 1996 (Cloudy 90.01)
Fiore F., Elvis M., Mathur S., Wilkes B., McDowell J., 1993, ApJ,
415, 129
Fiore F., Elvis M., McDowell J. C., Siemiginowska A., Wilkes B.
J., 1994, ApJ, 431, 515
Fiore F., Elvis M., Siemiginowska A., Wilkes B.J., McDowell J.C.,
Mathur S., 1995, ApJ, 449, 74
Fiore F., Elvis M., 1997, in proceedings of the 1994 30th COSPAR
meeting "High energy radiation from Galactic and extragalac-
tic black holes", Adv. Space Res. Vol. 19 No 1, p. 85
Fiore F., Matt G., Nicastro, F., 1997a, MNRAS, 284, 731
Fiore F., Laor A, Elvis M., Nicastro F, Giallongo E., 1998a, ApJ,
submitted
Fiore F., et al., 1998b, A&A in preparation
Gehrels 1986, ApJ, 303, 336
Gendreau K., Yaqoob T., 1997, ASCA News, n. 5, p. 8
Grandi P., et al., 1997, A&A, A&A, 325, L17
Grupe D., 1996, PhD Thesis, University of Gottingen
Kato T., 1976, ApJS, 30 397
Laor A., Fiore F., Elvis E., Wilkes B.J., McDowell J.C., 1994,
ApJ, 435, 611
Laor A., Fiore F. Elvis E., Wilkes B.J., McDowell J.C., 1997,
ApJ, 477, 93
Lawrence A., Elvis M., wilkes B.J., McHardy I., Brandt N., 1997,
MNRAS, 285, 879
Leighly K.M., Mushotzky R., Nandra K., Forster K. 1997, ApJ,
489, L25
Leighly, K.M., et al., 1998a, in preparation
Leighly, K.M., et al., 1998b, in preparation
Madejsky G.M., Mushotzky R.F., Weaver K.A., Arnaud K.A.,
1991, ApJ 370, 198
Matt G., Fabian A.C., Ross R.R., 1993, MNRAS, 264, 839
Mathur S., Wilkes B.J., Elvis M., Fiore F. 1994, ApJ, 434, 493
Nandra P., George I.M., Mushotzky R.F., Turner T.J., Yaqoob
T., 1997, ApJ, 477, 602
Netzer H., 1996, ApJ, 473, 781
Nicastro F., Fiore F., Perola G.C., Elvis M., 1998, ApJ, submitted
Otani C., Kii T., Miya K., 1996, MPE Rep. 263, p.491
Pounds K.A., Done C., Osborne J.P., 1995, MNRAS, 277, L5
Ptak A., Yaqoob T., Serlemitsos P.J., Mushotzky R. Otani C.
1994, ApJ, 436
Reynolds C.S., 1997, MNRAS, 287, 513
Tanaka Y., Inoue H, Holt S.S., 1994, PASJ, 46, L37
Turner T.J., Weaver K.A., Mushotzky R.F., Holt S.S., Madejsky
G.M., 1991, ApJ, 381, 85
Walter R., Fink H., 1993, A&A, 274, 105
Zycki P.T., Krolik J.H., Zdziarski A.A., Kallman T.R., 1994, ApJ,
437, 597
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0510289 | 1 | 0510 | 2005-10-10T17:57:59 | Effects of Type I Migration on Terrestrial Planet Formation | [
"astro-ph"
] | Planetary embryos embedded in a gas disc suffer a decay in semimajor axis -- type I migration -- due to the asymmetric torques produced by the interior and exterior wakes raised by the body (Goldreich & Tremaine 1980; Ward 1986). This presents a challenge for standard oligarchic approaches to forming the terrestrial planets (Kokubo & Ida 1998) as the timescale to grow the progenitor objects near 1 AU is longer than that for them to decay into the Sun. In this paper we investigate the middle and late stages of oligarchic growth using both semi-analytic methods (based upon Thommes et al. 2003) and N-body integrations, and vary gas properties such as dissipation timescale in different models of the protoplanetary disc. We conclude that even for near-nominal migration efficiencies and gas dissipation timescales of ~1 Myr it is possible to maintain sufficient mass in the terrestrial region to form Earth and Venus if the disc mass is enhanced by factors of ~2-4 over the minimum mass model. The resulting configurations differ in several ways from the initial conditions used in previous simulations of the final stages of terrestrial accretion (e.g. Chambers 2001), chiefly in (1) larger inter-embryo spacings, (2) larger embryo masses, and (3) up to ~0.4 Earth masses of material left in the form of planetesimals when the gas vanishes. The systems we produce are reasonably stable for ~100 Myr and therefore require an external source to stir up the embryos sufficiently to produce final systems resembling the terrestrial planets. | astro-ph | astro-ph |
Effects of Type I Migration on Terrestrial Planet Formation
Douglas McNeil and Martin Duncan
Department of Physics and Astronomy, Queen's University,
Kingston, Ontario, Canada K7L 3N6
[email protected]
Harold F. Levison
Southwest Research Institute, Boulder, Colorado, USA 23489
ABSTRACT
Planetary embryos embedded in a gas disc suffer a decay in semimajor axis --
type I migration -- due to the asymmetric torques produced by the interior and
exterior wakes raised by the body (Goldreich & Tremaine 1980; Ward 1986). This
presents a challenge for standard oligarchic approaches to forming the terrestrial
planets (Kokubo & Ida 1998) as the timescale to grow the progenitor objects near
1 AU is longer than that for them to decay into the Sun. In this paper we inves-
tigate the middle and late stages of oligarchic growth using both semi-analytic
methods (based upon Thommes et al. 2003) and N-body integrations, and vary
gas properties such as dissipation timescale in different models of the protoplan-
etary disc. We conclude that even for near-nominal migration efficiencies and
gas dissipation timescales of ∼ 1 Myr it is possible to maintain sufficient mass in
the terrestrial region to form Earth and Venus if the disc mass is enhanced by
factors of ∼ 2 − 4 over the minimum mass model. The resulting configurations
differ in several ways from the initial conditions used in previous simulations of
the final stages of terrestrial accretion (e.g. Chambers 2001), chiefly in (1) larger
inter-embryo spacings, (2) larger embryo masses, and (3) up to ∼ 0.4M⊕ of ma-
terial left in the form of planetesimals when the gas vanishes. The systems we
produce are reasonably stable for ∼ 100 Myr and therefore require an external
source to stir up the embryos sufficiently to produce final systems resembling the
terrestrial planets.
Subject headings: planets, formation
-- 2 --
1.
Introduction
A standard model for forming the terrestrial planets has emerged, and is divided broadly
into three stages (see references in Canup et al. 2000). The first stage is poorly understood,
but the formation of the Sun is thought to leave a large but thin protoplanetary disc of
both gas and solids, with coagulation or gravitational instability producing rocky objects
('planetesimals') in the metre-to-km size range. In the second stage, interactions between
the planetesimals result in an early phase of 'runaway growth', to be discussed below, where
the single largest body in a given region breaks away from the mass spectrum and becomes
significantly larger than the second-largest. Over ∼ 105 − 106 years, this gives rise to an
oligarchic phase, also discussed below, which produces many well-spaced comparable-mass
embryos. In the third stage, the embryos interact with each other and what remains of the
the disc, merging and producing the terrestrial planets we observe today on a timescale of
∼ 107 − 108 years.
The basic paradigm for the second stage, the 'planetesimal problem', is described in
Wetherill & Stewart (1993). One considers a very large number of small bodies of compara-
ble mass, moving in low-eccentricity, low-inclination orbits around the Sun. These objects
will suffer close encounters with each other, changing the velocity distribution, and occa-
sionally collide and either merge or fragment, changing the mass spectrum as well. Energy
equipartition in the system decreases the random velocities of the larger bodies, and increases
the velocities of the smaller ones, an effect known as 'dynamical friction'. It has been shown
(Wetherill & Stewart 1989) that dynamical friction leads to runaway growth of the largest
body, as the decrease in velocity increases the collisional cross section and thus the accretion
rate. Later simulations (e.g. Ida & Makino 1992a,b; Kokubo & Ida 1996, 1998) confirmed
this, and demonstrated that the end result of runaway growth was an 'oligarchic' phase where
the accretion process is dominated by the gravitational stirring of well-separated embryos of
roughly equal masses.
All simulations published to date of the second stage have either neglected any interac-
tion between the gas disk and the solid bodies or have included only the aerodynamic gas
drag acting on the solid bodies as described by Adachi et al. (1976). However, it is well
known that the interactions between the protoplanetary disc and the protoplanets can also
cause the latter to migrate. This can happen in several ways: type I, where the migration is
driven by the asymmetry between the torques generated by the interior and exterior wakes
of the body (Goldreich & Tremaine 1980; Ward 1986); type II, where the protoplanet is large
enough to open up a gap in the disc, coupling it to the viscous evolution of the disc (Pa-
paloizou & Lin 1984); and recently type III (Masset & Papaloizou 2003; Artymowicz 2004),
a very fast migration mode which occurs when the embryo begins to migrate quickly enough
-- 3 --
that corotating material cannot librate. (The solid material in the disc can also be scattered
by the embryos, leading to embryo migration directly; this 'type 0' migration (Fernandez &
Ip 1984) is thought to play an important role in the evolution of the outer solar system.)
In most models, protoplanets of terrestrial mass are incapable of opening a gap and
therefore do not suffer type II migration, and likewise type III should not be relevant (Masset
& Papaloizou 2003). Accordingly we restrict ourselves to considering the effects of type
I migration, specifically the tidal interactions between the planet and disc which result in
semimajor axis decay and random velocity decay (both eccentricity and inclination damping.)
In the models we are considering, the typical effect is to decrease the semimajor axis, which
leads to difficulties for accretion: the type I migration rate scales with the mass of the object,
so as an object grows it moves inwards faster. Indeed, the type I migration timescales for
the embryos produced near 1 AU are typically comparable to the timescales for their growth;
this problem is discussed in Tanaka et al. (2002).
In a sense, we seek to refine the initial conditions for the third stage, by including the
effects of the tidal interaction during the second stage. This is very similar in spirit to
Kominami & Ida (2004), but they include only the effects of random velocity damping, not
of semimajor axis decay, and study only the third stage itself. To this end we will investigate
the effects of different assumptions about the disc and the migration efficiency on the state of
the protoplanetary system when the gas vanishes, and thus our ability to produce terrestrial
planets from what material remains. Note that we concentrate here on forming the Earth
and Venus. Mercury has a small enough mass (0.055 M⊕) that it is easily considered debris
at this scale, and the problem of forming Mars has its own special difficulties involving the
asteroid belt, the possible presence of other large embryos, and influences from Jupiter, which
we defer. We will not discuss here the third stage resulting from this work, because it is
probably dependent on the details of the stirring of the embryos near the end of the second
stage. This is likely to involve the perturbations introduced by the formation of Jupiter
and Saturn, for which models are sufficiently uncertain that we reserve a discussion of these
effects for a subsequent paper.
In §2 we develop a semi-analytic model for oligarchy during type I migration in the ter-
restrial region and use the results to motivate the N-body simulations which follow. In §3 we
discuss our numerical N-body methods, and present the results of our N-body simulations in
§4. Section 5 contains a further discussion of our scenario and we summarize our conclusions
in §6.
-- 4 --
2. Models
In this section we construct a model of oligarchic formation by adapting the model of
Thommes et al. (2003), which built on the earlier work of Kokubo & Ida (1998, 2000, 2002).
We content ourselves with summarizing their results and noting our modifications. We then
discuss the differences between the physics in this semi-analytic model and the physics we
use in our N-body simulations.
We use the semi-analytic model for two distinct purposes: first, to provide a quick
way to explore parameter space and determine what regimes are good candidates for more
detailed N-body investigation; and second, to generate initial conditions for said N-body
simulations. This allows us to begin our runs later in the process than would otherwise be
possible, thereby avoiding the expense of handling very large numbers of embryos at early
times.
2.1. Semi-analytic Model
Thommes et al. (2003) represent the oligarchic system by a two-component model con-
sisting of embryos and field planetesimals, where the former accrete but the latter do not.
The model variables are the embryo mass M and planetesimal surface density Σm as func-
tions of semimajor axis a, and we seek to evolve M(a) and Σm(a) over time t.
Note that since M(a) is continuous but is meant to represent a discrete population,
it corresponds not to the mass of material in the form of embryos in a given region, but
to the mass that an embryo would have were one present at that semimajor axis.
It is
also assumed that the embryo separation b measured in single-planet Hill units (defined by
rH = (M/3M⊙)1/3a) is constant and contains all the information about the effects of any
embryo-embryo mergers in the model. This assumption is justified in Kokubo & Ida (1998)
as the consequence of an equilibration between the decrease in b caused by accretion of small
objects and an increase in b caused by two-body scattering followed by eccentricity damping
due to dynamical friction from the field of planetesimals. The properties of the resulting
(non-continuous) embryo population can be determined from M(a) by discretizing the curve
appropriately. Thus we construct an effective embryo surface density ΣM = M/(2π a ∆a),
setting the embryo spacing ∆a = b rH.
The embryos are assumed to have uniform physical density ρM . It is also assumed that
the embryos are kept in near-circular orbits by dynamical friction with the field (or, when
active, tidal damping from the gas) and therefore the important random velocity is that of
the field particles.
-- 5 --
The field population is similarly approximated as being comprised of planetesimals of
uniform mass m and density ρm (and therefore uniform radius rm) with (root-mean-square)
eccentricity em and inclination im satisfying em = 2im. Thommes et al. (2003) determine
the eccentricity, after Ida & Makino (1993), by equating the stirring timescale of the field
by the embryos with the damping timescale due to aerodynamic drag on the field particles.
The stirring timescale in a dispersion-dominated regime can be written as
Tstir ≃
1
40(cid:18)Ω2a3
GM (cid:19)2 e4
m M
ΣM a2Ω
,
(1)
where G is the gravitational constant and Ω the orbital frequency. We use the damping
timescale due to aerodynamic drag from Adachi et al. (1976), where ρgas is the density of
the gas and CD is a drag efficiency factor (≃ 1) dependent upon the Reynolds number of the
disc. We neglect resonant interactions. In our notation,
Taero =
1
e2
m
m
(CD/2) πr2
m ρgas a Ω
.
Equating these two timescales yields an equilibrium eccentricity of
eeq ≃
1.72 m1/15 M 1/3 ρ2/15
gas a1/5 b1/5
M⊙
1/3 C 1/5
D ρ1/5
m
.
Thommes et al. (2003) derive an expression for the growth of the embryo mass
dM
dt
=
3.93 M 1/6
⊙ G1/2 Σm M 2/3 C 2/5
ρ1/3
M a1/10 m2/15 ρ4/15
m
D ρ2/5
gas
,
(2)
(3)
(4)
and a corresponding decrease in planetesimal surface density from conservation of mass
dΣm
dt
=
−M 1/3
⊙
32/3 b πa2 M 1/3
dM
dt
.
(5)
(The aerodynamic drag also affects Σm, as we discuss later.)
We restrict consideration of initial disc conditions at T = 0 to those resembling the
minimum mass model (Hayashi 1981). Specifically, we assume the surface density in solids
has the form Σsolid = k(r/AU)p; the nominal values are k = 7.1 g cm−2, p = −1.5 and
we introduce an enhancement factor fenh so that Σsolid = fenh (7.1 g cm−2) (r/AU)p. We
-- 6 --
adopt a gas-to-solid ratio of 240, so that Σgas = 240 Σsolid. We assume the disc has an
exponential vertical structure such that ρgas(r, z) = ρgas(r) exp(−z2/z2
0), with ρgas(r) taken
at the midplane and z0(r) the disc half-thickness. (Integrating the z-dependent term over
all z gives the relation ρgas = Σgas/√πz0.) The half-thickness varies as z0(r) = Z1(r/AU)5/4,
where we take the thickness at 1 AU to be Z1 = 0.07 AU. We model the dissipation of the gas
by introducing an exponential decay with a characteristic time τdecay which is fixed for a given
model and does not vary through the disc, ρ ∝ exp(−t/τdecay), where we allow τdecay = ∞.
We define a parameter η which measures the degree of pressure support of the disc (and
thus deviations of circular orbits from Keplerian velocity): η = (π/16)(α + β)(z0/a)2 where
α = 5/4 − p and β = 1/2; typically η ≃ 0.001. We assume the gas is in cylindrical rotation
such that vgas = vkep√1 − 2η.
For the planetesimal migration introduced by aerodynamic drag, we adopt the orbit-
averaged approximation valid for small e, i, and η due to Adachi et al. (1976):
vm =
da
dtaero ≃ −2
a
Taero em (cid:18)5
8
e2
m +
1
2
i2
m + η2(cid:19)1/2
nη +(cid:18)α
4
+
5
16(cid:19) e2
m +
1
8
i2
mo.
(6)
We also add a prescription for type I migration based on that of Papaloizou & Larwood
(2000), assuming that the embryos are on circular orbits (see also Tanaka et al. 2002). They
derive a timescale for semimajor axis damping ta:
ta =
1
ca (cid:18) a3
GM⊙(cid:19)1/2(cid:18) Z1
AU(cid:16) a
AU(cid:17)1/4(cid:19)2(cid:18)Σgasπa2
M⊙ (cid:19)−1 (cid:18) M
M⊙(cid:19)−1
,
(7)
where ca is a migration efficiency parameter, with ca = 1 being nominal and ca = 0 yielding
ta = ∞. By varying ca in our simulations we can account for an uncertainty of order unity
in the rate of migration.
For the purpose of this model we can construct a rough radial velocity for protoplanet
material
a
ta
= −
(8)
vM =
da
dttidal
We emphasize that since it is not clear (to choose one assumption among many) whether
the aerodynamic drag coefficient CD should be ≃1 or ≃2 (Adachi et al. 1976), the above
model is only expected to be accurate to within factors of order unity. This is true even if
we neglect consequences of assuming that all field particles have the same fixed mass and
the issue of the evolution of inter-embryo spacing (see §4.4).
τ
d
e
c
a
y
=
∞
.
F
i
g
.
1
.
--
E
v
o
l
u
t
i
o
n
o
f
n
o
m
n
a
l
i
m
o
d
e
l
w
i
t
h
o
u
t
t
i
d
a
l
m
i
g
r
a
t
i
o
n
,
k
=
7
.
1
g
c
m
−
2
,
p
=
−
1
.
5
,
]
E
M
[
s
s
a
m
o
y
r
b
m
e
0
1
.
0
8
0
.
0
6
0
.
0
4
0
.
0
2
0
.
0
0
0
.
0
t = 0 Myr
t = 1 Myr
t = 5 Myr
t = 10 Myr
--
7
--
0.0
0.5
1.0
1.5
2.0
2.5
3.0
semimajor axis [AU]
-- 8 --
2.2. Results from Semi-analytic Model
We now integrate the model described in the previous section forward in time. We have
two variables, the embryo mass M(a) and the planetesimal surface density Σm(a), whose
evolution (in the absence of aerodynamic or tidal migration) is specified by equations (4)
and (5). We introduce migration through equations (6), (7), and (8). These five equations
are simultaneously integrated across a zone from a = 0.2 AU to a = 5.0 AU, subject to our
time-dependent gas profile.
Figure 1 shows the evolution of the standard minimum-mass model (k = 7.1 g cm−2,
fenh = 1.0, p = -1.5, τdecay = ∞) assuming that b = 10. We arbitrarily set M0 = 10−6M⊕
as the seed mass for the embryos at all a, so that the total fraction of mass in embryos is
. 5 × 10−4 that of the mass in the planetesimal field; accretion at early times is so quick
that the precise value is unimportant as long as M0 ≪ Mfinal. As is evident on the left part
of the curves, the embryo mass nearly asymptotes toward a power-law in semimajor axis a:
except for deviations introduced by planetesimal migration and the nonzero initial mass M0
of the embryo, the limiting mass is set solely by the amount of planetesimal material and
the chosen b, Mlim = p8/(3M⊙)(Σm π b a2)3/2. (The right side of the curve will reach the
same asymptote eventually.)
Figure 2 shows the minimum-mass model with tidal migration added (keeping τdecay =
∞): several differences from the previous case are immediately apparent. The maximum
embryo mass is substantially reduced, from 0.086M⊕ at 10 Myr to 0.027M⊕. Without
migration there is a sharp drop-off in embryo mass past the peak of the accretion front, but
with tidal migration there is a wide band in semimajor axis where the embryo mass is within
25% of the maximum. With τdecay = ∞ there is nothing to prevent embryo material from
continuing to migrate inwards as it accretes, and thus eventually all embryo mass falls into
the Sun.
Figure 3 is the result of introducing gas dissipation of τdecay = 1 Myr, which produces an
intermediate system. The maximum embryo mass reached is ∼ 0.043M⊕, down by a factor
of 2 from the no-migration case, but the gas decay has several effects: type I migration is
halted, preserving embryo mass; the increase in planetesimal eccentricity as a result of the
decrease in aerodynamic drag increases the accretion time; and as a result of this increase,
a fair amount of material is left in the form of planetesimals. However, the total amount of
material in the terrestrial region (0.5AU ≤ a ≤ 1.5AU) is insufficient to later accrete into
the Earth and Venus during stage three.
The dissipation timescale of ∼ 1 Myr was chosen based on studies of the disc lifetimes
around stars in young clusters by Haisch et al. 2001. They find that roughly half the stars
F
i
g
.
2
.
--
E
v
o
l
u
t
i
o
n
o
f
n
o
m
n
a
l
i
m
o
d
e
l
w
i
t
h
m
i
g
r
a
t
i
o
n
,
k
=
7
.
1
g
c
m
−
2
,
p
=
−
1
.
5
,
τ
d
e
c
a
y
=
∞
.
]
E
M
[
s
s
a
m
o
y
r
b
m
e
0
1
.
0
8
0
.
0
6
0
.
0
4
0
.
0
2
0
.
0
0
0
.
0
t = 0 Myr
t = 1 Myr
t = 5 Myr
t = 10 Myr
--
9
--
0.0
0.5
1.0
1.5
2.0
2.5
3.0
semimajor axis [AU]
p
=
−
1
.
5
,
τ
d
e
c
a
y
=
1
M
y
r
.
F
i
g
.
3
.
--
E
v
o
l
u
t
i
o
n
o
f
n
o
m
n
a
l
i
m
o
d
e
l
w
i
t
h
m
i
g
r
a
t
i
o
n
a
n
d
d
i
s
s
i
p
a
t
i
o
n
,
k
=
7
.
1
g
c
m
−
2
,
]
E
M
[
s
s
a
m
o
y
r
b
m
e
0
1
.
0
8
0
.
0
6
0
.
0
4
0
.
0
2
0
.
0
0
0
.
0
t = 0 Myr
t = 1 Myr
t = 5 Myr
t = 10 Myr
--
1
0
--
0.0
0.5
1.0
1.5
2.0
2.5
3.0
semimajor axis [AU]
-- 11 --
lose their discs in . 3 Myr, and the overall disc lifetime was ∼ 6 Myr. (In the N-body
simulations to be discussed later, we will consider discs with both larger and greater τdecay.)
This suggests that in order to obtain the appropriate amount of material in the terrestrial
region with nominal migration, one could (1) introduce a dissipating disc and (2) enhance
the disc above the minimum mass model. Figure 4 compares the end state after 100 Myr
of evolution for the minimum mass model with various τdecay values with the same for a
disc enhanced by a factor of three. The enhanced case both moves the point of maximum
embryo mass outwards and raises the maximum mass reached by almost a factor of ∼ 3, but
is otherwise qualitatively similar to the minimum mass results. From this and other runs
(not shown) we conclude that if we believe the nominal migration rates, then to within the
expected reliability of the model, disc enhancements of a factor of several and dissipation
times of ∼1 Myr should suffice to keep enough prototerrestrial mass in the region.
Unfortunately these analytic models are sufficiently simplified that they miss many in-
teresting dynamical interactions and so cannot predict the detailed behaviour of the embryos.
This is especially true at the later stages when the embryo eccentricities are no longer being
damped by the gas. The assumption that b∼ 10 also breaks down (see §4.4) as the embryo
mass increases, and therefore semi-analytic estimates of the mass in embryos remaining at
Myr timescales may be unreliable. We therefore want to construct N-body realizations of
these models and integrate them directly, as we discuss next.
2.3. N-body modifications to Semi-analytic model
For consistency, we carry as much of the above approach as possible through to our
N-body code. There are nevertheless modifications in both the aerodynamic and the tidal
drag formulae.
First, instead of the approximation for the aerodynamic drag in eq. (6), we consider the
planetesimal migration induced by aerodynamic drag
aaero =
dv
dt
=
v − vgas
τaero
,
(9)
where v is the Cartesian velocity of the object, vgas the velocity of the gas at the object's
position, and τaero is given by eq. (2). As before, we assume that vgas = vkep√1 − 2η.
Second, for the tidal drag, we use the full approach of Papaloizou & Larwood (2000),
which the authors developed to handle the case where a protoplanet's eccentricity can be
F
i
g
.
4
.
--
E
v
o
l
u
t
i
o
n
o
f
n
o
m
n
a
l
i
a
n
d
e
n
h
a
n
c
e
d
m
o
d
e
l
s
w
i
t
h
m
i
g
r
a
t
i
o
n
a
n
d
d
i
s
s
i
p
a
t
i
o
n
.
]
E
M
[
s
s
a
m
o
y
r
b
m
e
]
E
M
[
s
s
a
m
o
y
r
b
m
e
0
2
.
0
5
1
.
0
0
1
.
0
5
0
.
0
0
0
.
0
0
2
.
0
5
1
.
0
0
1
.
0
5
0
.
0
0
0
.
0
minimum mass model at T = 100 Myr
t decay = 0.1 Myr
t decay = 1 Myr
t decay = 10 Myr
3x minimum mass model at T = 100 Myr
t decay = 0.1 Myr
t decay = 1 Myr
t decay = 10 Myr
--
1
2
--
0.0
0.5
1.0
1.5
2.0
2.5
3.0
semimajor axis [AU]
-- 13 --
greater than the scale height-to-semimajor axis ratio. They derive timescales for semimajor
axis damping ta and for eccentricity damping te:
ta =
1
GM⊙(cid:19)1/2(cid:18) Z1
M⊙ (cid:19)−1
AU(cid:17)1/4(cid:19)2(cid:18) Σgasπa2
ca (cid:18) a3
M⊙(cid:19)−1 (cid:18) 1 + (e/1.3) (Z1/AU)−5 (a/AU)−5/4
1 − (e/1.1) (Z1/AU)−4 (a/AU)−1 (cid:19) ,
(cid:18) M
AU(cid:16) a
te =
1
GM⊙(cid:19)1/2(cid:18) Z1
ce (cid:18) a3
AU(cid:16) a
M⊙(cid:19)−1 1 +
(cid:18) M
AU(cid:17)1/4(cid:19)4(cid:18)Σgasπa2
AU(cid:19)−3
4 (cid:18) Z1
M⊙ (cid:19)−1
AU(cid:17)−3/4! ,
(cid:16) a
e
(10)
(11)
where (as before) ca is the migration efficiency, and ce is the analogous damping efficiency.
They also argue that if the inclination damping timescale is not significantly shorter
than the eccentricity damping timescale then it plays little role in the equilibrium state; we
set ti = te for simplicity. From these we can find the acceleration on an object due to tidal
damping of semimajor axis and random velocity, namely
atidal = −
v
ta −
2(v·r)
r2 te −
2(v·k)k
ti
,
(12)
where r, v, and a are Cartesian position, velocity, and acceleration vectors, respectively (with
r as the magnitude of the radial vector) and k is the unit vector in the vertical direction.
In the code (to be discussed in §3), it is equations (9) through (12) which are used to
incorporate the interaction between the embryos and planetesimals and the gas disc.
3. N-body Methods
We perform the numerical integrations using a parallel implementation of SyMBA (Dun-
can et al. 1998) called miranda. SyMBA is a mixed-variable symplectic integrator based on
the N-body map of Wisdom & Holman (1991) (see also Kinoshita et al. 1991) which has
been improved to accurately handle close encounters between particles. As is done in the
analytic model, we consider two classes of object, embryos (which gravitate and can accrete
planetesimals) and field planetesimals (where we neglect all self-interactions.) We have made
-- 14 --
several modifications to the original algorithm, mainly to include the effects of the gas disc.
As introduced in §2.3, both embryos and planetesimals suffer the aerodynamic drag of eq. (9)
and the embryos further suffer semimajor axis decay and random velocity damping accord-
ing to the prescriptions of (10), (11), and (12). Accelerations resulting from the gas disc
are treated as operators surrounding the (possibly recursively subdivided) drift operator in
the basic democratic heliocentric step (Duncan et al. 1998) after the manner of Thommes
(2001). However, in practice they are not applied during close encounters as the model for
the gas density is almost certainly inapplicable in such situations, and we want to prevent
artificial formation of tight binaries. Note that we neglect the gravitational potential due to
the gas disc.
One important approximation made in our N-body treatment of the field is the use of
'super-planetesimals'. That is, in representing the field by a discrete number of particles, we
use a large mass msp for each object and imagine it represents an aggregate of underlying
planetesimals of mass m. The aerodynamic drag that the object feels is that which the
underlying m-mass object would, but when a merger occurs the mass msp is used. Thus
the integrated field objects are serving as dynamical tracers. Theoretical considerations
(e.g. Binney & Tremaine 1987) show that the accretion rate of an embryo embedded in a
planetesimal field should depend only the product of the number density of the planetesimals
and their mass, provided that the collision timescale is much shorter than the embryo growth
timescale. Numerical experiments varying the ratio msp/m over orders of magnitude confirm
that the accretion is insensitive to the exact value of msp in our regime given that the
characteristic embryo mass M ≫ msp and the number of super-planetesimals is large.
It remains true that as a consequence of this approximation, at early times when the
ratio of typical embryo mass M to msp is at its lowest, we are inaccurately treating the
effects of dynamical friction. However, the embryos are on circular, non-inclined orbits in
any event due to the strong tidal eccentricity damping, and the aerodynamic drag is damping
the planetesimals' random velocities, which suppresses the frictional stirring. As the drag
becomes less effective due to the dissipation of the disc, the M/msp ratio quickly increases,
thereby reducing the inaccuracy.
We adopt as starting conditions M/msp ≥ 5 (where M is the characteristic initial
embryo mass). Note that this differs from the approximation often made by Kokubo and
Ida where they instead increase the accretion radius of their particles by a factor f , thereby
artificially speeding up their simulations by a factor f −β for some β ∈ [1, 2] (c.f. Kokubo &
Ida 1996; e.g. Kokubo & Ida 2002).
We also assume perfect accretion, which avoids the problem of collisional cascades (and
the resulting increase in particle number, catastrophic in a nonstatistical code) at the cost
-- 15 --
of inaccuracy in the mass spectrum; effects of this and other simplifications are discussed in
§5.
4. Simulations
4.1. Generating initial conditions for the N-body simulations
Since computer power is limited, we wish to begin our integrations as late in the second
stage as possible. We will use the semi-analytic model of §2 to evolve the disc from our nomi-
nal T=0 start until the number of embryos (i.e. the predicted number of embryos; recall that
M(a) is continuous) is computationally tractable. This naturally produces a self-consistent
gradient in embryo masses with the appropriate field population. We then discretize M(a)
and Σm(a), yielding a set of embryos and a large number of super-planetesimal field parti-
cles, typically ∼ 8500 (subject to the constraints on mass discussed in the previous section),
which constitute our N-body realization of the initial conditions for our simulations.
To be specific, we set embryo and planetesimal densities ρM = ρm = 3.0 g cm−3, and
take the underlying physical planetesimal mass to be m = 2.1· 10−6M⊕ (i.e. planetesimal
radius rm = 100 km at 3.0 g cm−3). We set the efficiencies of aerodynamic drag and tidal
eccentricity damping to unity, CD = 1, ce = 1, and set η = 0.001 at 1 AU. We take M0,
the initial embryo mass at T = 0 for the semi-analytic model, to be 0.0015M⊕, which is
much larger than m but two orders of magnitude below the likely final values M ≥ 0.1M⊕.
By introducing this larger M0 we advance the evolution (especially of the embryos at larger
a) forward in time. This procedure is justifiable to the extent that the evolution we are
bypassing is merely local accretion prior to any significant embryonic tidal migration. Thus,
embryos in a simulation starting with smaller embryonic seeds would largely 'catch up' in
mass to those in our simulations by the time significant migration sets in.
We evolve this model until T = 0.3 Myr, which produces a satisfactory (∼ 40 − 50)
number of embryos, and is early enough that the use of a fixed b of 10 is acceptable. In this
early phase, the runaway growth timescale (to return the system to oligarchy if the spacing
is too large) and the scattering timescale (to do the same if the spacing is too small) are
both short. Relaxation to the empirical equilibrium conditions begins almost immediately
in the simulations, and we do not believe that our results are particularly sensitive to the
fine details at the start.
From this T = 0.3 Myr frame, objects are constructed from a0 = 0.75 AU to a1 ≈ 2.5
AU with embryo masses derived from M(a) (objects placed inwards, from 0.50 AU to 0.75
AU, quickly fall off the inner edge of the simulation at 0.40 AU.) This sets the maximum
-- 16 --
mass of a field object msp (see §3) and from Σm(a) the field objects can be built. Initial
eccentricities and inclinations for embryos are set to 0.001 and 0.0005, respectively. Plan-
etesimal eccentricities are drawn from a Rayleigh distribution of scale eeq using eq. (3), and
we enforce i = e/2. (We observe that eq. (3) consistently predicts an equilibrium eccentricity
which is slightly larger than the value to which the simulations rapidly relax. Experiments
suggest that this is due to an overprediction of the stirring, whether the viscous or friction
term, and not an underprediction of the effects of aerodynamic drag. This is consistent with
the comments of Stewart & Ida (2000) on stirring efficiency.) All angles are randomized.
As discussed above, the implications of the simplified models of §2 are that to preserve
sufficient mass we will need to increase the surface density, and therefore (keeping in mind
possible enhancements in the outer solar system) we focus on discs ∼ 3 to 4 times the
minimum mass disc. Specifications of the resulting conditions are summarized in table 4.1.
We use values for τdecay of 0.5, 1.0, and 2.0 Myr, and migration efficiency ca of 0.25, 0.50,
and 1.00. For concreteness we choose the τdecay = 1 Myr, ca = 0.5 case to discuss below.
It was not obvious to what absolute time T the simulations would need to be advanced:
after some experiments we chose a value of T = 20 Myr, when the gas has effectively vanished
in all our simulations and the inter-embryo spacing is large enough that the timescale for
mutual interaction is long. Note that we do not include Jupiter and Saturn in this simulation.
For example, fig. 5 shows the initial (i.e. T = 0.3 Myr) conditions for the N-body
simulation produced by the model for run C2. We see the clear decrease in embryo mass
with increasing a characteristic of early times. In the innermost region, 0.70-0.80 AU, roughly
50% of the material is contained in embryos, and in the outermost region (2.4-2.5 AU) this
proportion drops to less than 5%. (We define femb as the fraction of total mass in a region
which is in the form of embryos.)
The majority of the simulations were run at the Canadian Institute for Theoretical
Astrophyics on the McKenzie parallel machine, a Beowulf cluster of ∼ 256 dual-processor
2.4 GHz Linux boxes, and the remainder were integrated on local machines at Queen's
University.
-- 17 --
Table 1: Simulation Parameters
Name
C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
C11
C12
C13
C14
fenh
3.0
3.0
3.0
3.0
3.0
3.0
3.0
3.0
3.0
4.0
4.0
4.0
4.0
4.0
p
τdecay [Myr]
-1.0
-1.0
-1.0
-1.0
-1.0
-1.0
-1.0
-1.0
-1.0
-1.5
-1.5
-1.5
-1.5
-1.5
1.0
1.0
1.0
2.0
2.0
2.0
0.5
0.5
0.5
1.0
1.0
1.0
2.0
2.0
ca NM Nm
7987
0.25
7987
0.50
1.00
7987
7876
0.25
7876
0.50
1.00
7876
8451
0.25
8451
0.50
1.00
8451
9570
0.25
9570
0.50
1.00
9570
9120
0.25
0.50
9120
47
47
47
47
47
47
48
48
48
45
45
45
44
44
Note. -- fenh and p are defined through Σsolid = (fenh · 7.1 g cm−2)(a/AU)p where Σsolid is the original
surface density in solids at T=0. τdecay is the e-folding time of the gas dissipation, ca is the migration
efficiency factor, and NM and Nm are the embryo and field particle counts. Each field particle is a super-
planetesimal as described in §3.
-- 18 --
4.2. Results
As an example, time slices in (a, M) and (a, e) from C2, a representative run which
demonstrated most of the typical behaviours, are displayed in figures 5 through 10. In this
simulation, Σm = 21.3 g cm−2 (a/AU)−1, ca = 0.5, and τdecay = 1 Myr.
After 0.7 Myr, in figure 6, at T = 1 Myr (1 e-folding time of the gas decay), the system
has undergone considerable evolution. Many embryos have undergone mutual mergers and
some have migrated beyond the inner edge of the simulation and been removed. Figure 6
shows wide ranges of roughly oligarchic behaviour where neighbouring objects have compa-
rable mass. In the region from 0.5 to 1.6 AU, for example, most of the embryos have mass
∼ 0.18M⊕ ± 0.02M⊕. At this time the region where femb ≃ 0.5 extends from 1.0-1.5 AU.
The variation of evolution timescale with semimajor axis is apparent, but both embryo-field
and embryo-embryo mergers have occurred beyond 1.5 AU.
In zones where an embryo has achieved some separation from its neighbours, 'Jacobi
wings' can be formed (see Tanaka & Ida 1997). Jacobi wings are the characteristic config-
urations -- a curved V -- produced in the (a, e) plane by an embryo embedded in a disc of
planetesimals as it scatters them, so called because the shape of the scattering paths of the
field particles are determined by the the conservation of the Jacobi integral in the restricted
circular three-body problem. One such wing is apparent interior to the embryo near 0.7 AU
in fig. 7, where the Jacobi wing has swept up a considerable amount of mass (0.2M⊕) during
migration.
At this time, 0.43M⊕ of material has been removed from the simulation for approaching
the Sun too closely (0.4 AU being the limit). Ultimately 2.1M⊕ will be removed, and 90%
of this is accomplished by 3 Myr. The probable fate of the removed material is discussed
further in §4.5; there are good reasons to believe that most of the material which escapes
makes it to the solar surface. With migration, in order to leave enough mass in the terrestrial
region, it becomes necessary to make a substantial sacrifice to the Sun.
At 2 Myr, in figure 7, when the gas has dropped to ∼ 1/7.4 of its original density, more
objects have been lost to the inner edge, and those that remain from 0.5 to 1.5 AU retain
comparable mass (i.e. the deviation from the mean is less than a factor of two) and have
grown to a mean mass just over 0.2M⊕. By 5 Myr (figure 8) the femb . 0.5 region is now
beyond 2.0 AU and we are approaching the end of the transition phase and the beginning
of our end stage (where embryo-planetesimal accretion no longer plays a significant role.)
The mean embryo mass in the 0.5-1.5 AU region is ∼ 0.25M⊕, brought down somewhat by
the objects from 1.3 to 1.4 AU. At this particular time the embryo spacing from 0.5 to 0.8
AU is anomalously small, and this tightly-packed configuration quickly evolves to a wider
-- 19 --
one. At this time, after 5 e-foldings, the gas density has dropped to ∼ 1/150 of its original
value and gas drag (whether aerodynamic or tidal) is not important. The corresponding
increase in eccentricity pushes the accretion timescale up substantially, and the dominant
growth mechanism will now be embryo-embryo mergers.
Between 10 Myr and 20 Myr (figures 9 and 10) we arrive at a long-lived configuration
of embryos: only one embryo merger occurs between 0.5 AU and 1.5 AU, and it occurs
before 14 Myr. Evolution is continuing beyond 1.75 AU, and although limited, the remaining
planetesimals are gradually being consumed. The mean embryo mass in the region of interest
(0.5-1.5 AU) is ∼ 0.45M⊕, and the spacing varies from b ∼ 15 to b ∼ 25, which is very well-
separated. The eccentricities of the bodies (∼ 0.02 − 0.04), which have been increasing over
time, are now many times above their tidally-suppressed values. The region contains 2.58M⊕
in the form of embryos and a residual planetesimal field of 0.23M⊕ for a total of 2.81M⊕ in
the zone from 0.5-1.5 AU.
As previously noted, roughly 2M⊕ of material escapes the simulation to the inside.
A substantial amount of mass remains beyond the zone of interest, however: 3.55M⊕ of
material is located beyond 1.5 AU. This is problematic, as there is nothing approaching a
4M⊕ object between the Earth and Mars; we return to this problem in §4.5.
Viewing the embryo evolution in (M, a, t) space is instructive, as shown in fig. 11. Three
different regimes are apparent: for a < 1.0 AU the embryos head directly for the Sun without
scattering; for 1.0 AU < a < 1.5 AU the embryos 'slide' (i.e. collectively move in an orderly
fashion, without orbit crossing) toward their final destinations, slowing as the gas decay
lowers the migration rate; and for a > 1.5 AU the evolution is highly chaotic. Figure 12
magnifies the innermost regions.
At early times, when the embryo masses are low, the embryos can reorder themselves
substantially and move several tenths of an AU without suffering a merger with another
embryo. However, as time passes and the masses of the embryos grow (whether through
accretion of planetesimals or direct embryo mergers), the migration pulls objects inwards
at different rates. This leads to embryo segregration, and in some cases (typically where
migration rates are high) to the formation of groups of two to four members which we call
tidal convoys.
(Although not mentioned explicitly, this effect also appears to have been
present in the simulations of Papaloizou & Larwood 2000; see figure 6 in their paper.)
At any time in these convoys the objects are typically ordered such that the mass
increases outwards. This is self-selection: if an inner object is more massive than an outer
one, differential migration will cause the objects to diverge, breaking the convoy. As one
would expect, the migration rate of the convoys is above that of the less massive objects and
-- 20 --
run C2 : time = 3.00e+05 [yr]
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
y
t
i
c
i
r
t
n
e
c
c
e
0
.
1
8
.
0
6
.
0
4
.
0
2
0
.
0
0
.
0
1
0
.
8
0
0
.
6
0
0
.
4
0
0
.
2
0
0
.
0
0
.
0
0.5
1.0
1.5
2.0
2.5
semimajor axis [AU]
Fig. 5. -- Initial conditions at T = 0.3 Myr for model C2. Large circles correspond to
embryos (of area proportional to mass) and small circles to planetesimals; the lines through
the large circles indicate a width of 10 Hill radii. The broken lines on the top graph indicate
the amount of planetesimal material (in Earth masses) in a semimajor axis bin of width 0.1
AU.
-- 21 --
run C2 : time = 1.00e+06 [yr]
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
y
t
i
c
i
r
t
n
e
c
c
e
0
.
1
8
.
0
6
.
0
4
0
.
2
0
.
0
0
.
0
1
0
.
8
0
0
.
6
0
0
.
4
0
0
.
2
0
.
0
0
0
.
0
0.5
1.0
1.5
2.0
2.5
semimajor axis [AU]
Fig. 6. -- Configuration at T = 1 Myr for model C2.
-- 22 --
run C2 : time = 2.00e+06 [yr]
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
y
t
i
c
i
r
t
n
e
c
c
e
0
.
1
8
.
0
6
.
0
4
0
.
2
0
.
0
0
.
0
1
0
.
8
0
0
.
6
0
0
.
4
0
0
.
2
0
.
0
0
0
.
0
0.5
1.0
1.5
2.0
2.5
semimajor axis [AU]
Fig. 7. -- Configuration at T = 2 Myr for model C2.
-- 23 --
run C2 : time = 5.00e+06 [yr]
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
y
t
i
c
i
r
t
n
e
c
c
e
0
.
1
8
.
0
6
.
0
4
0
.
2
0
.
0
0
.
0
1
0
.
8
0
0
.
6
0
0
.
4
0
0
.
2
0
.
0
0
0
.
0
0.5
1.0
1.5
2.0
2.5
semimajor axis [AU]
Fig. 8. -- Configuration at T = 5 Myr for model C2.
-- 24 --
run C2 : time = 1.00e+07 [yr]
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
y
t
i
c
i
r
t
n
e
c
c
e
0
.
1
8
.
0
6
.
0
4
0
.
2
0
.
0
0
.
0
1
0
.
8
0
0
.
6
0
0
.
4
0
0
.
2
0
.
0
0
0
.
0
0.5
1.0
1.5
2.0
2.5
semimajor axis [AU]
Fig. 9. -- Configuration at T = 10 Myr for model C2.
-- 25 --
run C2 : time = 2.00e+07 [yr]
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
y
t
i
c
i
r
t
n
e
c
c
e
0
.
1
8
.
0
6
.
0
4
0
.
2
0
.
0
0
.
0
1
0
.
8
0
0
.
6
0
0
.
4
0
0
.
2
0
.
0
0
0
.
0
0.5
1.0
1.5
2.0
2.5
semimajor axis [AU]
Fig. 10. -- Configuration at T = 20 Myr for model C2.
-- 26 --
run C2 : overview
0.1 ME
0.2 ME
0.3 ME
0.4 ME
0.5 ME
0.6 ME
0.7 ME
0
.
3
5
.
2
0
.
2
5
.
1
0
.
1
5
.
0
i
j
s
x
a
r
o
a
m
m
e
s
i
0
5
10
15
20
time [Myr]
Fig. 11. -- Evolution of embryos with time in C2. The area of the circles scales linearly with
embryo mass as indicated in the legend. For each embryo three lines are drawn, corresponding
to the osculating values of the perihelion, semimajor axis, and aphelion.
-- 27 --
run C2 : convoying and sliding regions
0
1
.
9
.
0
8
.
0
7
.
0
6
.
0
5
.
0
4
.
0
i
j
s
x
a
r
o
a
m
m
e
s
i
0.5
1.0
1.5
2.0
2.5
3.0
time [Myr]
Fig. 12. -- The interior regions of C2 at early times.
-- 28 --
below that of the most massive object as the outer objects 'push' on the inner ones. In many
of our simulations, as in C2, these groupings are only apparent on the inner edges (where the
object masses and gas density are highest) as the decay of the gas disc and corresponding
weakening of the migration makes the convoys harder to form and maintain.
If we set
τdecay = ∞, tidal convoying becomes nearly inevitable, which raises serious questions about
the applicability of analytic oligarchy models to the middle-to-late stage transition unless
the gas is presumed to have vanished.
In our simulations, while the gas remains, type I migration is remarkably efficient at
capturing embryos into first-order mean motion resonances with each other, and the convoys
usually show resonant relationships (two-object pairs are especially susceptible.) This en-
hancement of resonance capture probability due to migration has been discussed in a different
context by previous authors (see Peale & Lee 2002 for a possible explanation for the Laplace
resonance among the Galilean satellites.) Most two-member tidal convoys are observed to
be stable, but arguments to that end (e.g. Gladman 1993) are not easily generalized to the
multiple-planet case, and in practice convoys involving three or more objects of different
masses are eventually observed to merge.
Beyond the convoying region, fig. 12 shows part of the 'sliding' region (and, conveniently,
a pair showing intermediate behaviour.) As is apparent in fig. 11, when the masses and gas
densities are such that the spacings will stay roughly constant during the migration, a group
of embryos can move smoothly from further out in the disc suffering only a few mutual
mergers. (Admittedly, in C2 the surface density of the gas disc varied as Σ ∝ a−1, and one
can show (see section 4.5) that da/dt is constant for a fixed mass in this case. However, in
the sliding region the masses vary by a factor of several, the gas density was decaying, and
similar behaviours are seen for the Σ ∝ a−3/2 case. The effect is not a peculiarity of this
particular density law, although it is a consequence of the near-constant migration rates that
it and similar laws provide.)
Finally, in the lowest-mass region, chaotic behaviour persists for the entire length of
the simulation, although it does begin to settle down by the end. In one case, the embryo
excursion reached 1.3 AU. Even in this chaotic regime, the broad outlines of the oligarchic
model (roughly equally spaced embryos of comparable mass) were still respected, despite the
fact that the simplest oligarchic picture is inapplicable as the embryos are being constantly
reordered. This may imply that other mechanisms underly the oligarchic model's impressive
robustness (Goldreich et al. 2004).
These three regimes (convoying, sliding, and chaotic) were observed throughout the
simulations -- although their locations and strengths varied with mass, gas density, and
stochastically-set spacing, as expected -- and are likely frequent.
-- 29 --
4.3. Final Characteristics
Time slices of the final configurations for simulations C1-C14 are plotted in figures 13,
14, and 15.
The resulting mass in the 0.5-1.5 AU region in the form of embryos ranges from 0.81M⊕
to 3.27M⊕, and for planetesimals from 0.14M⊕ to 0.44M⊕. Some of the runs (C2, C3, C4,
C6) yield very oligarchic-looking outcomes where the variation in mass between embryos
is low, and the characteristic mass varies from 0.2 to 0.4 M⊕. Others (C1, C7, C8, C9,
C10) produce systems with embryos of mass approaching 1M⊕, at least some of which are
candidate planets. The remaining simulations are difficult to classify, although most show
regions oligarchic in appearance with a few outliers (e.g. C12, C14.) C5 produced the least
mass (embryo mass 0.81M⊕, field mass 0.44M⊕) for reasons to be discussed in §4.6, and C1
the most (3.27M⊕, 0.16M⊕). By accident, C1 produced two dominant objects of terrestrial
mass -- Venus and Earth analogues.
Figures 16 and 17 show the resulting amounts of embryo and field mass for all simulations
at 1, 10, and 20 Myr. Most simulations show an increase in embryo mass with time, and
a substantial number of simulations landed in our nominal target region of 1.5 to 2.5M⊕ in
embryo mass and ≤ 0.5M⊕ in planetesimal mass.
The original expectation was for all the simulations to produce oligarchic-looking results
(as in C3), not direct planet production (as in C1), but it is not surprising that with more
massive embryos at least some embryo-embryo mergers would occur as the gas vanished,
and it takes only a few such mergers at late times to produce a substantial planet. Of
the fourteen simulations, four produced less than two Earth masses (our nominal target) of
material in the region and the remaining ten produced more, leading one to suspect that
our enhancements may have been too high. However, this depends entirely on how much
of the material is removed during whatever processes turn the resulting configurations into
terrestrial systems: they are quite widely spaced.
4.4. Evolution of inter-embryo spacing
As noted by Goldreich et al. (2004), the argument offered in Kokubo & Ida (1998) to
derive an estimate for the inter-embryo spacing b as a function of M and Σ is of dubious
applicability. Recall that the spacing b in their model is derived by equating the decrease in b
due to embryo growth (due to planetesimal accretion) with an increase in b due to scattering.
The two-body scattering formula of Petit & Henon (1986) has a problem in the oligarchic
context: namely, there are embryos on both sides of our scattering pair who are instead
-- 30 --
MT = 3.27
mT = 0.16
C1
MT = 2.58
mT = 0.23
C2
MT = 2.34
mT = 0.23
C3
MT = 2.57
mT = 0.14
C4
MT = 0.81
mT = 0.44
C5
MT = 1.54
mT = 0.14
C6
2
.
1
0
.
1
8
.
0
6
.
0
4
.
0
2
.
0
0
.
0
2
1
.
0
1
.
8
0
.
6
0
.
4
0
.
2
0
.
0
0
.
2
1
.
0
1
.
8
0
.
6
0
.
4
0
.
2
.
0
0
.
0
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
2
.
1
0
.
1
8
.
0
6
.
0
4
.
0
2
.
0
0
.
0
2
1
.
0
1
.
8
0
.
6
0
.
4
0
.
2
0
.
0
0
.
2
1
.
0
1
.
8
0
.
6
0
.
4
0
.
2
.
0
0
.
0
0.50
0.70
0.90
1.10
1.50
0.50
1.30
semimajor axis [AU]
0.70
0.90
1.10
1.30
1.50
Fig. 13. -- Final configurations in mass and semimajor axis for runs C1-C6. MT and mT
indicate (in Earth masses) the total mass in embryos and in planetesimals, respectively, in
the region [0.5 AU, 1.5 AU].
-- 31 --
MT = 2.92
mT = 0.34
C7
MT = 2.77
mT = 0.35
C8
MT = 2.45
mT = 0.45
C9
MT = 2.24
mT = 0.21
C10
0
1
.
8
.
0
6
0
.
4
.
0
2
.
0
0
.
0
0
.
1
8
.
0
6
.
0
4
.
0
2
0
.
0
0
.
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
0
1
.
8
.
0
6
0
.
4
.
0
2
.
0
0
.
0
0
.
1
8
.
0
6
.
0
4
.
0
2
0
.
0
0
.
0.50
0.90
1.30
0.50
0.90
1.30
semimajor axis [AU]
Fig. 14. -- Same as fig.13, but for runs C7-C10.
-- 32 --
MT = 1.32
mT = 0.29
C11
MT = 2.35
mT = 0.31
C12
MT = 1.31
mT = 0.28
C13
MT = 2.02
mT = 0.35
C14
0
1
.
8
.
0
6
0
.
4
.
0
2
.
0
0
.
0
0
.
1
8
.
0
6
.
0
4
.
0
2
0
.
0
0
.
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
0
1
.
8
.
0
6
0
.
4
.
0
2
.
0
0
.
0
0
.
1
8
.
0
6
.
0
4
.
0
2
0
.
0
0
.
0.50
0.90
1.30
0.50
0.90
1.30
semimajor axis [AU]
Fig. 15. -- Same as fig.13, but for runs C11-C14.
-- 33 --
C1
C7
C1 C2
C4
C2
C3
C3
C6
C6
C5
]
E
M
[
s
s
a
m
o
y
r
b
m
e
0
5
3
.
5
2
3
.
0
0
3
.
5
7
.
2
0
5
.
2
5
2
.
2
0
0
.
2
5
7
.
1
0
5
.
1
5
2
.
1
0
0
.
1
5
7
.
0
0
5
.
0
5
2
.
0
0
0
.
0
C7
t = 1 Myr
t = 10 Myr
t = 20 Myr
C4
C2
C5
C4
C1
C7
C6
C3
C5
0.00
0.25
0.50
0.75
1.00
1.25
1.50
1.75
2.00
planetesimal mass [ME]
Fig. 16. -- Embryo mass versus planetesimal mass within the 0.50-1.50 region at 1, 10, and
20 Myr for runs C1-C7. The nominal target zone (from 1.5 to 2.5 M⊕ in embryo mass and
≤ 0.5M⊕ in planetesimal mass) is shaded.
-- 34 --
t = 1 Myr
t = 10 Myr
t = 20 Myr
C8
C9
C10
C12
C10
C8
C9
C14
C12
C14
C11
C13
C11
C13
C10
C13
C8
C9
C14
C11
C12
]
E
M
[
s
s
a
m
o
y
r
b
m
e
0
5
3
.
5
2
3
.
0
0
3
.
5
7
.
2
0
5
.
2
5
2
.
2
0
0
.
2
5
7
.
1
0
5
.
1
5
2
.
1
0
0
.
1
5
7
.
0
0
5
.
0
5
2
.
0
0
0
.
0
0.00
0.25
0.50
0.75
1.00
1.25
1.50
1.75
2.00
planetesimal mass [ME]
Fig. 17. -- Same as fig.16, but for runs C8-C14.
-- 35 --
trying to scatter the two objects closer together. Naively one would expect this to turn the b
growth from a monotonic process into a much slower random walk. The formula also assumes
that the eccentricities of the two objects are near zero at conjunction, an assumption which
breaks down in our model as the damping timescale increases with the dissipation of the gas.
(Simple experiments confirm both suspicions: the multiple-embryo case shows substantially
slower b growth than the two-embryo case, and introducing a decay in the strength of the
eccentricity damping in the two-embryo case also washes out the growth.) Finally, when
type I migration is active, differential migration of equal-mass objects will tend to increase
b, and this is a non-negligible effect in our regime.
Nevertheless, our numerical experiments suggest that b ≃ 10 is a reasonable approxi-
mation during the period we apply the semi-analytic model, possibly accidentally: embryos
quickly scatter when b < 5 and growth is a weak function of M (b ∝ M 2/15 in Kokubo
& Ida 1998) so substantial masses are required before b = 10 becomes unacceptable.
In
any case, corrections to the 'wrong' chosen b occur on short timescales, and during testing
no significant correlations between initial b (granted b ∼ 10) and the results were observed,
suggesting that the equilibration process leaves little signature.
Figure 18 shows the evolution of the mean mass-weighted b in the region a = 0.50 −
1.50AU for all simulations. Although the specific evolution varies from integration to in-
tegration, the trend of increase in b from ∼ 10 to ∼ 20 is clear. The trend to large b is
accentuated by the tidal migration: in our control runs without tidal migration (not shown)
we find values of b∼ 10 − 15 were common in the late stages. The spikes near 4 and 6 Myr
are the result of unusual structures being formed; the 4 Myr spike in run C5 which reaches
off the scale is to be further discussed in §4.6. One concern is that we find spacings of more
than 20 Hill radii are typically sufficient to produce stable systems (at least on timescales of
100 Myr, the expected formation timescale). Preliminary simulations suggest that the intro-
duction of Jupiter and Saturn stirs the system enough to produce embryo interaction, and
resonance sweeping involving the removal of the disc potential (here neglected: see Nagasawa
et al. 2003) is another possibility. We will return to this issue in a subsequent paper.
4.5. Material Transport
The gas drag (both aerodynamic and tidal) tends to drive embryo and field material
inwards. Scattering amongst embryos, or between embryos and field objects, can counteract
this and move material outwards. The question of how the transport processes affect the
resulting proportions of mass in the final embryos therefore arises (i.e. what amount of
mass, from where.) Figures 19 and 20 show the resulting material fraction by source region
-- 36 --
5
3
0
3
5
2
0
2
5
1
0
1
]
b
[
i
g
n
c
a
p
s
n
a
e
m
0
5
10
time [Myr]
15
20
Fig. 18. -- Evolution of mean inter-embryo spacing measured in units of single-planet Hill
radii in region 0.5-1.5 AU.
-- 37 --
(whether embryo or planetesimal.)
Substantial amounts of the material which ends up in the 0.5-1.5 AU region comes from
beyond 1.5 AU, and even from beyond 2 AU. In some rare cases (e.g. the embryo at 1.2
AU in C1, or at 0.87 AU in C8), an embryo in fact consists mostly of material from beyond
2 AU. It is not surprising there are general trends with τdecay. Indeed, in fig. 19, we see
in runs C4, C5, and C6 (each with τdecay = 2 Myr), more material from beyond 2 AU is
incorporated than in C1-C3 (with decay timescale 1 Myr.) A somewhat weaker trend with
migration efficiency ca is also evident. Comparatively little material originally from 1 AU
becomes part of embryos beyond 1.2 AU. We do not believe that our lack of embryos and
planetesimals beyond 2.5 AU at T = 0.3 Myr plays a significant role in these results: in the
majority of the resulting embryos, material from beyond 2.25 AU contributes little, and the
arguable exception of C6 is the simulation with the most migration (the result of a 2 Myr
dissipation timescale and an efficiency ca = 1).
The large degree of inward mixing suggests that present-epoch local chemistry may
be probing much further out in the planetary disc than might have been expected without
migration, at least if countervailing effects such as disc turbulence are limited. Consequences
of different models of material transport during terrestrial planet formation on presently
observed water abundances are discussed in Lunine et al. (2003) and Raymond et al. (2004).
Recently Weidenschilling, in unpublished work, has also been investigating terrestrial
formation in enhanced discs with migration using a hybrid dynamical and statistical tech-
nique and reached broadly consistent conclusions (private communication, DPS 2004.) How-
ever, he reports a definite tendency for material to accumulate in the inner regions.
As we remove objects when they get below 0.4 AU, and do not simulate any objects
inside of 0.75 AU at our start of 0.3 Myr, we are insensitive to the production of a terrestrial-
mass Mercury. This decision was motivated by our early experiments, which demonstrated
that the oligarchic predictions of the §2 model were reliable (except for certain details of the
b spacing) over a large mass range before type I migration played a role and were tolerable,
although less accurate, at tidally significant masses. The resulting embryos were massive
enough that they invariably migrated inside of 0.4 AU before the gas vanished. Although
these early embryos escaped our region of interest, this is not a guarantee that they would
succeed in migrating all the way to the Sun. To investigate the likely future of embryos
whose dynamics we stopped tracing, we can take the objects removed for falling off the inner
edge of our simulation and integrate the tidal migration equations (8) and (7). For Σ ∝ a−1,
then da/dt ∝ exp(−t/τdecay), and for Σ ∝ a−3/2, da/dt ∝ a−1/2 exp(−t/τdecay).
Figure 21 shows the projected evolution of the escaped embryos, considered one at a
c
o
r
r
e
s
p
o
n
d
n
g
i
t
o
s
o
u
r
c
e
r
e
g
i
o
n
s
a
s
i
n
d
i
c
a
t
e
d
o
n
t
h
e
l
e
g
e
n
d
.
F
i
g
.
1
9
.
--
M
a
t
e
r
i
a
l
f
r
a
c
t
i
o
n
s
f
o
r
C
1
-
C
6
.
P
i
e
s
c
o
r
r
e
s
p
o
n
d
t
o
e
m
b
r
y
o
s
a
t
2
0
M
y
r
w
i
t
h
s
l
i
c
e
s
l
n
o
i
t
a
u
m
S
i
C6
C5
C4
C3
C2
C1
Material Transport
Source
[AU]
2.5
2.0
1.5
1.0
--
3
8
--
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
1.3
1.4
1.5
semimajor axis [AU]
F
i
g
.
2
0
.
--
M
a
t
e
r
i
a
l
f
r
a
c
t
i
o
n
s
f
o
r
C
7
-
C
1
4
.
l
n
o
i
t
a
u
m
S
i
C14
C13
C12
C11
C10
C9
C8
C7
Material Transport
Source
[AU]
2.5
2.0
1.5
1.0
--
3
9
--
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
1.3
1.4
1.5
semimajor axis [AU]
-- 40 --
time. The majority of objects (80%) succeed in making it to the solar radius (∼ 0.005
AU), and so our removal of them from the simulation is defensible.
In over half of the
simulations, no embryo which was removed would have survived. In several runs (C1, C2,
C4, C9, C10, C11), multiple embryos would have survived, although in only four of these
(C2, C4, C10, C11) would an embryo have survived inside of 0.2 AU. Note this analysis is
restricted to the embryos which were actually integrated and then removed, not to embryos
which should have been there but never were (e.g. embryos initially located at 0.5 AU). A
fortiori most of the embryos we did not include inside of 0.75 AU would have also made it
to the surface of the Sun, at least for the 1 Myr and 2 Myr decay timescales. In the 0.5
Myr case, with its resulting lower migration, although we are not missing much material
which should end up in the 0.5-1.5 AU region, we could be missing a substantial amount
between 0.2 and 0.4 AU by our choice of initial embryo range. As for planetesimals, given
the effectiveness of migrating Jacobi wings as a sweeping mechanism, it seems likely that
many of the planetesimals not incorporated into the interior embryos would move with the
flow, and given the strong dependence on semimajor axis of formation timescale there are
unlikely to be many left at 0.4 AU when tidally-significant embryos are emerging at 1 AU
and beyond.
In summary, for τdecay ≥ 1 Myr, we expect that we are missing only a few
embryos, and do not believe our choice of inner edge significantly affects our conclusions.
The outer regions present their own set of difficulties: an average of 3M⊕ of material is
left beyond 1.5 AU (recall that the initial outermost embryo in each simulation is located
at ∼ 2.5 AU), which is an order of magnitude above the mass of Mars (0.11M⊕). As is
evident from the discussion of §4.3, we have provided an upper bound on the necessary disc
enhancement to survive type I migration, as our enhancements consistently produce more
mass than desired. Decreasing the enhancement to reduce this overproduction will mitigate
this problem somewhat, but even a one-third decrease will still leave ∼ 2M⊕ of material.
This problem is not unique to our migration models, however: the naive minimum mass
model, Σsolid = 7.1 g cm−2(r/AU)−1.5, itself produces ∼ 1.1M⊕ of material between 1.5 and
2.5 AU. The majority of the unwanted material in this region in our simulations is simply a
reflection of how much was there to begin with. One can deal with this problem in several
ways: assume a primordial decrease in original surface density in the region between Earth
and Mars; increase the material transport rates in some fashion, e.g. collisional grinding
producing large amounts of dust which leaves the region very quickly due to aerodynamic
drag; and so on. Any approach which solves the problem for non-enhanced models without
migration will have a natural analogue in our enhanced models with migration, and therefore
the problem is no worse in our model than for the standard scenario.
Of course, it is unlikely that a simple power-law profile accurately describes the density
-- 41 --
post−removal evolution
C1
C2
C4
C9
C10
C11
other
4
.
0
3
0
.
2
0
.
1
0
.
0
.
0
i
s
x
a
j
r
o
a
m
m
e
s
i
0.5
1.0
2.0
5.0
10.0
time [Myr]
Fig. 21. -- Projected future evolution of embryos removed from the simulation.
-- 42 --
of the protoplanetary disc (either solids or gas) to within a few radii of the Sun. If the solid
density reaches a local maximum near 1 AU, it may be possible to avoid any interior and
exterior problems entirely.
4.6. Rings
An unexpected side effect of neglecting the self-interactions of the field particles became
apparent during some of the simulations. Occasionally, due to an early embryo-embryo
merger, an object is produced which is prematurely massive relative to its neighbours. The
increase in mass results in a higher migration rate than the next outer embryo, and a gap
opens up between them where the only objects are planetesimals. This region devoid of
embryos can also be created by a major scattering event. In a weak sense this process occurs
during every merger, but most of the time the gap is quickly filled by another embryo,
whether through migration or scattering. However, under certain circumstances, this gap is
not closed quickly with the entrance of an embryo from further out in the disc, but instead
persists. In the absence of embryos there is nothing to stir the field, and the aerodynamic
drag decreases the field eccentricity, resulting in a thin ring of planetesimals (except where
an exterior embryo arrives to form a new Jacobi wing.) In extreme cases the width of the
ring is substantial (e.g. 0.2 AU), corresponding to a large amount of mass (∼ 1.0M⊕). At this
point it can serve as a wall, and act as a barrier to the entry of inward-migrating embryos
from beyond the ring.
There are several possibilities for the resulting behaviour. Sometimes the ring is quickly
disrupted, especially if it is small. At other times the ring endures until an exterior embryo
pushes through. This often leads to a surprisingly quick transition of the embryo through
the ring, a consequence of the process of forming Jacobi wings. The ring may or may not
survive this embryo crossing. Sometimes the ring endures, and the evolution stalls as the
embryo masses reached exterior to the ring are insufficient to push through the barrier.
An example of an extreme scenario (from run C5) is demonstrated in figure 22, the
most massive (by a factor of several) and longest-lived (also by a factor of several) ring in
our simulations. At 2 Myr, the system is unremarkable: between 1.0 and 1.7 AU there are
four embryos separated by ∼ 0.1 AU, with planetesimals throughout. By 2.5 Myr, however,
scattering events have resulted in a region from 1.3 AU to 1.6 AU without any embryos, but
with almost an Earth mass of planetesimal material. After another half million years, the
innermost embryos have migrated out of the region, and only one embryo of ∼ 0.2M⊕ remains
to control the dynamics. (This wide embryo-free region is responsible for the highest spike
in b in fig. 18.) Over the next 8 Myr, the ring moves from ∼ 1.5 AU to ∼ 1.25 AU as a result
-- 43 --
of both aerodynamic drag (over the first several million years, at any rate; the dissipation
timescale in C5 is 2 Myr) and the more significant push given by the exterior embryo. During
this push, the embryo consumes enough of the ring during the Jacobi scattering to grow from
0.2 to 0.3M⊕. Finally, an encounter between the shepherd embryo and and its neighbour
scatters the shepherd into the ring by 11.5 Myr. When the embryo enters, it quickly migrates
through and disrupts the ring. By 12 Myr the ring is gone.
We emphasize that this is an extreme case, and that when the rings do not form or are
quickly disrupted by the embryos (before field interactions would be expected to do so) the
two-component embryo/field approximation we make remains acceptable.
Could these rings be physical, or are they merely artifacts of our simplified treatment
of the planetesimal field? It is true that some early mergers and scattering events are
going to occur, and differential migration will then attempt to produce a gap. However,
in a system with a more realistic mass spectrum (with objects of intermediate mass) the
second-tier objects which are not accreted during the migration of the overmassive embryo
might ascend to be the new oligarchs in the absence of competition. Furthermore, eventually
mutual planetesimal-planetesimal encounters (which we neglect) should produce new seed
objects for oligarchic growth within the ring. Rings which last for less than the timescale
for internal disruption seem possible, and may occur in real planetary systems. In this case,
their chief effect would seem to be slowing the migration process for embryos with exterior
semimajor axes.
5. Discussion
Various authors, looking with dismay at the unhappy consequences of type I migration
for forming giant cores, have concluded that the effect simply does not occur. Some recent
work (Menou & Goodman 2004) has also argued that the timescale for migration in some
circumstances may be substantially larger than we have assumed here. This may ultimately
prove correct, but several points are worth making in response. First, given the complexity
of the local gas physics near the embryo, it is probably premature to rule out near-nominal
migration; a clear understanding of the behaviour will need to wait for the next generation
of hydrodynamic simulations. For example,
it is entirely possible that whatever effects
hypothetically shut off type I migration before the opening of a gap only become significant
for embryo masses near or beyond Earth mass. If so, the intuition that type I is migration
is too destructive to have occurred in our system could be completely justified regarding
the giant planets but not relevant for terrestrial formation. Disc properties such as opacity
on which migration may sensitively depend are extremely poorly constrained, and saying
-- 44 --
t = 2.0 Myr
t = 11.0 Myr
t = 2.5 Myr
t = 11.5 Myr
t = 3.0 Myr
t = 12.0 Myr
.
2
1
0
1
.
8
0
.
.
6
0
4
.
0
2
0
.
0
.
0
2
.
1
0
.
1
8
.
0
6
.
0
4
.
0
2
.
0
0
.
0
2
.
1
0
.
1
8
.
0
6
.
0
4
.
0
2
.
0
0
0
.
y
t
i
c
i
r
t
n
e
c
c
e
2
1
.
0
8
0
0
.
4
0
.
0
0
0
.
0
2
1
.
0
8
0
.
0
4
0
.
0
0
0
.
0
2
1
.
0
8
0
.
0
4
0
.
0
0
0
0
.
]
s
e
s
s
a
m
h
t
r
a
E
[
s
s
a
m
d
e
l
i
f
d
e
n
n
b
i
1.00
1.20
1.40
1.60
1.00
1.20
1.40
1.60
semimajor axis [AU]
Fig. 22. -- Extreme example of the formation of a planetesimal ring. The ring endures from
2.5 Myr to 11.0 Myr in run C5. The left axis shows the eccentricity of the objects (dots)
while the right axis shows the mass of planetesimals in bins of width 0.1 AU (solid lines).
-- 45 --
anything firm will require improved proplyd observations.
That said, it seems possible that the gas disc is far less smooth than we have assumed,
and type I migration may be halted by abrupt changes in gradient, opacity, or density at
unknown locations. For example, Matsuyama et al. (2003) develops models for the pho-
toevaporation of the disc which result in a zone of depleted gas which could protect the
terrestrial region from marauding proto-Jovian embryos. Although these specific models
have dynamical consequences which make them difficult to accept at face value, the possi-
bility of significant disc inhomogeneities causing similar effects remains. Unfortunately all of
this will clearly be very difficult to model. Disc turbulence can wash out the effect of tidal
migration on short timescales (Laughlin et al. 2004; Nelson & Papaloizou 2004) but may not
remove a net drift; we hope to study oligarchy in a turbulent disc in future work.
Several compromises made to complete the integrations with the available resources may
affect the results. The neglect of field self-interaction contributes to the formation of rings,
although to the degree that the behaviour of the system truly is oligarchic, the embryo-
field model should be a reasonable approximation. Nevertheless, our representation of the
mass spectrum is limited, and no field object can promote itself to an embryo. We have
chosen 100 km as the underlying planetesimal size, which is both plausible (the formation
timescale for such objects is comparable to the absolute time when we begin the simulation)
and conventional in terrestrial formation simulations. Smaller planetesimal particles, such as
those produced in a fragmentation spectrum, would lead to different accretion rates (Rafikov
2004) in potentially complicated ways, and would also be likely to result in higher migration
rates for planetesimals due to aerodynamic drag, possibly requiring greater disc enhancement.
The artificial increase of the seed mass for the embryos may also cause problems. It
changes the gradient in embryo mass, and means that at a given time, the outermost embryos
are more massive relative to their inner neighbours than they should be. When the evolution
is local, this is defensible, but during migration it will lead to an overestimation of the
transport. The overestimation should be mild in practice as the outermost objects are sub-
migratory at T = 0.3 Myr, when our N-body simulations begin.
6. Conclusion
We have investigated the likely conditions at the beginning of the terrestrial endgame
as a result of type I migration during the mid-to-late transition. For reasonable values of the
parameters we find that plausible progenitors for the terrestrial planets can survive despite
type I migration at near the nominal rate, provided the disc density is enhanced above the
-- 46 --
minimum model.
In particular, in order to get the right amount of material left in the
terrestrial region with a gas dissipation timescale of ∼ 1 Myr, one must work with discs
several times more massive than the minimum model, although a lower enhancement than
we used here would likely suffice. (Preliminary investigations to be reported elsewhere using
discs of ∼ 2 times the minimum mass model are encouraging, and show good agreement
with expectations.) As a consequence, the resulting systems at the beginning of the late
phase show significant differences from the conditions normally considered (e.g. Chambers
& Wetherill 1998; Chambers 2001): (1) there is often a large amount, ∼ 0.40M⊕ of material
remaining in the form of planetesimals (supporting the direction taken in Chambers 2001)
at the time of dissipation; (2) the embryos tend to have larger masses, with mode 0.4M⊕;
and (3) the embryos are separated by ∼ 20 − 25 single-planet Hill radii.
In a forthcoming paper we take the next step and study what must be done to turn our
new endgame conditions into terrestrial planets. As our final configurations are stable when
left alone, we will require the introduction of new events (e.g. the formation of Jupiter and
Saturn as in Kominami & Ida 2004) -- to complete the terrestrial construction project.
The authors thank the sponsoring agencies of the McKenzie project, the Canadian In-
stitute for Theoretical Astrophysics, the Canada Foundation for Innovation, and the Ontario
Innovation Trust. The authors are very grateful for the generous allocations of time granted
to the project by the McKenzie external users program at CITA, without which this re-
search would not have been possible. The authors acknowledge useful discussions with Ed
Thommes and Paul Wiegert. DSM is grateful for the assistance of Robin Humble and Chris
Loken at CITA. MJD gratefully acknowledges ongoing support from the Natural Science and
Engineering Research Council of Canada. HFL acknowledges support from NASA's TPF
Fundamental Research Program.
-- 47 --
REFERENCES
Adachi, I., Hayashi, C., & Nakazawa, K. 1976, Progress of Theoretical Physics, 56, 1756
Artymowicz, P. 2004, Debris Disks and the Formation of Planets: A Symposium in Memory
of Fred Gillett, ASP Conference Series 324
Binney, J., & Tremaine, S. 1987, Princeton, NJ, Princeton University Press
Canup, R. M., Righter, K., & et al. 2000, Origin of the earth and moon, edited by R.M. Canup
and K. Righter and 69 collaborating authors. Tucson : University of Arizona Press ;
Houston : Lunar and Planetary Institute, c2000. (The University of Arizona space
science series)
Chambers, J. E. 2001, Icarus, 152, 205
Chambers, J. E., & Wetherill, G. W. 1998, Icarus, 136, 304
Duncan, M. J., Levison, H. F., & Lee, M. H. 1998, AJ, 116, 2067
Fernandez, J. A., & Ip, W.-H. 1984, Icarus, 58, 109
Gladman, B. 1993, Icarus, 106, 247
Goldreich, P., & Tremaine, S. 1980, ApJ, 241, 425
Goldreich, P., Lithwick, Y., & Sari, R. 2004, ARA&A, 42, 549
Haisch, K. E., Lada, E. A., & Lada, C. J. 2001, ApJ, 553, L153
Hayashi, C. 1981, Progress of Theoretical Physics Supplement, 70, 35
Ida, S., & Makino, J. 1992, Icarus, 96, 107
Ida, S., & Makino, J. 1992, Icarus, 98, 28
Ida, S., & Makino, J. 1993, Icarus, 106, 210
Kinoshita, H., Yoshida, H., & Nakai, H. 1991, Celestial Mechanics and Dynamical Astron-
omy, 50, 59
Kokubo, E., & Ida, S. 1996, Icarus, 123, 180
Kokubo, E., & Ida, S. 1998, Icarus, 131, 171
Kokubo, E., & Ida, S. 2000, Icarus, 143, 15
-- 48 --
Kokubo, E., & Ida, S. 2002, ApJ, 581, 666
Kominami, J., & Ida, S. 2004, Icarus, 167, 231
Laughlin, G., Steinacker, A., & Adams, F. C. 2004, ApJ, 608, 489
Lunine, J. I., Chambers, J., Morbidelli, A., & Leshin, L. A. 2003, Icarus, 165, 1
Masset, F. S., & Papaloizou, J. C. B., ApJ, 588, 494
Matsuyama, I., Johnstone, D., & Murray, N. 2003, ApJ, 585, L143
Menou, K., & Goodman, J. 2004, ApJ, 606, 520
Nagasawa, M., Lin, D. N. C., & Ida, S. 2003, ApJ, 586, 1374
Nelson, R. P., & Papaloizou, J. C. B. 2004, Astronomical Society of the Pacific Conference
Series, 321, 367
Papaloizou, J. C. B., & Larwood, J. D. 2000, MNRAS, 315, 823
Papaloizou, J., & Lin, D. N. C. 1984, ApJ, 285, 818
Peale, S. J., & Lee, M. H. 2002, Science, 298, 593
Petit, J.-M., & Henon, M. 1986, Icarus, 66, 536
Rafikov, R. R. 2004, AJ, 128, 1348
Raymond, S. N., Quinn, T., & Lunine, J. I. 2004, Icarus, 168, 1
Stewart, G. R., & Ida, S. 2000, Icarus, 143, 28
Tanaka, H., & Ida, S. 1997, Icarus, 125, 302
Tanaka, H., Takeuchi, T., & Ward, W. R. 2002, ApJ, 565, 1257
Thommes, E. W. 2001, Ph.D. Thesis
Thommes, E. W., Duncan, M. J., & Levison, H. F. 2003, Icarus, 161, 431
Ward, W. R. 1986, Icarus, 67, 164
Wetherill, G. W., & Stewart, G. R. 1989, Icarus, 77, 330
Wetherill, G. W., & Stewart, G. R. 1993, Icarus, 106, 190
-- 49 --
Wisdom, J., & Holman, M. 1991, AJ, 102, 1528
This preprint was prepared with the AAS LATEX macros v5.0.
|
0712.1122 | 1 | 0712 | 2007-12-07T11:56:26 | Variation in the primary and reprocessed radiation from an orbiting spot around a black hole | [
"astro-ph"
] | We study light curves and spectra (equivalent widths of the iron line and some other spectral characteristics) which arise by reflection on the surface of an accretion disc, following its illumination by a primary off-axis source - an X-ray 'flare', assumed to be a point-like source just above the accretion disc resulting in a spot with radius dr/r<1. We consider General Relativity effects (energy shifts, light bending, time delays) near a rotating black hole, and we find them all important, including the light bending and delay amplification due to the spot motion. For some sets of parameters the reflected flux exceeds the flux from the primary component. We show that the orbit-induced variations of the equivalent width with respect to its mean value can be as high as 30% for the observer's inclination of 30 degrees, and much more at higher inclinations. We calculate the ratio of the reflected flux to the primary flux and the hardness ratio which we find to vary significantly with the spot phase mainly for small orbital radii. This offers the chance to estimate the lower limit of the black hole spin if the flare arises close to the black hole. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 10 (2007)
Printed 29 October 2018
(MN LATEX style file v2.2)
Variation in the primary and reprocessed radiation from
an orbiting spot around a black hole
M. Dovciak,1 V. Karas,1 G. Matt2 and R.W. Goosmann1
1 Astronomical Institute, Academy of Sciences of the Czech Republic, Bocn´ı II, CZ-140 31 Prague, Czech Republic
2 Dipartimento di Fisica, Universit`a degli Studi "Roma Tre", Via della Vasca Navale 84, I-00146 Roma, Italy
Accepted ... 2007, Received ... 2007
ABSTRACT
We study light curves and spectra (equivalent widths of the iron line and some other
spectral characteristics) which arise by reflection on the surface of an accretion disc,
following its illumination by a primary off-axis source -- an X-ray 'flare', assumed
to be a point-like source just above the accretion disc resulting in a spot with radius
∆r/r . 1. We consider General Relativity effects (energy shifts, light bending, time
delays) near a rotating black hole, and we find them all important, including the light
bending and delay amplification due to the spot motion. For some sets of parameters
the observed reflected flux exceeds the observed flux from the primary component.
We show that the orbit-induced variations of the equivalent width with respect to its
mean value can be as high as 30% for an observer's inclination of 30◦, and much more
at higher inclinations. We calculate the ratio of the reflected flux to the primary flux
and the hardness ratio which we find to vary significantly with the spot phase mainly
for small orbital radii. This offers the chance to estimate the lower limit of the black
hole spin if the flare arises close to the black hole.
Key words:
line: profiles -- relativity -- galaxies: active -- X-rays: galaxies
1
INTRODUCTION
X-ray spectral measurements of the iron line and the un-
derlying continuum provide a powerful tool to study accre-
tion discs in active galactic nuclei (AGN) and Galactic black
holes (for a review see Fabian et al. 2000; Reynolds & Nowak
2003). If a line originates by reflection of the primary con-
tinuum, then its observed characteristics may reveal rapid
orbital motion and light bending near the central black hole.
Spectral characteristics can be employed to constrain the
black hole mass and angular momentum. A particularly im-
portant role is played by the equivalent width (EW), which
reflects the intensity of the line versus the continuum flux as
well as the role of General Relativity effects in the source.
In order to reduce the ambiguity of results one needs to per-
form spectral fitting with self-consistent models of both the
line and continuum.
Some AGN are known to exhibit EW greater than ex-
pected for a "classical" accretion disc. Enhanced values for
the EW can be obtained by assuming an anisotropical dis-
tribution of the primary X-rays (Ghisellini et al. 1991), sig-
nificant ionization of the disc matter (Matt, Fabian & Ross
1993) or iron overabundance (George & Fabian 1991). Mar-
tocchia & Matt (1996) and Martocchia et al. (2000) found,
using an axisymmetric lamp-post scheme, an anticorrelation
between the intensity of the reflection features and the pri-
mary flux. When the primary source is at a low height on
the disc axis, the EW can be increased by up to an order of
magnitude with respect to calculations neglecting General
Relativity effects. When allowing the source to be located
off the axis of rotation, an even stronger enhancement is ex-
pected (Dabrowski & Lasenby 2001). Miniutti et al. (2003)
and Miniutti & Fabian (2004) have realised that this so-
called light bending model can naturally explain the puz-
zling behaviour of the iron line of MCG -- 6-30-15, when the
line saturates at a certain flux level and then its EW starts
decreasing as the continuum flux increases further. Clear
understanding of the interplay between the primary and the
reprocessed components is therefore highly desirable.
In our previous paper (Dovciak et al. 2004a), we have
proposed that the orbiting spot model could explain the
origin of transient narrow lines, which have been reported
in some AGN X-ray spectra (Turner et al. 2002; Guainazzi
2003; Yaqoob et al. 2003) and widely discussed since then.
The main purpose of the current paper is to present accu-
rate computations of time-dependent EWs and other spec-
tral characteristics within the framework of the spot model,
taking into account a consistent scheme for the local spec-
trum reprocessing. The main difference from previous papers
is that the current one connects the primary source power-
2 M. Dovciak, V. Karas, G. Matt and R.W. Goosmann
law continuum with the reprocessed spectral features. Both
components are further modified by relativistic effects as the
signal propagates towards an observer.
In Section 2 we describe the model and we summarize
the equations and the approximation used. In Section 3 we
present the results of our calculations. The final conclusions
are drawn in Section 4.
2 MODEL SET-UP AND EQUATIONS
2.1 Model approximations and limitations
We examine a system composed by a black hole, an accretion
disc and a co-rotating flare with the spot underneath (Collin
et al. 2003; see Fig. 1). The gravitational field is described
in terms of Kerr metric (Misner, Thorne & Wheeler 1973).
Both static Schwarzschild and rotating Kerr black holes are
considered. The co-rotating Keplerian accretion disc is ge-
ometrically thin and optically thick, therefore we take into
account only photons coming from the equatorial plane di-
rectly to the observer. We further assume that the matter
in the accretion disc is cold and neutral.
A flare is supposed to arise in the disc corona due to a
magnetic reconnection event (e.g. Galeev, Rosner & Vaiana
1979; Poutanen & Fabian 1999; Merloni & Fabian 2001; Cz-
erny et al. 2004). Details of the formation of the flare and its
structure are not the subject of the present paper instead we
assume that the flare is an isotropic stationary point source
with a power-law spectrum. It is located very near above
the disc surface and it co-rotates with the accretion disc.
We also assume that the single flare dominates the intrinsic
emission for a certain period of time.
The question of the flare height above the accretion
disc is still unresolved. Although there are some similari-
ties between the reconnection events that are responsible
for solar flares and those in accretion discs (Romanova et
al. 1998; Czerny et al. 2004), it is not yet clear to what
height the magnetic loops can rise in the latter case. For ex-
ample, Dabrowski & Lasenby (2001) assumed a large height
(h(r) ≃ 1 -- 2Rg) of the flare above the inner disc region and
they performed the ray tracing from the flare towards the
disc surface and further to the observer. On the other hand,
here we assume that the flare height is less than the grav-
itational radius, which seems to be substantiated by the
condition of equipartition between the magnetic and gas
pressure, but the confirmation will need coupled radiation-
magnetohydrodynamic computations, which have been only
recently started (e.g. Blaes et al. 2006). The small height al-
lows us to simplify the calculations of the disc irradiation by
neglecting the light-bending on primary rays, so that we can
study off-axis flares and manage the time-dependent evolu-
tion of the observed flux, EW and other characteristics.
The spot represents the flare-illuminated part of the
disc surface. The flare can be viewed as a lamp. The repro-
cessed photons are re-emitted from only those parts of the
disc that are illuminated by the flare. Therefore the spot
does not share the differential rotation with the disc mate-
rial. It is considered to be a rigid two dimensional circular
feature, with its centre directly below the flare. However,
the matter in the disc lit by the flare is in Keplerian orbits
at the corresponding radii, thus it has different velocities at
to the
observer
accretion
disc
flare
spot
black hole
Figure 1. A sketch of the model geometry (not to scale). A
localized flare occurs above the disc, possibly due to magnetic
reconnection, and creates a spot by illuminating the disc surface.
The resulting 'hot spot' co-rotates with the disc and contributes
to the final observed signal by reprocessing the primary X-rays.
different parts of the spot. This is important when calcu-
lating the transfer function for the observer in the infinity.
Because the flare is very close to the disc, the spot does not
extend far from below the flare. Only photons emitted into
the lower hemisphere reach the disc. We assume the open-
ing angle of the illuminating cone to be 89◦ which means
we lose less than 2% of these photons. We neglect the pho-
tons emitted into the upper hemisphere directed towards the
black hole that could possibly reach the disc behind it. We
approximate the photon trajectories between the flare and
the spot by straight lines. We do not consider the energy
shift and abberation due to the different motion of the flare
and the illuminated disc matter. As far as the illumination
is concerned we consider the flare and the spot to be in the
same co-moving reference frame. Thus the illumination of
the spot is approximated by a simple cosine law in the lo-
cal Keplerian frame co-moving with the matter in the spot
and decreasing with the distance from the flare (i.e. illumi-
i /h2, where µi is the cosine of the incident
nating flux ∼ µ3
angle in the local Keplerian frame and h is the height of the
flare above the disc). Furthermore we neglect the time delay
between the photon's emission from the flare and its later
re-emission from the spot.
The intrinsic (local) spectra from the spot were com-
puted by Monte Carlo simulations considering multiple
Compton scattering and iron line fluorescence in a cold, neu-
tral, constant density slab with solar iron abundance. We
used the NOAR code for these computations (see Section 5
of Dumont et al. 2000 and Chapter 5 of Goosmann 2006).
The local flux depends on the local incident and local emis-
sion angles, hence the flux changes across the spot. Here and
elsewhere in the text we refer to the quantities measured in
the local frame co-moving with the matter in the disc as
"local".
The local flux consists of only two components -- the
flux from the primary source (the flare) and the reflected
flux from the spot. The latter one consists of the reflection
continuum (with the Compton hump and the iron edge as
the main features) and the neutral Kα and Kβ iron lines.
No other emission is taken into account.
As far as the photon trajectories from the spot to the
observer are concerned, all general relativistic effects -- en-
ergy shift, aberration, light bending, lensing and relative
time delays -- are taken into account. We assume that only
the gravity of the central black hole influences photons on
their travel from the disc to the observer. This allows us
to define a relatively simple scheme in which different in-
tervening effects remain under full control and can be well
identified.
2.2 Model predictions for the observed flux
The observed energy flux from the finite size spot on the
accretion disc, F , can be computed from the local energy
flux, f (Dovciak 2004; Dovciak et. al. 2004b),
F (E, t) = ZΣ
dS f (E/g; r, ϕ; te) G(r, ϕ),
where dS = r dr dϕ is an area element on the disc,
te(t, r, ϕ) = t − δt(r, ϕ)
(1)
(2)
is the emission time, δt is a relative time delay with which
photons emitted at different places on the disc reach the ob-
server (we use Boyer-Lindquist spheroidal coordinates; Mis-
ner et al. 1973). Note that we use the specific energy flux (not
the photon flux) per unit solid angle (not per unit area), i.e.
F ≡ E dN (E)/(dt dE dΩ). Hence the units are keV/s/keV
for spectra (keV/s for integrated energy flux in light curves).
One can get the flux per unit area of the detector by dividing
our results by the distance of the source squared.
In a classical (non-relativistic) case, when there is no
light bending and aberration, the relative time delay would
be δt(r, ϕ) = r sin θ sin ϕ (the observer is located in the di-
rection ϕ = −90◦). The transfer function (see Cunningham
1975) for an extended source is G = g2 l µe, where g denotes
the combined gravitational and Doppler shift, l is lensing
and µe is the cosine of the local emission angle at the disc
(measured with respect to the frame co-moving with the
disc). All these functions depend on the place of the emission
(r and ϕ on the disc). Because the local flux is coming from
an orbiting spot it is useful to use coordinates co-moving
with the spot's centre. The observed flux is then
F (E, t) = ZΣ
dSe f (E/g; r, ϕ; te) G(r, ϕ) kt(r, ϕ),
(3)
Spectral variations from an orbiting spot
3
flare, which we suppose is a point-like source
Fp(E, t) = f (E/g; te) Gp(r, ϕ) kt(r, ϕ).
(6)
Here, Gp = g3 l is the transfer function for a point-like source
and the emission time te together with the azimuthal coordi-
nate ϕ are both functions of the observer's time t as defined
in eqs. (2) and (4) with ϕe = 0.
2.3 The intrinsic emission of the flare and the
spot
We define the primary emission to be
fp(E) ≡ E
dNp(E)
dt dE dΩp
= E1−Γ,
where Γ is a photon number density power-law index.
The energy flux of reflected photons is
fr(E, µi, µe, Φe − Φi) ≡
E dNr(E)
dt dE dS⊥dΩe
= nr E
µ3
i
h2
1
µe
,
where
nr(E, µi, µe, Φe − Φi) =
dNr(E)
dt dE dΩpdΩe
.
(7)
(8)
(9)
is the photon number density flux of the reflected radia-
tion emitted into the solid angle dΩe if the incident light
rays come from the solid angle dΩp. In eq. (8) we used the
fact that the solid angle dΩp corresponds to the area dS⊥
perpendicular to the light ray emitted from the spot in the
following way:
dS⊥ =
dS⊥
dS
dS
dSi⊥
dSi⊥
dΩp
dΩp = µe
h2
µ3
i
dΩp.
(10)
Here, h is the height of the primary source above the disc,
dS is the area of the spot lit by the emission coming from
the solid angle dΩp, dSi⊥ is the corresponding area perpen-
dicular to the incident light ray and µi is the cosine of the
incident angle. All these quantities are evaluated in the local
reference frame co-moving with the disc.
As mentioned earlier, the emitted photon flux nr was
calculated by Monte Carlo simulations. In our computations
we used pre-calculated tables of nr(E, µi, µe) which were av-
eraged over the difference between incident and emitted az-
imuthal angles Φ = Φe − Φi.
where dSe = r dr dϕe with
ϕe = ϕ − ωe [t − δt(r, ϕ)].
2.4 The equivalent width, ratio of reflected and
(4)
primary fluxes, hardness ratio
This equation defines ϕ in eq. (3) as a function of ϕe and
t. The angular velocity ωe is the angular velocity of the
spot. The factor kt arises from the coordinate transforma-
tion ϕ → ϕe (it is connected with the time delays). This is
due to the spot moving in its orbit towards or away from
the observer and due to the fact that the light rays emitted
from different places on the disc are differently bent, thus
acquiring different travel times. Therefore we will refer to
this factor as the delay amplification,
kt(r, ϕ) = »1 + ωe
∂(δt)
∂ϕ
(r, ϕ) -- −1
.
(5)
Because the flare is very near above the disc the spot receives
the light emitted downward to almost the whole half-space.
The local equivalent width of the spectral line is
EWloc(µe) =
=
r (EL) + fp(EL)
r (E)
R dSµeR dE f L
R dSµef C
R dE ¯nL
r (E, µe) E
r (EL, µe)EL + E1−Γ
¯nC
,
/2π
(11)
L
where we defined an average of both the line (nL
tinuum (nC
source into the half-space as
r ) and con-
r ) part of the local photon flux emitted by a point
Similarly as for the spot emission one gets an expression
for the observed energy flux from the primary source, the
¯nr(E, µe) ≡
1
2π Z 2π
0
dΩp nr(E, µi, µe, Φ)
4 M. Dovciak, V. Karas, G. Matt and R.W. Goosmann
]
V
e
[
c
o
l
W
E
1100
1000
900
800
700
120
100
80
60
40
20
]
V
e
[
c
o
l
W
E
0
0.2
0.4
0.6
0.8
1
µ
e
0
0
0.2
0.4
0.6
0.8
1
µ
e
Figure 2. Left: The local equivalent width without taking the
primary flux into account as a function of the direction of emis-
sion. Right: The same as in the left panel but with the flux from
the primary source included.
0.12
0.1
0.08
0.06
0.04
0.02
)
l
a
c
o
l
(
p
F
/
r
F
0
0
0.2
0.4
0.6
0.8
1
µ
e
)
l
a
c
o
l
(
s
F
/
h
F
0.62
0.61
0.6
0.59
0
0.2
0.4
µ
e
0.6
0.8
1
Figure 3. Left: The ratio of the locally emitted energy flux in
the direction µe to the primary flux. The fluxes are integrated in
the energy range 3 -- 10 keV. Right: The local hardness ratio of
the fluxes in the ranges 6.5 -- 10 keV (Fh) and 3 -- 6.5 keV (Fs).
=
1
2π Z 1
0
dµiZ 2π
0
dΦ nr(E, µi, µe, Φ) .
(12)
The dependence of the local equivalent width on the emis-
sion angle is shown in Fig. 2.
In the relativistic case the equivalent width is
energy shift g and the overall amplification Gkt change only
slightly throughout the spot and can be represented by their
values at the centre of the spot, gs and Gsks
t, respectively.
The local ratio of the reflected emission to the primary
radiation within the energy range hE1, E2i is
Fr
Fploc
dE fr(E)
(E1, E2, µe) = R dSµeR E2
2π(2 − Γ)R E2
R E2
=
E1
E1
E2−Γ
E1
dE fp(E)
2 − E2−Γ
1
dE ¯nr(E, µe)E
. (15)
The dependence of the ratio of the reflected and primary
radiation on the emission angle is shown in the Fig. 3. In
the relativistic case we find
Fr
Fp
(E1, E2, t) = R dSeGktR E2
t R E2
2π(2 − Γ)R E2
pks
Gs
≈
E1
dE fr(E/g)
E1
dE fp(E/gs)
dE ¯nr(E/g, µe)E
E1
2 − E2−Γ
1
)
gΓ(E2−Γ
Fr
kr(E1, E2, t)
=
g(t)Γ
Fploc
with the coefficient kr(E1, E2, t) being
kr(E1, E2, t) ≡ R E2
R E2
E1
E1
dE ¯nr(E/g(t), µe(t))E
dE ¯nr(E, µe(t))E
(E1, E2, µe(t))
(16)
.
(17)
Again, the approximation in eq. (16) holds true only when
the spot is small.
The last property we will discuss is the hardness ratio.
The dependence of the local hardness ratio, i.e. the ratio of
the hard flux Fh in the energy range hE2, E3i to the soft flux
Fs in the energy range hE1, E2i,
(µe) = k
Fh
Fs loc
Fr
Fr
Fploc
Fploc
(E2, E3, µe) + 1
(E1, E2, µe) + 1
,
(18)
EW(t) =
≈
R dSe GktgR dE f L
r (Ec/g) + Gs
r (E)
pks
t fp(Ec/gs)
R dSe Gkt f C
gR dE ¯nL
r (E, µe) E
r (EL, µe)EL + E1−Γ
¯nC
L
/2π
= g(t) EWloc(µe(t)) .
on the emission angle is shown in Fig. 3. Note that the pri-
mary and both components of the reflected spectrum, line
as well as continuum, are taken into account. The k factor
is defined as the ratio of the hard and soft primary fluxes,
(13)
k =
Fp(E2, E3)
Fp(E1, E2)
.
(19)
A factor g in the numerator accounts for integration of the
line flux over local energy. Note that the amplification func-
tion for the primary flux is taken at the spot centre (denoted
by index 's'), whereas this function for the line and reflected
continuum emission is changing throughout the spot. In the
transformation from coordinate area dSe to the local solid
angle dΩp we used the fact that the local area will differ from
the coordinate one by the g-factor. The centroid energy in
eq. (13) is defined as
In the relativistic case we get
Fh
Fs
(t) = k
Fr
Fp
Fr
Fp
(E2, E3, t) + 1
(E1, E2, t) + 1
≈ k
kr(E2, E3, t) g(t)−Γ Fr
kr(E1, E2, t) g(t)−Γ Fr
Fploc
Fploc
(E2, E3, µe(t)) + 1
(E1, E2, µe(t)) + 1
.
(20)
Ec(t) = R dE F L
R dE F L
r (E)E
r (E)
≈ g(t)EL
(14)
3 RESULTS
with F L
r (E) being the observed energy flux in the line. The
approximations in eqs. (13) and (14) hold only for a very
small "effective" spot (by which we mean a patch of the
disc from which photons arrive simultaneously; the observed
spot shape is deformed by the time delays). In that case the
3.1 Effects of General Relativity
The transfer function and delay amplification tell us how
much the local flux from the flare and the spot is amplified
when observed at infinity. Both of these functions depend
on the location on the disc where the observed photons are
a = 0 , r = 7
a = 0 , r = 7
a = 0 , r = 7
a = 0 , r = 7
Spectral variations from an orbiting spot
5
60
30
t
e
ω
−
ϕ
0
-30
-60
g
1.5
1.0
0.5
0.0
0.0
0.2
0.4
0.6
0.8
1.0
0.0
0.2
0.4
0.6
0.8
1.0
t/T
a = 0.998 , r = 3
t/T
a = 0.998 , r = 3
60
30
t
e
ω
−
ϕ
0
-30
-60
g
1.5
1.0
0.5
0.0
0.0
0.2
0.4
0.6
0.8
1.0
0.0
0.2
0.4
0.6
0.8
1.0
l
l
8
7
6
5
4
3
2
1
0
0.0
8
7
6
5
4
3
2
1
0
0.0
e
µ
1.0
0.8
0.6
0.4
0.2
0.0
0.2
0.4
0.6
0.8
1.0
0.0
0.2
0.4
0.6
0.8
1.0
t/T
a = 0.998 , r = 3
t/T
a = 0.998 , r = 3
e
µ
1.0
0.8
0.6
0.4
0.2
0.0
0.2
0.4
0.6
0.8
1.0
0.0
0.2
0.4
0.6
0.8
1.0
t/T
t/T
t/T
t/T
Figure 4. Several important functions changing within one orbital timescale T for the Schwarzschild (top panels) and the Kerr (bottom
panels) black hole. The radius of the orbit is at 7 GM/c2 for the Schwarzschild black hole and 3 GM/c2 for the Kerr black hole. The
initial time corresponds to the detection of the first photon. The flare and the spot are initially moving from the observer (ϕ = 0◦). The
dashed, dotted and solid lines correspond to the inclination of the observer 30◦, 60◦ and 85◦. Left: The lag (gain) angle −ωeδt due to
the positive (negative) time delay δt with which photons arrive to the observer. The solid and dotted lines correspond to the relativistic
and classical cases, respectively. The lines for 30◦, 60◦ and 85◦ are shown in the order of an increasing amplitude. Middle left: The
relativistic energy shift. Middle right: The lensing. Right: The cosine of the emission angle. The cosine of the observer's inclination
angle is also shown by the straight lines.
a = 0 , r = 7
a = 0 , r = 7
a = 0 , r = 7
3.0
2.5
2.0
t
k
1.5
1.0
0.5
0.0
0.0
0.2
0.4
0.6
0.8
1.0
t/T
a = 0.998 , r = 3
3.0
2.5
2.0
t
k
1.5
1.0
0.5
0.0
0.0
0.2
0.4
0.6
0.8
1.0
t/T
t
k
G
t
k
G
8
7
6
5
4
3
2
1
0
0.0
8
7
6
5
4
3
2
1
0
0.0
0.2
0.4
0.6
0.8
1.0
t/T
a = 0.998 , r = 3
0.2
0.4
0.6
0.8
1.0
t/T
t
k
p
G
t
k
p
G
12
10
8
6
4
2
0
0.0
12
10
8
6
4
2
0
0.0
0.2
0.4
0.6
0.8
1.0
t/T
a = 0.998 , r = 3
0.2
0.4
0.6
0.8
1.0
t/T
Figure 5. Left: Function kt describing the time delay amplification. Middle: The overall amplification of the local specific energy flux
for an extended source of emission. Right: The overall amplification of the local specific energy flux for a point source of emission.
6 M. Dovciak, V. Karas, G. Matt and R.W. Goosmann
]
1
-
s
×
V
e
k
[
x
u
F
l
]
1
-
s
×
V
e
k
[
x
u
F
l
100
10
1
0.1
0.01
0.001
100
10
1
0.1
0.01
0.001
a=0 , r = 7 , θ
o = 30°
lines
reflected
primary
0
0.2
0.4
0.6
0.8
1
t/T
a=0.998 , r = 3 , θ
o = 30°
lines
reflected
primary
0
0.2
0.4
0.6
0.8
1
t/T
]
1
-
s
×
V
e
k
[
x
u
F
l
]
1
-
s
×
V
e
k
[
x
u
F
l
100
10
1
0.1
0.01
0.001
100
10
1
0.1
0.01
0.001
a=0 , r = 7 , θ
o = 60°
lines
reflected
primary
0
0.2
0.4
0.6
0.8
1
t/T
a=0.998 , r = 3 , θ
o = 60°
lines
reflected
primary
0
0.2
0.4
0.6
0.8
1
t/T
]
1
-
s
×
V
e
k
[
x
u
F
l
]
1
-
s
×
V
e
k
[
x
u
F
l
100
10
1
0.1
0.01
0.001
100
10
1
0.1
0.01
0.001
a=0 , r = 7 , θ
o = 85°
lines
reflected
primary
0
0.2
0.4
0.6
0.8
1
t/T
a=0.998 , r = 3 , θ
o = 85°
lines
reflected
primary
0
0.2
0.4
0.6
0.8
1
t/T
Figure 6. The light curves of the observed emission from the flare and the spot for the energy range 3 -- 10 keV for the Schwarzschild
(top) and the Kerr (bottom) black hole and observer's inclination angles 30◦, 60◦ and 85◦ (from left to right). The primary emission,
spot's continuum emission and spot's emission in Kα and Kβ lines are denoted by solid, dashed and dotted graphs, respectively.
]
1
-
V
e
k
×
1
-
s
×
V
e
k
[
x
u
F
l
a=0 , r = 7 , θ
o = 30°
line+reflected
reflected
primary
1
0.1
0.01
a=0 , r = 7 , θ
o = 60°
line+reflected
reflected
primary
1
0.1
0.01
]
1
-
V
e
k
×
1
-
s
×
V
e
k
[
x
u
F
l
a=0 , r = 7 , θ
o = 85°
line+reflected
reflected
primary
1
0.1
0.01
]
1
-
V
e
k
×
1
-
s
×
V
e
k
[
x
u
F
l
0.001
3
6
9
12
E[keV]
0.001
3
6
9
12
E[keV]
0.001
3
6
9
12
E[keV]
a=0.998 , r = 3 , θ
o = 30°
line+reflected
reflected
primary
1
0.1
0.01
]
1
-
V
e
k
×
1
-
s
×
V
e
k
[
x
u
F
l
a=0.998 , r = 3 , θ
o = 60°
line+reflected
reflected
primary
1
0.1
0.01
a=0.998 , r = 3 , θ
o = 85°
line+reflected
reflected
primary
1
0.1
0.01
]
1
-
V
e
k
×
1
-
s
×
V
e
k
[
x
u
F
l
]
1
-
V
e
k
×
1
-
s
×
V
e
k
[
x
u
F
l
0.001
3
6
9
12
E[keV]
0.001
3
6
9
12
E[keV]
0.001
3
6
9
12
E[keV]
Figure 7. The observed spectra averaged over one orbit computed for the same set of parameters as in Fig. 6. Here, the observed line
flux is shown on top of the spot continuum emission.
a=0 , r = 7
a=0 , r = 7
a=0 , r = 7
a=0 , r = 7
Spectral variations from an orbiting spot
7
200
160
120
80
40
0
0
0.2
0.4
0.6
0.8
1
t/T
0.3
p
F
0.2
/
r
F
0.1
0
0
0.2
0.4
0.6
0.8
1
t/T
i
n
m
p
F
)
/
r
F
(
:
x
a
m
p
F
)
/
r
F
(
100
10
1
3
100
10
i
n
m
p
F
)
/
C
r
F
(
:
x
a
m
p
F
)
/
C
r
F
(
1
3
6
9
12
E [keV]
6
9
12
E [keV]
a=0.998 , r = 3
a=0.998 , r = 3
a=0.998 , r = 3
a=0.998 , r = 3
200
160
120
80
40
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
1000
100
10
1000
100
10
i
n
m
p
F
)
/
C
r
F
(
:
x
a
m
p
F
)
/
C
r
F
(
i
n
m
p
F
)
/
r
F
(
:
x
a
m
p
F
)
/
r
F
(
p
F
/
r
F
]
V
e
[
W
E
]
V
e
[
W
E
0
0
0.2
0.4
0.6
0.8
1
0
0
0.2
0.4
0.6
0.8
1
1
3
t/T
t/T
6
9
12
E [keV]
1
3
6
9
12
E [keV]
Figure 8. Left: The time variation of the observed equivalent width of the Kα line is shown for three different observer's inclination
angles -- 30◦ (dashed), 60◦ (dotted) and 85◦ (solid). The EW integrated over the whole orbit is shown in horizontal lines. Middle left:
The ratio of the observed reflected emission to the observed primary emission. Both energy fluxes are integrated in the energy range
3 -- 10 keV. The line styles are the same as in the left panel. Middle right: The ratio of the maximum of the ratio of the observed reflected
and primary emission Fr/Fp to its minimum. The solid line represents the energy dependence of this ratio, the dotted line shows this
ratio for fluxes integrated in the energy ranges 3 -- 6 keV, 6 -- 9 keV and 9 -- 12 keV. The inclination of the observer is 30◦ (bottom lines),
60◦ (middle lines) and 85◦ (top lines). Right: The same as in the middle right panel but without the flux originating in the Fe lines.
The top (bottom) panels correspond to the Schwarzschild (Kerr) black hole.
a=0 , r = 7
a=0 , r = 7
a=0 , r = 7
s
F
/
h
F
s
F
/
h
F
0.75
0.7
0.65
0.6
0
0.2
0.4
0.6
0.8
1
t/T
a=0.998 , r = 3
10
1
0.1
]
1
-
s
×
V
e
k
[
h
F
0.01
0.01
0.1
1
Fs [keV × s-1]
10
]
1
-
s
×
V
e
k
[
s
F
k
-
h
F
1
0.1
0.01
0.001
0.1
1
Fs [keV × s-1]
10
a=0.998 , r = 3
a=0.998 , r = 3
1
0.9
0.8
0.7
0.6
10
1
0.1
]
1
-
s
×
V
e
k
[
h
F
0.01
1
0.1
0.01
]
1
-
s
×
V
e
k
[
s
F
k
-
h
F
0.001
0
0.2
0.4
0.6
0.8
1
t/T
0.001
0.001
0.01
0.1
1
10
Fs [keV × s-1]
0.01
0.1
1
Fs [keV × s-1]
10
Figure 9. Left: The hardness ratio of the flux Fh in the energy range 6.5 -- 10 keV to the flux Fs in the energy range 3 -- 6.5 keV. The
flux in the Fe lines is also included. The dashed, dotted and solid lines correspond to the inclination of the observer being 30◦, 60◦ and
85◦. Middle: The hard flux Fh versus the soft flux Fs. The solid line has a slope of k = 0.59 (see text for more details). Right: The
same as in the middle panel but the line with the slope k = 0.59 is subtracted from the hard flux. The loops are for different observer's
inclinations (inner for 30◦, middle for 60◦ and outer for 85◦). The points are separated by the time 0.005 T . The top (bottom) panels
correspond to the Schwarzschild (Kerr) black hole.
8 M. Dovciak, V. Karas, G. Matt and R.W. Goosmann
emitted. Therefore one needs to know where the spot was at
the time when a particular observed photon was emitted, in
other words one must know the dependence ϕ(ϕe, t) given by
eq. (4). If the spot is locally moving with a constant velocity,
the observer at infinity sees that the spot changes its veloc-
ity, due to its motion and different time delays with which
the emitted photons are observed. Thus the spot seems to
either fall behind or overrun the position it should have if all
photons arrive with the same time delay (see the left panel
in Fig. 4 for ϕe = 0). This is true even in the classical case.
The differences between the classical and relativistic cases
can be very large especially for the closer orbits of the spot,
larger observer's inclination angle and faster rotating black
hole.
The transfer function consists of several components --
the energy shift g, the lensing l and for an extended source
also the emission angle µe. The contribution of each of these
components to the transfer function for our set of parameters
can be seen in Fig. 4 (taken for the position of the flare or
centre of the spot).
We show the magnitude of General Relativity effects
for the Schwarzschild (a = 0 GM/c3) black hole and the
extremally spinning Kerr1 (a = 0.998 GM/c3) black hole,
assuming different inclinations of the observer and setting
the primary flare at a small height h above the disc (typically
h = 0.015 GM/c2, resulting in the spot radius 0.86 GM/c2).
The power-law photon index of the primary radiation is Γ =
1.9 and the spot is illuminated by photons emitted within a
downwards directed cone (with the half-opening angle 89◦).
The energy shift, g-factor, can either amplify or dimin-
ish the observed flux. The latter is the case for the whole
orbit for the inclination θo = 30◦ which is due to the fact
that the spot orbits very close to the black hole and gravi-
tational redshift prevails over the Doppler shift.
The lensing effect has the largest amplification contri-
bution for the inclination 85◦ whereas we can neglect it for
the lowest inclination of 30◦. In the middle case (θo = 60◦)
its effect is comparable to other amplification parameters.
The emission angle is important for the extended source
not only as part of the transfer function but also as a variable
that the reflected component fr of the local flux depends on
heavily, as can be seen from the left panel in the Fig. 3 --
the local flux changes by one order of magnitude for the full
range of angles. The emission angle can change dramatically
for higher inclinations during the whole orbit, and for θo =
85◦ in the Kerr case it acquires almost all possible values.
The delay amplification plays an important role in mod-
ifying the local flux as can be seen from the left panels in
Fig. 5. As described earlier, according to the observer, the
spot spends more time in certain parts of the orbit than in
the others and, as a consequence, less photons per unit ob-
servers time is detected. The kt factor has a larger maximum
in the Kerr case. It is mainly due to the fact that the spot
orbits with a larger velocity being closer to the black hole.
From the middle and right panels in Fig. 5 one can see
1 We use the value a = 0.998 GM/c3 because it is usually ac-
cepted as astrophysically the most extremal case of a rotating
black hole. The exact value for the extremal spin of the astro-
physical black hole is rather model-dependent although close to
0.998 GM/c3 for a standard disc (see Thorne 1974).
what is the overall observed amplification of the local flux
from the centre of the spot (extended source) and from the
flare (point source). Because the overall amplification of the
flare's flux has one g-factor more, the flux from the flare
is more amplified where g > 1 and less in the other case.
For larger inclinations the overall amplification has a larger
maximum in the Kerr case, because of the larger lensing
effect, Doppler shift and delay amplification, whereas for
lower inclinations it is lower because of the much smaller
gravitational shift.
3.2 Spectral characteristics of the observed signal
The observed light curves computed for the 3 -- 10 keV en-
ergy range can be seen in the Fig. 6. The light curves are
influenced mainly by the overall amplification factor (trans-
fer function and delay amplification) and by the dependence
of the local flux on the emission angle. The primary emis-
sion dominates the observed flux as expected, meanwhile
the reflected flux in the Fe lines from the spot contributes
less. There is an exception in this behaviour, though, for
some parts of the orbit in the Kerr case, when the reflected
flux from the spot exceeds the flux of the primary (see cases
θo = 60◦, 85◦).
Fig. 7 shows the mean spectra taken over the whole or-
bit. The line is smeared when taken over the whole orbit. As
it is well known (Iwasawa et al. 1996) in the Schwarzschild
case the line stays above 3 keV, while in the Kerr case it
can be shifted even below this energy (as is the case for all
inclinations for the spot orbit at 3 GM/c2). The iron edge
is smeared in all studied cases and the dominance of the
primary emission is evident.
In order to quantify the properties of the observed spec-
tra let us look at the equivalent width, ratio of the observed
reflected and primary components, and the hardness ratio
(Figs. 8 and 9).
A closer look at the EW reveals that it does not much
differ from its local value except for the Kerr case with an
observer inclination of 85◦. From eq. (13) it follows that the
observed EW should be equal to the local one multiplied by
the g-factor for a small spot, but this is not true in our case
when we take also the primary emission into account. This
is due to the fact that the relationship Gkt ≈ Gsks
t does
generally not hold true. Only in the reflection the factors
Gkt for the continuum and the line components cancel each
other, hence if we compute the EW for the reflected emission
only, then it does behave as the local one times g.
For the spot close to the black hole (r = 3 GM/c2) the
EW is changing with respect to its mean value by 30% even
for a low inclination angle 30◦. For an almost edge-on disc
it can vary by as much as 200%.
The observed ratio of the reflected flux to the primary
flux is amplified when compared to the local one. Again, the
approximation of a small spot cannot be used. The amplifi-
cation is the highest in the Kerr case with θo = 85◦ -- the
ratio is increased by more than one order of magnitude. Note
that in the Kerr case, for the inclinations 60◦ and 85◦ the
ratio of the observed reflected flux to the observed primary
flux is larger than unity, meaning the reflected component
prevails over the primary one. Because the primary flux is
energy-dependent (a power-law), this ratio depends on en-
ergy, too. One can ask what is the maximum and the min-
imum value of this ratio over the whole orbit. The energy
dependence of this ratio is shown in the middle right (includ-
ing the line emission) and right (without the line) panels in
Fig. 8. If the reflected continuum were without features (Fe
lines and the edge), the ratio would just slightly decrease
with energy because of the power-law local primary flux (see
e.g. 5 -- 8 keV and above 10 keV regions for θo = 85◦, top
line, in the Schwarzschild case). The Fe edge causes more
complicated behaviour of this ratio when it influences the
minimum (see 8 -- 10 keV region for θo = 85◦, top line, in the
Schwarzschild case), or (together with the flux originating
in the iron line) the maximum (see e.g. 3.8 -- 6.3keV and 4.3 --
5.2keV region for θo = 30◦, bottom line, in Schwarzschild
case).
To evaluate the hardness ratio we compared the fluxes
in between 3 -- 6.5 keV (soft component, Fs) and 6.5 -- 10 keV
(hard component, Fh). The hardness ratio is amplified when
we compare it with the local hardness ratio (Fig. 3 and
the left panel in Fig. 9). The amplification is the largest
in Kerr case. The value of the k-coefficient in eqs. (18) and
(20) defined by the primary flux is for these energy ranges
k = 0.59. The local hardness ratio is larger than this value
only by a multiplicative factor 1.01 -- 1.05, whereas the ob-
served hardness ratio has a maximum larger by ≈ 1.2 in the
Schwarzschild case, and by more than 1.5 in the Kerr case.
The sudden increase in the observed hardness ratio around
the time t/T = 0.5 in Schwarzschild case and t/T = 0.4
in Kerr case is due to the fact that the Fe Kα line passes
from the soft to the hard component and back because of
the Doppler shift. The flux -- flux graph is shown in the mid-
dle panel of Fig. 9 where the line with the slope equal to
k = 0.59 is shown as well. Points for all inclinations are in-
cluded in this graph. The deviation from this line is shown
in the right panel of the same figure. The points in the flux
to flux graph (for different emission angles) for the local flux
lie close to the line with the slope ≈ 1.
4 SUMMARY AND CONCLUSIONS
We discussed the General Relativity effects in the observed
emission of the spot model. The primary flux was included
and the mutual normalizations were treated within the
framework of a simple yet self-consistent scheme. About half
of the isotropic primary flux hits the spot and is reprocessed
(this has been computed by a Monte-Carlo scheme), and
part of the reprocessed radiation is re-emitted towards the
observer. The radiation is influenced by the relativistic ef-
fects before reaching the observer (these have been treated
in terms of the transfer function of the KY code, see Dovciak
et. al. 2004b). As an example of the two sets of parameters
(a Schwarzschild black hole versus an extremally spinning
Kerr black hole) we find that all components of the transfer
function -- the energy shifts (Doppler and gravitational),
aberration effects (which are in the interplay with the limb
darkening/brightening laws) and the lensing -- are impor-
tant and need to be taken into account. As the spot is orbit-
ing rapidly in the inner regions of the accretion disc, timing
is also essential. Motion results in the delay amplification,
which is again in a complex interplay with the bending of
light near the black hole; this is described by eq. (5).
We would like to emphasize that whereas the signifi-
Spectral variations from an orbiting spot
9
cance of the energy shift and lensing effect has already been
widely studied, the effect of the finite light travel time has
not been discussed much, in spite of the fact that it signif-
icantly affects the resulting signal and causes an additional
enhancement of the observed flux, especially in the case of
a fast-rotating black hole. Although this light time effect
arises as an immediate consequence of the finite velocity of
light, its mutual interplay with the light bending and focus-
ing is rather complicated. We thus examined this time delay
amplification in some detail. Our results also confirm what
has been long suspected, namely, that the local reprocessing
in the disc medium and the signal propagation through the
curved spacetime are all mutually interconnected in a rather
complex manner.
We have demonstrated that various integral character-
istics of the spectral features (such as equivalent widths)
are at best moderately sensitive to the black hole rotation.
The equivalent width could be significantly amplified in our
model only if the primary emission were beamed towards the
disc, thus decreasing the observed primary emission. Both
the ratio of the observed reflected to the observed primary
flux and the hardness ratio are amplified when compared to
the values for the intrinsic (local) emission.
In addition to the shown cases, we also performed our
computations for some other values of the spin of the black
hole and orbital radius of the spot. We found out that the
results do not differ significantly for different spin when
the radius is kept the same. We examined the cases with
a = 0.998 GM/c3 and a = 0 GM/c3 with r = 7 GM/c2
and the cases with a = 0.998 GM/c3 and a = 0.93 GM/c3
with r = 3 GM/c2 (the spin in the latter case could not
be lower if the spot should be above the marginally stable
orbit). Moreover, the spectral characteristics depend on the
orbital radius mainly close to the black hole whereas father
away their variability either does not change significantly or
slowly decreases.
The main conclusions of this work are summarized as
follows.
(i) All general and special relativistic effects are impor-
tant, none of them can be neglected.
(ii) The EW, apart for the extreme cases of high inclina-
tions, does not differ significantly from the local EW, how-
ever it varies even for low inclination of 30◦ by up to 30%
when compared with its mean value for the whole orbit.
(iii) The variability of the reflected to the primary flux
ratio and hardness ratio changes rapidly with the radius if
the spot orbits close to the black hole.
(iv) The spin of the black hole affects significantly our
results only as far as it determines the location of the
marginally stable orbit.
It follows from the last two points that the flux ratios could
be used for estimating the lower limit of possible values of
the spin parameter if the flare arises in the close vicinity of
the black hole.
We must emphasize that nowadays X-ray satellite ob-
servatories usually do not have enough sensitivity to be suit-
able for real fitting with our model. However, future satellite
missions like XEUS or CONSTELLATION-X will enable to
spectroscopically follow the motion of individual spots close
to the innermost stable orbit in active galactic nuclei (Goos-
mann et al. 2007). The azimuthally-dependent effects on the
10 M. Dovciak, V. Karas, G. Matt and R.W. Goosmann
Misner C. W., Thorne K. S., Wheeler J. A. 1973, Gravita-
tion (Freeman, San Francisco)
Poutanen J., Fabian A. C. 1999, MNRAS, 306, L31
Reynolds C. S., Nowak M. A. 2003, Phys. Rep., 377, 389
Romanova M. M., Ustyugova G. V., Koldoba A. V.,
Chechetkin V. M., Lovelace R. V. E. 1998, ApJ, 500, 703
Turner T. J., Mushotzky R. F., Yaqoob T., George I. M.,
Snowden S. L., Netzer H., Kraemer S. B., Nandra K.,
Chelouche D., 2002, ApJ, 574, L123
Thorne K. S. 1974, ApJ, 191, 507
Yaqoob T., George I. M., Kallman T. R., Padmanabhan U.,
Weaver K. A., Turner T. J., 2003, Apj, 596, 85
iron line band that we investigate in this paper can thus
be used to constrain the emission structure and the metric
around supermassive black holes.
ACKNOWLEDGMENTS
This research is supported by the ESA PECS project
No. 98040. MD and VK gratefully acknowledge support
from the Czech Science Foundation grants 205/05/P525
and 205/07/0052. GM acknowledges financial support from
Agenzia Spaziale Italiana (ASI). RG is grateful
for fi-
nancial support to the Centre of Theoretical Astrophysics
(LC06014).
REFERENCES
Blaes O. M., Davis S. W., Hirose S., Krolik J. H.,
Stone J. M. 2006, ApJ, 645, 1402
Collin S., Coup´e S., Dumont A.-M., Petrucci P.-O.,
R´oza´nska A. 2003, A&A, 400, 437
Cunningham C. T. 1975, ApJ, 202, 788
Czerny B., R´oza´nska A., Dovciak M., Karas V., Dumont
A.-M. 2004, A&A, 420, 1
Czerny B., Collin S., Dovciak M., Dumont A.-M.,
Falewicz R., Goosmann R., Karas V., Ponti G., Pres P.,
Zycki P. T. 2005, Astro-
Siarkowski M., Sylwester J.,
physical Sources of High Energy Particles and Radiation,
AIP Conference Proceedings, Eds.: K. Girardi, M. Flikop,
American Institute of Physics, 801, 188
Dabrowski Y., Lasenby A. N., 2001, MNRAS, 321, 605
Dovciak M. 2004, PhD Thesis (Charles University, Prague),
arXiv:astro-ph/0411605
Dovciak M., Bianchi S., Guainazzi M., Karas V., Matt G.
2004a, MNRAS, 350,745
Dovciak M., Karas V., Yaqoob T. 2004b, ApJS, 153, 205
Dumont A.-M., Abrassart A., Collin S. 2000, A&A, 357,
823
Fabian A. C., Iwasawa K., Reynolds C. S., & Young A. J.
2000, PASP, 112, 1145
Galeev A. A., Rosner R., Vaiana G. S., 1979, ApJ, 229, 318
George I. M., Fabian A. C. 1991, MNRAS, 249, 352
Ghisellini G., George I. M., Fabian A. C., Done C. 1991,
MNRAS, 248, 14
Goosmann R. W. 2006, PhD thesis (Universitat Hamburg,
Germany)
Goosmann R. W., Mouchet M., Czerny B., Dovciak M.,
Karas V., R´oza´nska A., Dumont A.-M. 2007, A&A, 475,
155
Guainazzi M., 2003, A&A, 401, 903
Iwasawa K., Fabian A. C., Reynolds C. S., Nandra
K., Otani C., Inoue H., Hayashida K., Brandt W. N.,
Dotani T., Kunieda H., Matsuoka M., Tanaka Y. 1996,
MNRAS, 282, 1038
Martocchia A., Karas V., Matt G. 2000, MNRAS, 312, 817
Martocchia A., Matt G. 1996, MNRAS, 282, L53
Matt G., Fabian A. C., Ross R. R. 1993, MNRAS, 262, 179
Merloni A., Fabian A. C. 2001, MNRAS, 321, 549
Miniutti G., Fabian A. C. 2004, MNRAS, 349, 1435
Miniutti G., Fabian A. C., Goyder R., Lasenby A. N. 2003,
MNRAS, 344, L22
|
astro-ph/0204502 | 1 | 0204 | 2002-04-29T21:34:50 | Red Giant Branch Stars as Probes of Stellar Populations I: 2MASS Calibration and Application to 2MASS GC01 | [
"astro-ph"
] | The near-infrared behavior of the red giant branch (RGB hereafter) as a function of abundance is examined with an unprecedented large sample of 27 Galactic globular clusters with Two Micron All Sky Survey photometry. We propose a new simplified analyses, involving the zero point of the RGB slope fit, and derive calibrations for the RGB slope, zero point, and tip. The weak metallicity sensitivity of the zero point leads to a ``fan''-like diagram for obtaining the abundance distributions in resolved stellar systems, and reddening estimates. Finally, we apply the new calibrations to the recently discovered Galactic globular cluster 2MASS GC01, to derive [Fe/H]_H96=-1.19+/-0.38 mag. The uncertainty is dominated by the severe foreground contamination. We estimate an extinction of A_V=21.07+/-2.20 mag toward the cluster. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(DOI: will be inserted by hand later)
October 27, 2018
2
0
0
2
r
p
A
9
2
1
v
2
0
5
4
0
2
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Red Giant Branch Stars as Probes of Stellar Populations I:
2MASS Calibration and Application to 2MASS GC01
Valentin D. Ivanov1 and Jordanka Borissova2
1 European Southern Observatory, Ave. Alonso de Cordova 3107, Casilla 19, Santiago 19001, Chile
e-mail: [email protected]
2 Institute of Astronomy, Bulgarian Academy of Sciences, and Isaac Newton Institute of Chile, Bulgarian Branch,
72 Tsarigradsko Chauss`ee, BG -- 1784 Sofia, Bulgaria
e-mail: [email protected]
Received ...; accepted ...
Abstract. The near-infrared behavior of the red giant branch (RGB hereafter) as a function of abundance is
examined with an unprecedented large sample of 27 Galactic globular clusters with Two Micron All Sky Survey
photometry. We propose a new simplified analyses, involving the zero point of the RGB slope fit, and derive
calibrations for the RGB slope, zero point, and tip. The weak metallicity sensitivity of the zero point leads to a
"fan"-like diagram for obtaining the abundance distributions in resolved stellar systems, and reddening estimates.
Finally, we apply the new calibrations to the recently discovered Galactic globular cluster 2MASS GC01, to
derive [Fe/H]H96 = −1.19 ± 0.38 mag. The uncertainty is dominated by the severe foreground contamination. We
estimate an extinction of AV = 21.07 ± 2.20 mag toward the cluster.
Key words. globular clusters: general -- Galaxy: abundances -- Galaxies: abundances -- Galaxies: distances and
redshifts Stars: distances -- Stars: abundances
1. Introduction
The red giant branch (RGB hereafter) stars are among
the brightest red stars in stellar systems, older than a few
Gyrs. These stars appear in almost all galaxies, including
II Zw 40 ( Ostlin 2000), considered until recently as the
best candidate for a primeval galaxy. Therefore the red
giants are a promising tool for probing the parameters of
old populations and the history of star formation in any
galaxy.
Galactic globular clusters, with their single age and
metallicity, are the ideal sites for calibrating the RGB pa-
rameters. Since Da Costa & Armandroff (1990) provided
the first reliable calibration of the position of the RGB ver-
sus metallicity, there has been a significant advancement,
both because of the improvement of the astronomical in-
strumentation and the development of the corresponding
theory.
The infrared waveband is particularly compelling for
such studies, in comparison with the optical one, because
of the potential to probe the stellar populations of sys-
tems with high foreground and/or intrinsic extinction.
The relatively small size and field of view of the IR ar-
rays have made it more difficult to carry out photometry
of large areas, and to compile uniform samples, necessary
to calibrate the RGB parameters reliably, in comparison
with the optical region. The Two Micron All Sky Survey1
(2MASS hereafter) offered for the first time such an op-
portunity.
Previous calibrations suffered a number of drawbacks.
Kuchinski, Frogel, & Terndrup (1995) and Kuchinski &
Frogel (1995) studied only metal rich globular clusters,
with [Fe/H]≥ −1. Later on, Ivanov et al. (2000a) added
to their sample three metal poor globular clusters. Ferraro
et al. (2000) used exceptional quality data but in a pho-
tometric system, based on unpublished standards by Ian
Glass (South African Astronomical Observatory) with no
available transformations to any of the other systems. The
similarity of their final results to those of Ivanov et al., who
used the CIT system, leads to a conclusion that the two
photometric systems are not radically different.
To expand the basis of the RGB studies, we used high
quality uniform photometry of Galactic globular clusters
from the 2MASS point source catalog, and assembled a
sample of RGBs of clusters with well known distances and
reddening. We calibrated the behavior of the RGB slope,
zero point, and tip with metallicity in the 2MASS photo-
1 See http://www.ipac.caltech.edu/2mass/releases/second/
Send offprint requests to: V. D. Ivanov
doc/explsup.html
2
Ivanov & Borissova: 2MASS Calibration of the RGB Parameters
metric system. For the first time we offer such calibration
in a well defined photometric system with all-sky cover-
age. We present a "fan" diagram, suitable for abundance
distribution analysis. Our results complement the recent
work of Cho & Lee (2001) and Grocholski & Sarajedini
(2001). who explored the properties of the RGB bump.
2. 2MASS Calibration of the RGB Parameters
2.1. Sample
Infrared photometry of about 80 of the 147 globular clus-
ter, listed in Harris (1996, revision June 22, 1999) is cur-
rently available from the First and the Second 2MASS
incremental releases. However, many of the observed clus-
ters are rendered unsuitable for this project because they
suffer from one or more of the following drawbacks:
1. Severe foreground contamination. This refers mostly
to globular clusters in the general direction of the
Galactic center.
2. The photometry of some distant halo clusters or highly
reddened clusters is not sufficiently deep.
3. Compact globular clusters have unresolved cores, lim-
iting the number statistics, and leading to crowding
problems.
After a visual inspection of the luminosity functions
and color-magnitude diagrams (CMD hereafter) of all
available globular clusters, we selected a subset of 27 ob-
jects (Table 1) for which reliable estimates of at least one
RGB parameters could be obtained. To calculate the ab-
solute magnitudes, and the intrinsic colors of stars, we
adopted the reddening and distance estimates from the
compilation of Harris (1996), and the reddening law of
Rieke & Lebofsky (1985; used throughout this paper).
A special care was taken to assure the uniformity of
the metallicity data. Harris (1996) based his compila-
tion on the system established by Zinn & Wsst (1984),
Zinn (1985), and Armandroff & Zinn (1988). Carretta &
Gratton (1997) suggested that this scale may overestimate
the abundance of metal-rich globular clusters, and devel-
oped a new one. We list their measurements in Table 1 as
well. Values in brackets are estimates based on the trans-
formation they derive, to the Zinn & West scale (see Eq.
7 of Carretta & Gratton).
Ferraro et al. (1999) argued that the total amount
of heavy elements is a better metallicity measurement,
because it accounts naturally for the opacity variations,
which depend on the total metallicity, not just the iron
abundance (Salaris, Chieffi, & Straniero 1993). They de-
veloped a news scale, designed to measure the that. For
the clusters, absent in their sample, we adopted their pre-
scription for α-element enhancement (Sect. 3.4 in Ferraro
et al.; estimates bracketed in Table 1).
Since the metallicities originate from different sources,
and the corresponding uncertainties are often not quoted,
we adopted an uniform error of 0.20 dex for all measure-
ments. This is a typical value of the accuracy of the abun-
dance estimates.
Table 1. Globular cluster sample. The Columns give:
cluster ID, color excess EB−V , distance modulus, tidal ra-
dius Rt (all from the compilation of Harris 1996), and
abundances in various scales (see the references).
(m-M)v
[Fe/H]
NGC EB−V
mag
0.04
0.03
0.02
0.01
0.14
0.02
0.12
0.00
0.36
0.32
0.33
0.02
0.37
0.28
0.28
0.44
0.28
0.16
0.34
0.14
0.20
0.07
0.25
0.16
0.10
0.06
0.03
mag
ID
13.37
104
14.69
288
15.47
1851
15.59
1904
15.59
2298
16.38
5024
13.97
5139
16.15
5466
12.83
6121
16.06
6144
15.06
6171
14.48
6205
15.85
6273
16.70
6284
16.77
6356
16.62
6441
15.37
6624
15.16
6637
13.60
6656
17.61
6715
15.65
6779
13.87
6809
13.75
6838
16.87
6864
15.37
7078
15.49
7089
7099
14.62
References for [Fe/H]:
H96 - Harris (1996; Zinn scale);
CG97 - Carretta & Gratton (1997);
F99 - Ferraro et al. (1999)
H96
−0.76
−1.24
−1.22
−1.57
−1.85
−1.99
−1.62
−2.22
−1.20
−1.73
−1.04
−1.54
−1.68
−1.32
−0.50
−0.53
−0.42
−0.71
−1.64
−1.59
−1.94
−1.81
−0.73
−1.32
−2.25
−1.62
−2.12
′
F99
[M/H] Rt
[Fe/H]
CG97
−0.70 −0.59 47.25
−1.07 −0.85 12.94
−1.08 −0.88 11.70
−1.37 −1.22
8.34
−1.74 (−1.54) 6.48
(−1.82) (−1.62) 21.75
(−1.38) (−1.18) 44.85
(−2.14) (−1.94) 34.24
−1.19 −0.94 32.49
−1.49 (−1.29) 33.25
−0.87 −0.70 17.44
−1.39 −1.18 25.18
(−1.45) (−1.25) 14.50
(−1.10) (−0.90) 23.08
(−0.66) (−0.50) 8.97
(−0.67) (−0.51) 8.00
(−0.64) (−0.48) 20.55
−0.68 −0.55
8.35
(−1.41) (−1.21) 29.97
(−1.35) (−1.15) 7.47
(−1.75) (−1.55) 8.56
−1.61 −1.41 16.28
−0.70 −0.49
9.96
(−1.10) (−0.90) 7.28
−2.12 −1.91 21.50
(−1.38) (−1.18) 21.45
−1.91 −1.71 18.34
The first step before estimating the RGB parameters
was to eliminate statistically the foreground contamina-
tion. As many stars were removed randomly from the
CMD of the globular clusters, as the number of stars,
present in a nearby field with the same area as the glob-
ular cluster field. The statistical removal is not reliable in
case of severely contaminated globular clusters, where the
number of foreground stars was comparable to the number
of cluster members. A subset of CMDs for four clusters is
shown in Figure 1 and 2. In addition, we removed the stars
within 15 arcsec from the clusters centers, to minimize the
crowding effects.
2.2. RGB Slope and Zero Point
Da Costa & Armandroff (1990) prepared a set of stan-
dard RGBs on the V versus V-I CMD, for metallicity es-
timates of globular clusters in the optical. The technique
has been expanded toward the near infrared wavebands,
where the giant branch is linear, and therefore, more ob-
Ivanov & Borissova: 2MASS Calibration of the RGB Parameters
3
licity indicator. It is significantly less demanding in terms
of observing time and telescope collecting area than the
spectroscopic methods, but unlike them it can be applied
reliably only to uniform groups of stars.
We followed the procedure to determine the RGB
slope, described in Ivanov et al. (2000a), fitting the RGB
on the [Ks, (J − Ks)] CMD with a linear equation: J −
Ks = a × Ks + b. Only the stars above the horizontal
branch were included. We applied a least square method,
taking into account the uncertainties along both axis. Two
iterations were carried, and all the stars outside 5σ from
the first fit were removed from the calculations.
For the purpose of fitting the RGB of an individual
cluster, the zero point is equivalent to previous calibra-
tions of the RGB colors at fixed K-band magnitude levels
(e.g. Ivanov et al. 2000a; Ferraro et al. 2000), although
more straightforward. To derive the zero-point relation to
the abundance, we used only the clusters which suffer min-
imal extinction (EB−V ≤ 0.3 mag).
Notably, the fit to the zero point shows a very small
variation with metallicity: 0.11 mag for [Fe/H]H96 varying
from -2 to -0.5. For comparison, (J − KS)0 at KS = −5.5
mag varies by 0.32 mag for the same metallicity range
(Ferraro et al. 2000). This result offers the possibility for
a more reliable reddening estimates than in case of RGB
color calibrations.
The behavior with metallicity of the RGB slope and
zero point are demonstrated in Figure 3. Fitting coeffi-
cients for individual cluster RGBs are listed in Table 2.
2.3. RGB Tip
The tip of the RGB is a well known distance indicator.
Mould & Kristian (1986), Lee (1993) and Lee et al. 1993
pioneered the method in the optical, measuring the cor-
responding jump in the luminosity function in external
galaxies. Later it was applied to number of resolved stellar
systems, among the most challenging of which were some
Virgo cluster members (Harris et al. 1998) and NGC 3379
(Sakai et al. 1997).
This is a statistically demanding technique, requiring
50-100 stars per bin. The limited size of infrared arrays
explains the difficulty to apply it in the near infrared.
Although it does cover a large area, the 2MASS photom-
etry can not alleviate the intrinsic problem of the small
number of giants in globular clusters. Thus, we assume
that the brightest cluster member represents the RGB tip.
In most of the cases the tip is obvious, but sometimes
additional criteria had to be applied to determine the
brightest stars. We took advantage of the linearity of the
RGB in the near infrared, and excluded from the consid-
erations bright stars that deviated from the color of the
RGB at a given magnitude level, predicted by the RGB
slope fit by more than 0.5 mag. We also excluded some ex-
tremely bright stars, with luminosity higher than the rest
of the RGB by 2-3 mag. They were obvious foreground
Fig. 1. Color-magnitude diagrams for a subset of metal
poor globular clusters. The left panels show the raw data.
The right panels show the reddening corrected CMDs,
converted to absolute magnitudes. The foreground con-
tamination and stars with (J − KS)0 < 0.4 mag have
been removed.
Fig. 2. Same as Figure 1 for metal rich clusters.
vious and easier to define compared with the optical or
optical-infrared CMDs. The slope is related to the effec-
tive temperature of the stars along the RGB, and Teff in
turn depends on the opacity, and the heavy element abun-
dance. The slope is a reddening and distance-free metal-
4
Ivanov & Borissova: 2MASS Calibration of the RGB Parameters
Table 2. Linear fits to the RGB in the form of: J − Ks =
a×Ks +b. The radii Rmax of the region around the cluster
center used in the fit, and the root mean square of the fit
are given.
′
a
Slope
b
NGC Rmax
Zero Point
10 −0.125(0.002) 0.368(0.004)
−0.105(0.006) 0.341(0.018)
7
−0.098(0.004) 0.399(0.013)
5
−0.081(0.004) 0.356(0.017)
7
6
−0.055(0.005) 0.389(0.019)
11 −0.061(0.004) 0.417(0.014)
7
−0.085(0.001) 0.344(0.005)
10 −0.047(0.008) 0.468(0.030)
10 −0.094(0.003) 0.360(0.010)
4
−0.066(0.003) 0.431(0.014)
8
−0.101(0.003) 0.310(0.009)
10 −0.086(0.003) 0.354(0.010)
−0.074(0.003)
3
−0.110(0.005) 0.311(0.023)
2
−0.059(0.003)
3
−0.083(0.004) 0.343(0.021)
3
3
−0.048(0.005) 0.445(0.020)
10 −0.077(0.004) 0.396(0.013)
3
−0.099(0.003) 0.352(0.009)
−0.078(0.004) 0.411(0.020)
3
−0.080(0.003) 0.369(0.012)
9
8
−0.050(0.006) 0.450(0.024)
r.m.s.
mag
ID
0.054
104
0.048
288
0.095
1851
0.088
1904
0.104
2298
0.107
5024
0.063
5139
0.082
5466
0.091
6121
0.088
6144
0.096
6171
0.068
6205
0.099
6273
0.109
6356
0.074
6656
0.143
6715
0.069
6779
0.064
6809
0.085
6838
0.110
6864
0.090
7089
7099
0.047
Note: Fitting errors are given in brackets. The Zero
Points for NGC 6273 and 6656 are omitted because
of the large reddening toward this cluster.
Fig. 3. Relation of the [Fe/H] in various metallicity scales
versus RGB slope on [MKs, (J − Ks)0] diagram. The bars
indicate 1σ uncertainties. The dashed line is the fit of
Ferraro et al. (2000) drawn without transformation of col-
ors. The zero point errorbars show only the statistical er-
rors from the fits to the RGB, omitting the 0.2 mag added
in quadrature to account for the uncertain distance moduli
and reddening corrections.
contamination. Finally, the red variables from Clement et
al. (2001) were excluded .
The formal uncertainties of the stellar magnitudes
given in 2MASS were discarded, since they do not repre-
sent well the uncertainty in the tip magnitude. Instead, we
adopted the difference in the magnitudes of the two bright-
est stars, accounting for the possibility that the brightest
star may be a non-member. This led to typical error values
of 0.2-0.4 mag, much larger than the 2MASS errors.
The RGB tip magnitudes for 20 globular clusters are
given in Table 3. The behavior of the RGB tip with metal
abundance is shown in Figure 4.
2.4. Results
The behavior of the derived RGB parameters with metal-
licity was fitted with linear equations, taking into account
the errors along both axes. The coefficients are given in
Table 4. Figure 5 shows a "fan"-like grid of RGBs in a red-
dening and distance corrected CMD, for the three metal-
licity scales discussed in Section 2.1. It demonstrates that
the RGB behavior can be reduced to a simple rotation
around a nearly-fixed point. This comes to no surprise,
since the zero point of the RGB fit is almost independent
of the abundance (Section 2.2).
Fig. 4. Relation of the [Fe/H] in various metallicity scales
versus RGB tip absolute magnitude. The bars indicate 1σ
uncertainties, with the Y-axis errors including the uncer-
tainties due to the reddening and distance. The dashed
line is the fit of Ferraro et al. (2000) drawn without trans-
formation of colors.
Ivanov & Borissova: 2MASS Calibration of the RGB Parameters
5
Table 4. Polynomial fits to the metallicity behavior of the
RGB slope, zero point, bump, and tip: Y=a0+a1×X.
Variables
Y
X
Coefficients
r.m.s. Npts
a0(σa0 )
a1(σa1 )
H66 RGBSl −0.157(0.009) −0.051(0.006) 0.002
CG97 RGBSl −0.158(0.010) −0.058(0.007) 0.002
F99 RGBSl −0.149(0.009) −0.060(0.008) 0.002
H66 RGBZP +0.277(0.152) −0.070(0.104) 0.007
CG97 RGBZP +0.272(0.155) −0.082(0.120) 0.007
F99 RGBZP +0.285(0.137) −0.084(0.122) 0.007
−5.650(0.187) −0.323(0.121) 0.025
H66
Jtip
CG97 Jtip
−5.690(0.210) −0.387(0.151) 0.026
−5.631(0.191) −0.399(0.158) 0.026
F99
Jtip
−6.641(0.186) −0.486(0.121) 0.025
H66 Htip
−6.712(0.210) −0.594(0.153) 0.024
CG97 Htip
F99 Htip
−6.631(0.193) −0.620(0.162) 0.024
−7.032(0.212) −0.615(0.134) 0.032
H66 Ktip
−7.109(0.243) −0.739(0.171) 0.034
CG97 Ktip
F99 Ktip
−7.003(0.224) −0.768(0.181) 0.034
Note: Errors are given in brackets. Metallicity
notation is the same as in Table 1.
22
22
22
20
20
20
20
20
20
20
20
20
20
20
20
Table 3. Estimated RGB tip absolute magnitudes.
NGC
J
H
Ks
104 −5.616(0.171) −6.372(0.017) −6.862(0.209)
288 −5.578(0.621) −6.412(0.610) −6.672(0.656)
1851 −5.182(0.083) −6.092(0.024) −6.306(0.040)
1904 −5.055(0.189) −5.826(0.231) −5.940(0.211)
5024 −5.093(0.049) −5.708(0.006) −5.919(0.035)
5139 −5.124(0.106) −5.778(0.040) −6.015(0.003)
6144 −5.062(0.294) −5.803(0.318) −5.953(0.280)
6171 −5.326(0.605) −6.213(0.565) −6.471(0.711)
6205 −5.156(0.050) −5.901(0.092) −6.027(0.020)
6284 −5.267(0.277) −6.073(0.133) −6.283(0.156)
6441 −5.521(0.246) −6.413(0.235) −6.788(0.288)
6624 −5.459(0.019) −6.295(0.027) −6.623(0.072)
6637 −5.496(0.474) −6.369(0.401) −6.667(0.426)
6656 −5.146(0.023) −5.809(0.067) −5.984(0.059)
6779 −5.067(0.050) −5.765(0.036) −5.950(0.054)
6809 −5.020(0.385) −5.725(0.335) −5.870(0.367)
6864 −5.154(0.072) −6.012(0.047) −6.133(0.007)
7078 −4.890(0.136) −5.506(0.211) −5.626(0.087)
7089 −4.966(0.046) −5.711(0.072) −5.853(0.076)
7099 −5.105(0.183) −5.655(0.134) −5.832(0.142)
Note: The difference between the two brightest
stars is given in brackets.
The grid allows to obtain the metallicities of individual
sdtars in resolved systems in the infrared, and to obtain
the metallicity distributions as it has been done before in
the optical (e.g. Saviane et al. 2000). The linear represen-
tation of the RGB in absolute KS magnitude versus the
intrinsic (J − KS)0 color is:
MKS = RGBZP + RGBSl × (J − KS)0
(1)
Both the slope RGBSl and the zero point RGBZP were
calibrated as functions of the abundance:
RGBSl = aSl
0 + aSl
1 × [Fe/H]
RGBZP = aZP
0 + aZP
1 × [Fe/H]
(2)
(3)
After substituting Eq. 2 and 3 in 1, and solving for the
stellar metallicity:
] / [aSl
1 ×(J −KS)0+aZP
1
](4)
Fig. 5. "Fan Diagram" - a grid of RGBs on the reddening
and distance corrected color-magnitude diagram, for the
three metallicity scales, as indicated.
[Fe/H] = [MKS −aSl
0 ×(J −KS)0−aZP
0
This equation should be used with caution. Small pho-
tometric errors are crucial for accurate [Fe/H] estimates.
The upper part of the infrared CMD is better suited for
determining individual metallicities because near the root
of the RGB the abundance relation degenerates. The ex-
act faint limit is determined by the quality of the pho-
tometry, and by the presence of the horizontal branch at
MKS ∼ −1.5 mag.
3. 2MASS GC01
Hurt et al. (1999, 2000) reported the serendipitous discov-
ery of 2MASS GC01 (hereafter GC01) in 2MASS data.
It is a heavily obscured globular cluster, laying in the
Milky Way disk, in the general direction of the Galactic
center. Ivanov, Borissova, & Vanzi (2000b) determined
(m − M)0 = 12.4 − 14.0, and AV = 20.9 − 18.8, assuming
[Fe/H] = −0.5 and −2.0, respectively. The main source of
uncertainty in these estimates was the unknown metallic-
ity of the cluster, although the location of GC01 suggested
that it might be a metal rich object.
6
Ivanov & Borissova: 2MASS Calibration of the RGB Parameters
The first step toward a metallicity estimate of GC01
was to remove the foreground stars contamination. Unlike
the clusters we used to derive the RGB parameters calibra-
tions, the contamination here is severe, reaching ∼ 30% in
the RGB region. We performed 2000 foreground substrac-
tions, and estimated the RGB slope for each realization
separately. This method yields distributions of the RGB
slope and tip, and the widths of these distributions mea-
sure the respective uncertainties.
To carry out this procedure we defined the CMD area,
encompassing the RGB: 8.0 ≤ KS ≤ 14.0 mag, and
2.5 ≤ J − KS ≤ 6.0 mag. Then we divided it into 0.2
mag square bins. Experiments with different bin sizes in-
dicated that any value between 0.2 and 1.0 mag leads to
the same conclusions.
Next, we constructed CMDs for the cluster field, and
for a surrounding field with an equal area. To minimize the
crowding effects we omitted the stars within 15 arcsec from
the cluster center. The outer limit of the cluster field was
constrained by the cluster diameter (3.3±0.2 arcmin, Hurt
et al. 1999, 2000). We carried out our calculations to two
values of the outer radii: 1.0 and 1.5 arcmin. Smaller values
limit the number statistics, and larger ones increase the
fraction of the foreground contamination. The foreground
field was defined as a circular annulus with 10.49 − 10.54
or 10.43 − 10.54 arcmin size, respectively for 1.0 and 1.5
armin cluster fields.
Finally we counted the stars in each bin, and sub-
tracted randomly from the "cluster" bins as many stars
as were present in the "field" bins. If the latter bin had
more stars than the former one, we subtracted stars from
the nearby bins, again in a random way.
We carried out a linear fit on the RGB stars in the
foreground subtracted CMD in the same manner as for
the calibration clusters (see Section 2.2). For stars without
error measurements in 2MASS we adopted σ = 0.20 mag.
Using the linearity of the RGB, we imposed a faint limit
of the stars, included in the fit, just above the horizontal
branch level. The luminosity function of GC01 (Ivanov et
al. 2000b, Figure 5) indicates that the horizontal branch
is at KS=13.0-13.2 mag. To minimize the uncertainties of
the RGB slope, we also imposed color limits on the stars
we used in the fit. The red one was set to J −KS=5.5 mag,
and has no effect on the slopes because of the negligible
number of stars to the red of the RGB. The results are
somewhat more sensitive to the blue limit. We choose to
impose J − KS=3.50 and 3.80 mag, because of the well-
defined limit of the RGB at this colors (Figure 6).
A summary of the results for the RGB slope and zero
point GC01 is presented in Figure 7. The RGB tip is
omitted because it is sensitive only to the adopted cluster
radius. Clearly, the effects from the assumed parameters
are smaller or comparable with the uncertainties origi-
nating due to the foreground contamination. We calcu-
lated the error-weighted averages for the realizations with
K lim
S = 12.5 − 13.0, to avoid the possible influence of the
cluster horizontal brunch. The determined RGB parame-
ters for GC01 are:
Fig. 6. 2MASS color-magnitude diagram of GC01. Solid
dots are stars from 0.25 to 1 arcmin from the cluster center
- cluster+foreground. X's indicate the stars within 10.49
to 10.54 arcmin from the center - pure foreground. 1σ
errors are indicated. The solid line represents the average
RGB fit, described in Section 3. Clearly, it is dominated
by the stars with minimal uncertainties. Dotted lines are
the adopted limits.
1. < Slope > = −0.0965 ± 0.0171
2. < ZeroPoint > = 5.40 ± 0.18
3. < Tip > = 8.87 ± 0.13
Using the equations, given in Table 4, we obtain:
[Fe/H]H96= −1.19±0.38, [Fe/H]CG97= −1.06±0.34, and
[M/H]F 99= −0.88±0.32. The errors given here include
both the uncertainties in the slope, and in the calibra-
tions. GC01 appears to resemble closely NGC 104.
The estimated absolute magnitude for the RGB tip, for
a cluster with such abundance is MKS = −6.30±0.35 mag,
using our new calibration for the Zinn metallicity scale.
This leads to a distance modulus of (m − M )0 + AKS =
15.17 ± 0.38 mag consistent with Ivanov et al. (2000b). We
refrain from further considerations based on the RGB tip
because of the poor statistics at the brighter end of the
RGB. Instead, we will adopt (m − M )0 + AKS = 15.0 ±
0.4 mag, a result of interpolation between the values for
[Fe/H]=−1.0 and −2.0, in Table 2 of Ivanov et al. (2000b).
We can also verify if the cluster reddening is consistent
with the previous estimates. First, we subtract the RGB
slope equations, written for the GC01 in apparent, and in
absolute magnitudes. Respectively:
(mJ − mKS ) = RGBapp
ZP + RGBSl × mKS
(MJ − MKS )0 = RGBabs
ZP + RGBSl × MKS
(5)
(6)
Ivanov & Borissova: 2MASS Calibration of the RGB Parameters
7
diagram, suitable for analyses of the metallicity spread in
resolved stellar systems.
The derived calibrations were applied to estimate
the metal abundance of the recently discovered globu-
lar cluster GC01. It is a particularly challenging object
because of the severe foreground contamination. We re-
moved it with a random procedure, and used the RGB
slope of the remaining pure cluster population to derive
[Fe/H]H96=−1.19±0.38. The uncertainty is dominated by
the foreground contamination, and albeit large, allows to
exclude extremely high abundances, expected from the
cluster location. GC01 is likely to be moderately metal
poor. The RGB tip and zero point yield distance modulus
and extinction, consistent with our previous estimates.
Acknowledgements. This publication makes use of data prod-
ucts from the Two Micron All Sky Survey, which is a joint
project of the University of Massachusetts and the Infrared
Processing and Analysis Center,
funded by the National
Aeronautics and Space Administration and the National
Science Foundation. The authors thank Dr. Ivo Saviane for
the useful discussions, and the referee Dr. M. G. Lee for the
comments that halped to improve the paper.
References
Armandroff, T.E. & Zinn, R., 1988, AJ, 96, 92
Carretta, E., & Gratton, R.G., 1997, A&AS, 121, 95
Clement, C.M., Muzzin, A., Dufton, Q., Ponnampalam, T.,
Wang, J., Burford, J., Richardson, A., Rosebery, T., Rowe,
J., Sayer Hog, H., 2001, AJ, 122, 2587
Cho, D.-H., & Lee, S.-G., astro-ph/0111204
Da Costa, G.S. & Armandroff, T.E., 1990. AJ, 100, 162
Ferraro, F.R., Messineo, M., Fusi Pecci, F., De Palo, M.A.,
Straniero, O., Chieffi, A., & Limongi, M., 1999, AJ, 118,
1738
Ferraro, F.R., Montegrifo, P., Origlia, L., & Fusi Pecci, F.,
2000, AJ, 119, 1282
Grocholski, A.J., & Sarajedini, A., 2001, astro-ph/0112251
Harris, W.E., 1996, AJ, 112, 1487
Harris, W.E., Durrell, P.R., Pierce, M.J., & Secker, J., 1998,
Nature, 359, 45
Fig. 7. Behavior of the RGB slope (top) and zero point
(bottom) for different parameters of the foreground sub-
traction and fits. The horizontal axis is the lower limit of
the stars used in the RGB fit. Triangles indicate cluster
radius of 1.0 arcmin, and circles indicate 1.5 arcmin. Open
symbols indicate a blue limit of J − KS=3.25 mag, and
solid symbols of J −KS=3.00 mag. Only some 1σ errors of
individual points are shown to avoid crowding. The solid
lines indicate the weighted averages, and the dashed lines
indicate their 1σ errors. The values to the right of the ver-
tical dotted line were discarded because they were affected
by the horizontal branch.
Here the notation is the same as in Table 4, and RGBabs
ZP =
0.36 ± 0.20 mag, for the [Fe/H]H96 given above. The result
of the subtraction is:
E(J −KS) = RGBapp
ZP −RGBabs
ZP +RGBSl×[(m−M )0+AKS ](7)
Substituting, we obtain: E(J − KS) = 3.58 ± 0.37 mag,
close to the values of 3.49 mag, interpolated as above from
Ivanov et al. (2000b). Our new estimate corresponds to
AV = 21.07 ± 2.20 mag, and AKS = 2.36 ± 0.25 mag.
4. Summary
The behavior of the RGB in the infrared was quantified
based on an unprecedented large sample of 2MASS pho-
tometry of Milky Way globular clusters. The RGBs were
fitted by straight lines. We produced new calibrations of
the RGB slope, tip, and - for the first time - zero point, as
functions of abundance. The introduction of the zero point
streamlines greatly the RGB analyses in comparison with
the traditional approach where RGB colors at given lev-
els were used. Notably, the zero point is fairly insensitive
to the abundance, varying by only 0.11 mag over a range
from [Fe/H]H96 = −2 to −0.5. We present a "fan"-like
Hurt, R.L., Jarrett, T., Cutri, R., Skrutskie, M., Schneider, S.,
& van Driel, W., 1999, AAS, 194.0711
Hurt, R.L., Jarrett, T.H., Kirkpatrick, J.D., Cutri, R.M.,
Schneider, S.E., Skrutskie, M., & van Driel, W., 2000, AJ,
120, 1876
Ivanov, V.D., Borissova, J., Alonso-Herrero, A., & Russeva, T.,
2000a, AJ, 119, 2274
Ivanov, V.D., Borissova, J., & Vanzi, L., 2000b, A&A, 362, L1
Kuchinski, L., Frogel, J.A., & Terndrup, D., 1995, AJ, 109,
1131
Kuchinski, L. & Frogel, J.A., 1995, AJ, 110, 2844
Lee, M.G., 1993, ApJ, 408, 409
Lee, M.G., Freedman, W.L., & Madore, B.F., 1993, ApJ, 417,
553
Mould, G. & Kristian, J., 1986, ApJ, 305, 591
Ostlin, G., 2000, ApJ, 535, L99
Rieke, G.H. & Lebofsky, M.J., 1985, ApJ, 288, 618
Sakai, S., Madore, B.F., Freedman, W.L., Lauer, T.R., Ajhar,
E.A., & Baum, W.A., 1997, ApJ, 478, 49
Salaris, M., Chieffi, A., & Straniero, O., 1993, ApJ, 414, 580
8
Ivanov & Borissova: 2MASS Calibration of the RGB Parameters
Saviane, I., Rosenberg, A., Piotto, G, & Aparicio, Al., 2000,
A&A, 355, 966
Zinn, R. & West, M.J., 1984, ApJS, 55, 45
Zinn, R., 1985, ApJ, 293, 424
|
0809.4857 | 1 | 0809 | 2008-09-28T18:15:48 | The Discovery of Stellar Oscillations in the K Giant $\iota$ Dra | [
"astro-ph"
] | $\iota$ Dra (HIP 75458) is a well-known example for a K giant hosting a substellar companion since its discovery by Frink et al. (2002). We present radial velocity measurements of this star from observations taken with three different instruments spanning nearly 8 years. They show more clearly that the RV period is long-lived and coherent thus supporting the companion hypothesis. The longer time baseline now allows for a more accurate determination of the orbit with a revised period of P=511 d and an additional small linear trend, indicative of another companion in a wide orbit. Moreover we show that the star exhibits low amplitude, solar like oscillations with frequencies around 3-4 d$^{-1}$ (34.7-46.3 $\mu$Hz). | astro-ph | astro-ph |
Astronomy&Astrophysicsmanuscript no. IotaDra
December 3, 2018
c(cid:13) ESO 2018
The Discovery of Stellar Oscillations in the K Giant ι Dra
Mathias Zechmeister1,2, Sabine Reffert3, Artie P. Hatzes2, Michael Endl4, and Andreas Quirrenbach3
1 Max-Planck-Institut fur Astronomie, Konigstuhl 17, 69117 Heidelberg, Germany
2 Thuringer Landessternwarte Tautenburg (TLS), Sternwarte 5, 07778 Tautenburg, Germany
3 ZAH-Landessternwarte, Konigstuhl 12, 69117 Heidelberg, Germany
4 McDonald Observatory, University of Texas, Austin, TX 78712, USA
Received; accepted
ABSTRACT
ι Dra (HIP 75458) is a well-known example for a K giant hosting a substellar companion since its discovery by Frink et al. (2002).
We present radial velocity measurements of this star from observations taken with three different instruments spanning nearly 8 years.
They show more clearly that the RV period is long-lived and coherent thus supporting the companion hypothesis. The longer time
baseline now allows for a more accurate determination of the orbit with a revised period of P = 511 d and an additional small linear
trend, indicative of another companion in a wide orbit. Moreover we show that the star exhibits low amplitude, solar like oscillations
with frequencies around 3-4 d−1 (34.7-46.3 µHz).
Key words. Stars: individual -- stars: planetary systems -- stars: oscillations
1. Introduction
Up to now long period radial velocity (RV) variations have been
detected in several K giants, e.g. β Gem (Hatzes et al. 2006;
Reffert et al. 2006), HD 47536 (Setiawan et al. 2003), HD 13189
(Hatzes et al. 2005) and ι Dra (Frink et al. 2002). Rotational
modulation can be excluded as a cause of these variations for
most of these giants due to lack of variations in photometry or
line bisectors with the RV period (Hatzes & Cochran 1998). The
most likely interpretation is orbiting, substellar companions.
Since the progenitor stars to planet hosting giant stars can
, these giant stars offer
have masses significantly larger than 1M
⊙
a way to investigate planet formation around massive stars. In
their evolved phase the Doppler induced effects of a planet are
easier to detect than in their main-sequence phase, due to cooler
effective temperatures and smaller rotational velocities.
But for evolved stars the determination of the mass is more
difficult. The most practical way for determining the stellar mass
is the use of evolutionary tracks, but those tracks converge in the
region of the giant branch for stars over a wide range of masses
and thus makes the derived masses uncertain.
Another possibility to measure the stellar mass is via astero-
seismology. This requires the investigation of stellar oscillations.
Photometric and RV variations with short periods from hours to
days have already been detected in some K and late G giants (e.g.
α Boo: Retter et al. 2003; Tarrant et al. 2007; α Ari: Kim et al.
2006; ǫ Oph: de Ridder et al. 2006; ξ Hya: Frandsen et al. 2002,
Stello et al. 2006; β UMi: Tarrant et al. 2008) and are consis-
tent with solar-like p-mode oscillations. Such oscillations have
also been measured for ensembles of red giants with photometry
of star rich regions or clusters (e.g. Gilliland 2008, Stello et al.
2008) which is a very efficient way to perform such asteroseis-
mologic studies.
Solar like oscillations have also been discovered in planet
hosting main sequence stars (e.g. µ Arae: Bazot et al. 2005;
ι Hor: Vauclair et al. 2008) and in the planet hosting K giant
β Gem for which Hatzes & Zechmeister (2007) gave, in combi-
nation with interferometric measurements of the angular diam-
eter, an estimation for the stellar mass completely independent
from evolutionary tracks. In a similar way we will do this here
for ι Dra.
2. Stellar Parameters
In Table 1 we summarize some direct measurements for the
K2III star ι Dra. The improved parallax from the Hipparcos cata-
log by van Leeuwen (2007) is given. Angular diameter estimates
based on spectrophotometry are available from the CHARM2
catalog (Richichi et al. 2005).
Table 1. Stellar parameters of ι Dra.
Parameter
Spectral typea
V a [mag]
MV [mag]
L [L
Parallaxb p [mas]
Distance d [pc]
Angular Diameterc θ [mas]
Radius R [R
⊙
]
]
⊙
Value
K2III
3.29
0.81
64.2 ± 2.1
32.23 ± 0.1
31.03 ± 0.1
3.73 ± 0.04
12.38 ± 0.17
a (ESA 1997)
b (van Leeuwen 2007)
c (Richichi et al. 2005)
The quantity which cannot be measured for single stars di-
rectly is the mass, which always relies on evolutionary tracks.
With the online tool1 from Girardi (see da Silva et al. 2006 and
Girardi et al. 2000 for a description), we derived some values
for M, log g and R. This tool uses as input parameters the visual
1 http://stev.oapd.inaf.it/cgi-bin/param
2
M. Zechmeister et al.: The Discovery of Stellar Oscillations in the K Giant ι Dra
Table 2. Stellar parameters of ι Dra from literature (T , log g and [Fe/H]) and derived with evolutionary tracks (M, log g and R).
Reference
Soubiran et al. (2008)
Prugniel et al. (2007)
Hekker & Mel´endez (2007)
Santos et al. (2004)
Gray et al. (2003)
Prugniel & Soubiran (2001)
Cenarro et al. (2001, 2007)
Allende Prieto & Lambert (1999)
McWilliam (1990)
Williams (1974)
T [K]
4552
4543
4605
4775 ± 113
4526
4491
4498
4466 ± 100
4490
4530 ± 100
log g
2.96
2.88
2.96
3.09 ± 0.40
2.64
2.57
2.38
2.24 ± 0.35
2.74
2.60 ± 0.25
[Fe/H]
0.16
0.19
0.07
0.13 ± 0.14
0.11
0.06
0.05
0.03 ± 0.11
0.29 ± 0.20
M [M
]
⊙
1.40 ± 0.24
1.41 ± 0.23
1.39 ± 0.24
1.71 ± 0.38
1.31 ± 0.24
1.24 ± 0.24
1.24 ± 0.24
1.05 ± 0.36
1.23 ± 0.24
1.40 ± 0.23
log g
2.40 ± 0.10
2.40 ± 0.10
2.45 ± 0.10
2.61 ± 0.12
2.37 ± 0.11
2.32 ± 0.10
2.33 ± 0.10
2.32 ± 0.10
2.39 ± 0.10
]
⊙
R [R
11.82 ± 0.63
11.94 ± 0.62
11.19 ± 0.58
10.34 ± 0.40
11.97 ± 0.65
12.23 ± 0.65
12.15 ± 0.66
12.22 ± 0.66
12.06 ± 0.64
magnitude V, parallax p, temperature T and metallicity Fe/H.
For the latter both quantities, T and [Fe/H], we adopt the values
in Table 2 as given by several authors. Allende Prieto & Lambert
(1999) gave no metallicity but derived themselves a mass of
M = 1.05M
which was used for the calculation of the com-
⊙
panion mass in Frink et al. (2002). This value is at the lower
mass range resulting from Table 2 (1.05 -- 1.71M
). This shows
⊙
how much the tracks depend on temperature and metallicity.
Furthermore, Table 2 allows a cross check of the log g-values
derived from spectroscopy and from evolutionary tracks. The
log g-values derived from spectroscopy are always higher, which
illustrates the difficulty of reliable mass determinations for K gi-
ants.
3. Observations
For our analysis we have RV measurements for ι Dra from
three independent instruments (Table 3 and Figure 1). The data
set from the 0.6m CAT (Coud´e Auxiliary Telescope) with the
Hamilton ´Echelle Spectrograph at Lick Observatory provides
nearly 8 years and thus the longest time baseline for the orbit
determination. The data up until March 2002 were already pub-
lished in Frink et al. (2002).
At
the Thuringia State Observatory (TLS -- Thuringer
Landessternwarte) ι Dra was observed on 43 nights spanning
200 d. The Alfred-Jensch 2m telescope is equipped with a Coud´e
´Echelle Spectrograph with a wavelength coverage from 4630
to 7370 Å. These observations primarily were carried out to
look for stellar oscillations with short periods. For this pro-
gram on some nights up to 20 spectra (even 61 on the night of
JD=2454193) were taken (Figure 3).
With a rather short time baseline of 8 days some spectra were
obtained in June 2005 with the 2.7 m telescope at the McDonald
Observatory and the "2dcoude" spectrograph. This instrument
provides a wavelength range from 3600 Å to 1 µm. The RVs
were calculated using the "Austral" program (Endl et al. 2000).
The whole data set is plotted in Figure 4.
All three instruments have a resolving power around R =
= 60 000 and utilize an iodine absorption cell for the wave-
λ
∆λ
length reference. Table 3 lists the time coverage, time span T ,
number of spectra N and average precision of the radial veloc-
ity measurements σRV. These measurement errors are internal
errors due to instrumental effects recorded with the iodine lines.
Comparing with point to point scatter, e.g. in Figure 3, the errors
appear reasonable.
Table 3. Journal of observations.
Data Set
Coverage
CAT
TLS
McD
2000.39 -- 2008.38
2006.99 -- 2007.59
2005.44 -- 2005.46
T [d]
2918
221
8
N
147
280
62
σRV [m/s]
4.3
3.3
3.9
4. Orbital Solution
The parameters of the orbit were determinated by weighted least
squares (χ2-Fit) with the differential correction method (Sterne
1941). We fitted simultaneously five Keplerian orbital elements,
a linear trend and three offsets for the combined data sets as the
measurements give relative velocities and the three instruments
have different zero points. When weighting the measurements
one has to take into account that not only the measurement errors
introduce an error into the parameter determination but also the
jitter due to pulsations: σ2
P. The stellar oscillations
are discussed in the next chapter but we mention here that σP
is of the order of 10 m/s. We adopt this value which is larger
than the typical measurement error and therefore leads to a more
equal weighting of all measurements in our fit.
RV,i + σ2
i = σ2
The resulting orbital elements are listed in Table 4 and the
RV curve is plotted in Figure 1. In the old solution the short time
baseline and the unrecognizable linear trend had resulted in an
overestimated period (P = 536 ± 6 d: Frink et al. 2002). The lin-
ear trend is −13.8±1.1 m/s/yr and may be part of a longer period
caused by a further companion. For the calculation of the com-
panion mass we assumed a stellar mass of 1.4M
(33% higher
⊙
than in Frink et al. 2002).
Table 4. Orbital parameters for the companion to ι Dra.
Parameter
Period P [d]
Amplitude K [m/s]
Periastron time T0 [JD]
Longitude of periastron ω [◦]
Eccentricity e
Linear trend [m/s/yr]
Mass functiona f (m) [M
⊙
Semi-major axis a [AU]
companion mass m sin i [MJup]
]
Value
510.88 ± 0.15
299.9 ± 4.3
88.7 ± 1.4
-13.8 ± 1.1
2452013.94 ± 0.48
0.7261 ± 0.0061
(4.64 ± 0.33) · 10−7
1.34
10.3
a
f (m) = (m sin i)3
(M+m)2 = P
2πG (K √1 − e2)3
Figure 2 illustrates the radial velocity phased to the orbital
period. The total remaining scatter is 13.9 m/s for all data sets
M. Zechmeister et al.: The Discovery of Stellar Oscillations in the K Giant ι Dra
3
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
Time [Year]
TLS
CAT
McD
400
300
200
100
0
−100
−200
−300
]
s
/
m
[
V
R
−400
51500
52000
52500
53000
53500
54000
54500
JD − 2400000
Time [Year]
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
TLS
CAT
McD
80
60
40
20
0
−20
]
s
/
m
[
V
R
−40
51500
52000
52500
53000
53500
54000
54500
JD − 2400000
Fig. 1. Radial velocity measurements for ι Dra from three data sets and the orbital solution as solid line. The lower panel shows the
residuals after subtraction of the orbital solution and the linear trend. (See online version for a colored figure.)
(and for the individual sets, which cover very different time
scales: 15.4 m/s (CAT), 13.9 m/s (TLS) and 11.4 m/s (McD)).
This scatter is a factor of 3-4 greater than the measurement er-
rors given in Table 3 and is largely due to stellar oscillations as
we will explain below.
5. Short Period Oscillations
For the analysis of variations on short time scales we subtracted
the trend and orbit from the data sets. The remaining residuals
during some nights for the TLS and McD data sets are shown in
Figures 3 and 4, respectively. Clearly one can see RV variations
with an amplitude of few 10 m/s.
We performed the frequency analysis of the orbit residuals
for each data set individually due to their different time sam-
pling. For this we used the prewhitening procedure provided by
the program Period04 (Lenz & Breger 2004). A Fourier trans-
formation was performed for the data, the dominant frequency
extracted and subtracted from the data. This procedure can be
repeated on the subsequent residuals until no further significant
frequencies can be found. The prewhitening has the advantage
that confusing alias peaks (here mainly 1 day aliases) of the ex-
tracted frequency are removed from periodograms.
Figures 5 and 6 show the periodograms in each prewhiten-
ing step for the TLS and McD residuals, respectively. The TLS
residuals show peaks at 0.040 d−1 and 0.004 d−1, which seem
to be 1 month-aliases (∆ f = 0.036 d−1) to each other. Neither
frequency is found in the other data sets. However, the corre-
sponding long periods (P = 25 d and P = 250 d) are not of
further interest for our investigation of short period oscillations.
Subtraction of the lower frequency ( f0 = 0.004 d−1, 12.33 m/s)
flattens the TLS orbit residuals and removes the 1 day aliases
of f0 (second panel in Figure 5). This prewhitening step reveals
effectively an excess at frequencies around 3.8 d−1. Extraction
of the two frequencies f1=3.45 d−1 (4.77 m/s) and f2=4.23 d−1
(4.26 m/s) lowers the excess power considerably. A third fre-
quency f3 =3.75 d−1 (3.86 m/s) is already near an amplitude sig-
nal to noise ratio of S/N=4 (the mean noise level is 0.865 m/s
in the range of 10-20 d−1). This threshold, as suggested by
Kuschnig et al. (1997), can be used as a criterion for stopping
the prewhitening procedure.
The fit with the frequencies is also drawn in Figure 3. Nights
with many data points fit well, while some discrepancies exist for
nights with sparse measurements. This may indicate that more
frequencies are present or that the modes have a finite lifetime
due to stochastic excitation mechanisms (e.g Barban et al. 2007).
M. Zechmeister et al.: The Discovery of Stellar Oscillations in the K Giant ι Dra
TLS
CAT
McD
]
s
/
m
[
V
R
30
20
10
0
-10
-20
-30
0
0.2
0.4
0.6
0.8
1
TLS
CAT
McD
53530
53531
53532
53533
53534
53535
53536
53537
53538
JD - 2400000
Fig. 4. Radial velocity during 8 consecutive nights measured at
McD (orbit subtracted).
4
]
s
/
m
[
V
R
]
s
/
m
[
V
R
300
200
100
0
−100
−200
−300
60
40
20
0
−20
−40
0
0.2
0.4
0.6
Orbital Phase
0.8
1
Fig. 2. Radial velocity measurements for ι Dra from three data
sets phased to the orbital period. The lower panel shows the orbit
residuals. (See online version for a colored figure.)
30
20
10
0
30
20
10
0
10
0
-10
-20
-30
10
0
-10
-20
-30
10
0
-10
-20
-30
]
s
/
m
[
V
R
30
20
10
0
10
0
-10
-20
10
0
-10
-20
-30
10
0
-10
-20
-30
192.3 192.4 192.5 192.6 192.7
251.3 251.4 251.5 251.6 251.7
253.3 253.4 253.5 253.6 253.7
278.3 278.4 278.5 278.6 278.7
191.3 191.4 191.5 191.6 191.7
193.3 193.4 193.5 193.6 193.7
252.3 252.4 252.5 252.6 252.7
276.3 276.4 276.5 276.6 276.7
307.3 307.4 307.5 307.6 307.7
JD - 2454000
The higher amplitudes of the frequencies in the McD data
may be an effect of the short time baseline and the finite mode
lifetime of solar-like oscillations seen in main sequence stars and
red giants (Stello et al. 2006; Carrier et al. 2007).
]
s
/
m
[
e
d
u
t
i
l
p
m
A
16
14
12
10
8
6
4
2
0
7
6
5
4
3
2
1
0
7
6
5
4
3
2
1
0
7
6
5
4
3
2
1
0
1
0.8
0.6
0.4
0.2
0
data-orbit
data-orbit-f0
data-orbit-f0-f1
f1
f2
data-orbit-f0-f1-f2
window
1
0.8
0.6
0.4
0.2
0
0
0.2
0.4
0.6
0.8
1
Fig. 3. Radial velocity during several good nights measured at
TLS (orbit subtracted). The solid curve is the fit with the TLS
frequencies ( f0, f1 and f2).
For the McD orbit residuals an excess of power (upper panel
in Figure 6) can be seen in the same region as for the TLS data.
The peaks at frequencies >15 d−1 in the McD data set are due to
the sparser sampling leading to a stronger aliasing at higher fre-
quencies. Extracted are two frequencies, f1 =3.81 d−1 (8.45 m/s)
and f2=4.03 d−1 (8.90 m/s), which are also illustrated along with
the periodogram of the remaining residuals in the third panel of
figure 6.
0
1
2
3
4
5
6
7
8
9 10 11 12 13 14 15 16 17 18 19 20
Frequency [1/d]
Fig. 5. Amplitude periodograms of the TLS RV data at each
prewhitening step. The fourth panel shows the periodogram of
the residuals in comparison with the extracted frequencies and
the last panel shows the window function. The inset is an en-
largement where the one month alias can be seen.
The CAT residuals show a 555 d period, which is close to
the orbital period. However the data set has insufficient time res-
olution to look for short period variability (< 1 day) since its
Nyquist frequency is only 0.5 d−1.
M. Zechmeister et al.: The Discovery of Stellar Oscillations in the K Giant ι Dra
5
data-orbit
A more accurate mass estimation could be done with the
scaling relation for the frequency splitting:
8
6
4
2
0
8
6
4
2
0
8
6
4
2
0
1
]
s
/
m
[
e
d
u
t
i
l
p
m
A
0.8
0.6
0.4
0.2
0
f1
f2
data-orbit-f1
data-orbit-f1-f2
window
∆ f0 = s M/M
(R/R
)3 · 11.66 d−1
⊙
⊙
(3)
since the radius can be obtained from angular diameter measure-
ments. With the assumed mass of 1.4M
for ι Dra the expected
⊙
frequency splitting is around 0.33 d−1. But the examination of
the frequency splitting needs more effort and much more data
for the identification of many modes. Unfortunately, our current
data sets are not suitable for this kind of analysis.
7. Conclusion
We revised the orbit solution for the companion of ι Dra. The
orbital period is P = 511 d, somewhat lower than the value in
the discovery paper by Frink et al. (2002). Furthermore there is a
linear trend of −13.8 m/s/yr present, possibly caused by another
companion.
An excess of power around 3.8 d−1 was found independently
in two data sets, taken at different times and with very different
sampling. The amplitude and the location of the power excess
are consistent with solar-like oscillations.
Our analysis of the short period oscillations indicates a
, considerably higher
somewhat high stellar mass of M = 2.2M
⊙
of Allende Prieto & Lambert (1999) or the
than the 1.05M
⊙
).
stellar masses derived from the Girardi track (1.2 − 1.7M
⊙
However, this is a preliminary result based on limited data.
Assuming 1.4M
for the mass ι Dra yields a minimum mass of
⊙
10MJup for the companion.
Our RV measurements for ι Dra indicate that it shows multi-
periodic stellar oscillations. This means that an asteroseismic
analysis can yield an accurate mass. This is best done by mea-
suring the frequency splitting in the p-mode oscillation spectrum
which requires more measurements and better sampling than the
data presented here.
Acknowledgements. We gratefully acknowledge assistance from Debra Fischer
and Geoffrey Marcy in obtaining precise radial velocities at Lick Observatory,
and we thank the staff at Lick Observatory for their extraordinary dedication and
support. In addition, we would like to thank the CAT observers David Mitchell,
Saskia Hekker and Christian Schwab for obtaining some of the observations used
in this work.
We thank the anonymous referee for constructive criticism and valuable
hints which helped improve this paper.
References
Allende Prieto, C. & Lambert, D. L. 1999, A&A, 352, 555
Barban, C., Matthews, J. M., de Ridder, J., et al. 2007, A&A, 468, 1033
Bazot, M., Vauclair, S., Bouchy, F., & Santos, N. C. 2005, A&A, 440, 615
Bedding, T. R. & Kjeldsen, H. 2003, Publications of the Astronomical Society
of Australia, 20, 203
Carrier, F., Kjeldsen, H., Bedding, T. R., et al. 2007, A&A, 470, 1059
Cenarro, A. J., Gorgas, J., Cardiel, N., et al. 2001, MNRAS, 326, 981
Cenarro, A. J., Peletier, R. F., S´anchez-Bl´azquez, P., et al. 2007, MNRAS, 374,
664
da Silva, L., Girardi, L., Pasquini, L., et al. 2006, A&A, 458, 609
de Ridder, J., Barban, C., Carrier, F., et al. 2006, A&A, 448, 689
Endl, M., Kurster, M., & Els, S. 2000, A&A, 362, 585
ESA. 1997, VizieR Online Data Catalog, 1239, 0
Frandsen, S., Carrier, F., Aerts, C., et al. 2002, A&A, 394, L5
Frink, S., Mitchell, D. S., Quirrenbach, A., et al. 2002, ApJ, 576, 478
Gilliland, R. L. 2008, AJ, 136, 566
Girardi, L., Bressan, A., Bertelli, G., & Chiosi, C. 2000, A&AS, 141, 371
Gray, R. O., Corbally, C. J., Garrison, R. F., McFadden, M. T., & Robinson, P. E.
2003, AJ, 126, 2048
0
1
2
3
4
5
6
7
8
9 10 11 12 13 14 15 16 17 18 19 20
Frequency [1/d]
Fig. 6. Amplitude periodograms of the McD RV data at each
prewhitening step. The third panel shows also the extracted fre-
quencies and the last panel shows the window function.
6. Discussion
We compare our results with the scaling relations for solar-like
oscillations from Kjeldsen & Bedding (1995). The oscillation
velocity amplitude vosc and the frequency fmax of the strongest
mode of the 5-minute-oscillations in the sun scales to other stars
as:
L/L
⊙
M/M
vosc =
⊙ · 0.234 m/s
fmax = 62 d−1 (T /5777K)3.5
vosc/ m s−1
.
(1)
(2)
These relations are valid for a wide range of convective stars
(Bedding & Kjeldsen 2003) and seem to be valid for other K gi-
ants (Stello et al. 2008). For a power excess frequency of around
3.8 d−1, which is present in two data sets, the expected ampli-
tude should be 6.75 m/s according eq.(2) (based on T = 4490 K,
McWilliam (1990)). This agrees with the excess power seen
in the periodograms and is comparable to the amplitude of f1
(4.77 m/s (TLS) and 8.90 m/s (McD)). So this frequency satisfies
the relation between amplitude and frequency typical for solar-
like oscillations.
Under the assumption that the simple scaling relations are
valid, the combination of both equations to L/M ∝ T 3.5/ fmax im-
for this fre-
plies a luminosity to mass ratio of L/M = 29L
from the Hipparcos catalog one
quency. Using L = 64.2 ± 2.1L
⊙
(or with fmax ∝ g/ √T a log g
would estimate a mass of 2.2M
⊙
of 2.54). Even if we roughly estimate an uncertainty of ±0.6M
⊙
due to strong 1 day aliasing (i.e. derived from ∆ fmax = ±1 d−1),
this value seems too high in comparison with Table 2. For a mass
the expected frequency is 2.4 d−1. We cannot exclude
of 1.4M
⊙
that we miss such a mode as 0.42 d periods are difficult to detect
in short nights with single site observations.
⊙/M
⊙
6
M. Zechmeister et al.: The Discovery of Stellar Oscillations in the K Giant ι Dra
Hatzes, A. P. & Cochran, W. D. 1998, in Astronomical Society of the Pacific
Conference Series, Vol. 134, Brown Dwarfs and Extrasolar Planets, ed.
R. Rebolo, E. L. Martin, & M. R. Zapatero Osorio, 312 -- +
Hatzes, A. P., Cochran, W. D., Endl, M., et al. 2006, A&A, 457, 335
Hatzes, A. P., Guenther, E. W., Endl, M., et al. 2005, A&A, 437, 743
Hatzes, A. P. & Zechmeister, M. 2007, ApJ, 670, L37
Hekker, S. & Mel´endez, J. 2007, A&A, 475, 1003
Kim, K. M., Mkrtichian, D. E., Lee, B.-C., Han, I., & Hatzes, A. P. 2006, A&A,
454, 839
Kjeldsen, H. & Bedding, T. R. 1995, A&A, 293, 87
Kuschnig, R., Weiss, W. W., Gruber, R., Bely, P. Y., & Jenkner, H. 1997, A&A,
328, 544
Lenz, P. & Breger, M. 2004, in IAU Symposium, Vol. 224, The A-Star Puzzle,
ed. J. Zverko, J. Ziznovsky, S. J. Adelman, & W. W. Weiss, 786 -- 790
McWilliam, A. 1990, ApJS, 74, 1075
Prugniel, P. & Soubiran, C. 2001, A&A, 369, 1048
Prugniel, P., Soubiran, C., Koleva, M., & Le Borgne, D. 2007, ArXiv
Astrophysics e-prints
Reffert, S., Quirrenbach, A., Mitchell, D. S., et al. 2006, ApJ, 652, 661
Retter, A., Bedding, T. R., Buzasi, D. L., Kjeldsen, H., & Kiss, L. L. 2003, ApJ,
591, L151
Richichi, A., Percheron, I., & Khristoforova, M. 2005, A&A, 431, 773
Santos, N. C., Israelian, G., & Mayor, M. 2004, A&A, 415, 1153
Setiawan, J., Hatzes, A. P., von der Luhe, O., et al. 2003, A&A, 398, L19
Soubiran, C., Bienaym´e, O., Mishenina, T. V., & Kovtyukh, V. V. 2008, A&A,
480, 91
Stello, D., Bruntt, H., Preston, H., & Buzasi, D. 2008, ApJ, 674, L53
Stello, D., Kjeldsen, H., Bedding, T. R., & Buzasi, D. 2006, A&A, 448, 709
Sterne, T. E. 1941, Proceedings of the National Academy of Science, 27, 175
Tarrant, N. J., Chaplin, W. J., Elsworth, Y., Spreckley, S. A., & Stevens, I. R.
2007, MNRAS, 382, L48
-- . 2008, A&A, 483, L43
van Leeuwen, F. 2007, Hipparcos,
the New Reduction of the Raw Data
(Hipparcos, the New Reduction of the Raw Data. By Floor van Leeuwen,
Institute of Astronomy, Cambridge University, Cambridge, UK Series:
Astrophysics and Space Science Library, Vol. 350 20 Springer Dordrecht)
Vauclair, S., Laymand, M., Bouchy, F., et al. 2008, A&A, 482, L5
Williams, P. M. 1974, MNRAS, 167, 359
List of Objects
'HIP 75458' on page 1
|
0711.2906 | 2 | 0711 | 2008-06-06T21:21:56 | Circumscribing Late Dark Matter Decays Model Independently | [
"astro-ph",
"hep-ph"
] | A number of theories, spanning a wide range of mass scales, predict dark matter candidates that have lifetimes much longer than the age of the universe, yet may produce a significant flux of gamma rays in their decays today. We constrain such late decaying dark matter scenarios model-independently by utilizing gamma-ray line emission limits from the Galactic Center region obtained with the SPI spectrometer on INTEGRAL, and the determination of the isotropic diffuse photon background by SPI, COMPTEL and EGRET observations. We show that no more than ~5% of the unexplained MeV background can be produced by late dark matter decays either in the Galactic halo or cosmological sources. | astro-ph | astro-ph | Circumscribing Late Dark Matter Decays Model Independently
Department of Physics, Ohio State University, Columbus, Ohio 43210 and
Center for Cosmology and Astro-Particle Physics, Ohio State University, Columbus, Ohio 43210
Hasan Yuksel and Matthew D. Kistler
(Dated: May 8, 2008)
A number of theories, spanning a wide range of mass scales, predict dark matter candidates that have lifetimes
much longer than the age of the universe, yet may produce a significant flux of gamma rays in their decays
today. We constrain such late decaying dark matter scenarios model-independently by utilizing gamma-ray line
emission limits from the Galactic Center region obtained with the SPI spectrometer on INTEGRAL, and the
determination of the isotropic diffuse photon background by SPI, COMPTEL and EGRET observations. We
show that no more than ∼5% of the unexplained MeV background can be produced by late dark matter decays
either in the Galactic halo or cosmological sources.
8
0
0
2
n
u
J
6
]
h
p
-
o
r
t
s
a
[
2
v
6
0
9
2
.
1
1
7
0
:
v
i
X
r
a
PACS numbers: 95.35.+d, 13.35.Hb, 12.60.-i
I.
INTRODUCTION
Dark matter continues to live up to its name [1], despite
accumulated evidence of its existence from observations of
large-scale structure formation, galaxy cluster mass-to-light
ratios, and galactic rotation curves. An attractive approach to-
wards revealing dark matter's particle identity is to search for
its signature in radiation backgrounds, either from the Milky
Way or in the isotropic diffuse photon background (iDPB),
which can contain both cosmological and Galactic halo con-
tributions. Dark matter might be comprised of particles that
can decay with finite lifetimes much longer than the age of
the universe. In such scenarios, the resultant fluxes of decay
products depend on the amount of dark matter present alone,
as opposed to self-annihilation, which, being dependent on
particle density squared, is very sensitive to assumptions con-
cerning details of dark matter clustering.
A wide variety of decaying dark matter models have been
examined in regards to their observable implications [2].
Among late decaying dark matter models, sterile neutrinos
with multi-keV masses have been extensively studied as dark
matter candidates [3], with strong constraints placed on their
decays, e. g. Refs. [3, 4, 5] and references therein. The de-
cay of moduli dark matter [6] with masses of several hundred
keV may contribute to the sub-MeV iDPB. The dark matter
model of Ref. [7], inspired from minimal universal extra di-
mensions or supersymmetry [8], with a mass scale of hun-
dreds of GeV, is advocated as the source of the iDPB in the
MeV range [9], which has yet to be accounted for with con-
ventional sources (e.g., supernovae [10] or active galactic nu-
clei [11]) or more exotic mechanisms [12]. Similarly, decay-
ing gravitino dark matter in R-parity breaking vacua [13], with
multi-GeV masses, has been suggested as an explanation of
iDPB spectral features in the GeV range [14, 15].
Rather than focusing on a particular model, we first con-
sider a generic decaying dark matter scenario in which the de-
cay of the parent particle is dominated by a monochromatic
photon emission. We assume that the lifetime of the par-
ent particle, τ, is much longer than the age of the universe
(τ0 ≃ 4.5 × 1017 s), thus its cosmological abundance has
not changed significantly since the time of dark matter decou-
pling. The decay under consideration is χ → χ′ + γ, where γ
is a monochromatic photon emitted with energy ε. In general,
1030
1028
1026
1024
1022
]
s
V
e
G
[
τ
m
χ
R
-
p
a
rit
y
b
r
e
a
k
i
n
g
v
a
c
u
a
m
U
E
D
ν
s
D
M
Excluded by
Milky Way
γ-ray Line Search
O verproduction of
the Isotropic Diffuse P hoton B ackground
τ and ε will depend on the masses of the parent and the daugh-
ter particles (mχ, mχ′) and their splitting, ∆m = mχ − mχ′.
The flux of photons from dark matter decays is inversely
proportional to both the particle lifetime (fixing the decay rate
per particle as specified by a particular theoretical model) and
the mass of an individual particle (yielding the total number of
particles in a fixed amount of dark matter). Thus, gamma-ray
observations allow us to place constraints only on the degen-
erate product mχτ versus ε, as we display in Fig. 1. As we
will discuss in detail, below the jagged line between 0.02 --
8 MeV, the gamma-ray line signal from the Galactic Cen-
ter (GC) region due to dark matter decays violates the corre-
sponding limit obtained with the SPI spectrometer on INTE-
1020
10-5
10-4
10-3
10-2
10-1
ε [GeV]
100
101
102
FIG. 1: Model-independent constraints on the product of mass and
lifetime, mχτ , versus the energy carried away by the monochromatic
photon emission, ε, for a generic late-decaying dark matter model:
′ + γ. Regions excluded by either the gamma-ray line emis-
χ → χ
sion limits from the Galactic Center region or overproduction of the
isotropic diffuse photon background are shown, together with pre-
ferred ranges of parameters from three well-studied models.
GRAL satellite [16]. Additionally, the iDPB, as determined
from SPI [17], COMPTEL [18] and EGRET [20, 21] data,
is overproduced (assuming it is fully accounted by late dark
matter decays in a given energy band) in the triangular region,
even disregarding any contributions from known astrophysi-
cal sources. We also show three representative scenarios, in-
spired by the theories of sterile neutrinos, R-parity breaking
vacua, and mUED. Since mχ, mχ′ and ∆m are not necessar-
ily predetermined, they may be adjusted to yield the displayed
curves relating mχτ and ε.
II. MILKY WAY GAMMA-RAY LINE SEARCH
A monochromatic line will be most detectable locally,
where cosmological redshifting is of no concern. Fortunately,
a search for diffuse gamma-ray line emission in the energy
range 0.02-8 MeV from the GC region has been conducted by
Teegarden and Watanabe using the SPI spectrometer on the
INTEGRAL satellite [16], which recovered the known astro-
physical diffuse line fluxes, such as the 511 keV positron an-
nihilation line [22]. The excellent energy resolution of SPI en-
abled them to place very strict constraints on potential uniden-
tified emission lines, with an energy dependent 3.5 σ flux
limit, Flim(E), from an angular region within a 13◦ radius
of the GC (which we refer as the GC region). This limit, re-
produced in the top panel of Fig. 2, can be compared to the
expected gamma-ray flux arising from late dark matter decays
in the GC region, which we calculate following the methods
in Ref. [5].
10-3
10-4
Excluded by MilkyWay
γ-ray Line Search
]
1
-
s
2
-
m
c
[
)
E
(
m
i
l
F
]
1
-
10-5
10-4
SPI
r
s
1
-
2
-
s
m
c
V
e
G
[
E
d
/
Φ
d
2
E
10-5
10-6
COMPTEL
EGRET
10-7
10-5
10-4
10-3
10-2
10-1
E [GeV]
100
101
102
FIG. 2: Top: Limits on the diffuse gamma-ray line emission from
the Galactic Center region (an angular region within a 13◦ radius) as
adopted from Ref. [16]. Bottom: Representative measurements of
the diffuse photon background from SPI [17], COMPTEL [18] and
EGRET [21] in the energy range around 0.01 MeV -- 100 GeV. The
thick solid line, summarizing the overall trend of the data, is to be
compared to predictions of decaying dark matter scenarios.
We first define a dimensionless line-of-sight integral at an
angle ψ relative to the GC,
2
J (ψ) =
1
ρscRsc Z ℓmax
0
dℓ ρ(cid:16)pR2
sc − 2 ℓ Rsc cos ψ + ℓ2(cid:17) ,
(1)
where ρ is the density of the dark matter in the halo as a func-
tion of the distance from the GC. This is normalized to the
dark matter density (ρsc = 0.3 GeV cm−3) at the solar circle
(Rsc = 8.5 kpc) so that ρscRsc ≃ 8 × 1021 GeV cm−2. Note
that this arbitrary normalization is needed to make J dimen-
sionless and will be canceled-out later. The upper limit of this
integration,
ℓmax = q(R2
MW − sin2 ψR2
sc) + Rsc cos ψ ,
(2)
depends on RMW , the assumed size of the halo. J is rela-
tively insensitive to ℓmax as long as RMW is large. The inten-
sity of photons (number flux per solid angle) from the same
direction,
I(ψ) =
ρscRsc
4πmχτ
J (ψ) ,
(3)
can be integrated over a circle of radius ψ around the GC (cov-
ering a patch of area ∆Ω = 2π(1 − cos ψ)) to obtain the cor-
responding total flux,
F = Z∆Ω
dΩ′ I(ψ′) =
ρscRsc
4πmχτ Z∆Ω
dΩ′ J (ψ′)
(4)
The limit reported in Ref. [16] has been obtained by sub-
tracting the average flux measured at regions away from the
GC region (ψ > 30◦) from the average flux measured in-
side the GC region (ψ < 13◦) to eliminate instrumental back-
grounds. Thus, the constraining power of this limit for decay-
ing dark matter scenarios depends on the enhancement of the
expected signal towards the GC region. Both theoretical and
observational studies strongly suggest that the central regions
of dark matter halos are significantly denser and, moreover,
the column depth is higher towards the GC direction relative
to off-axis lines-of-sight. We have reproduced the impact of
this subtraction (see Ref. [5] for details) by calculating a pa-
rameter
ζlim = Z∆Ω
dΩ′ [J (ψ′) − J >30◦ ] .
(5)
which ranges between ∼ 0.5 − 1.5 for various dark mat-
ter halo fitting profiles commonly used in the literature [23].
Here J >30◦ is the average of J away from the GC region.
We also note that the results that we adopted from Ref. [16]
are based on an assumption that the expected line signal has
a Gaussian source profile, while a flat source profile could
yield limits that are weaker by up to a factor of ∼2 (e.g., see
Fig. 12 of Ref. [29]). Moreover, one would expect to see these
limits improve as the amount of available data increases in
time [29]. In the rest of our study, we choose a conservative
value, ζlim ≃ 0.5, which can be realized only for profiles that
are rather flat inside the solar circle. While this mostly pro-
tects our conclusions from uncertainties in the halo profile, our
subsequent result can be easily rescaled for a different value.
The predicted gamma-ray emission line flux due to dark
matter decays at a given ε must not exceed the corresponding
limits from the GC region, thus
component is
3
(8)
Iiso =
ρscRsc
4πmχτ
Jiso .
ρscRsc
4πmχτ
ζlim < Flim(E = ε) .
(6)
Rearranging this equation yields our model independent con-
straint,
mχτ >
ρscRscζlim
4πFlim(ε)
≃
3 × 1020 GeV cm−2
Flim(ε)
,
(7)
as shown in Fig. 1 (region below the jagged line). The ex-
pected dark matter decay flux is inversely proportional to
mχτ, which leads to an overproduction of gamma rays for
mχτ . 1025 GeV s in the energy range 0.02-8 MeV. Thus
the area below the jagged line is excluded by the the diffuse
gamma-ray line emission limits from the GC region.
III.
ISOTROPIC DIFFUSE PHOTON BACKGROUND
While stringent limits on line emission from the GC re-
gion are only available in a rather limited energy range (0.02 --
8 MeV), the iDPB is measured over a broad range by many
instruments.
In the bottom panel of Fig. 2, we display
three recent determinations of iDPB in different ranges of
energy from SPI [17], COMPTEL [18] and EGRET [21],
which are consistent with others measurements (see, e.g.,
Ref. [19, 24]). The thick dotted line represents the global
trend of the data to be used for comparison. We choose the
terminology "isotropic diffuse" photon background (iDPB),
as opposed to "cosmic" or "extragalactic", since the contribu-
tion from sources in the Milky Way or its halo is not clear, and
iDPB can include gamma-ray line signals that could not have
been resolved by COMPTEL or EGRET. While it is gener-
ally thought that AGN are responsible for the emission in the
∼ keV [25] and ∼ GeV [26] ranges, the origin of the iDPB,
especially in the MeV regime, is far from being settled, with
various scenarios having been entertained [10, 11, 12]. It is
then of interest to determine just how much of the iDPB can
possibly be accounted for by late decaying dark matter.
A. Dark Matter Decays in the Halo
While the photon signal from dark matter decays in the
Galactic halo is enhanced towards the GC, as has been utilized
for our constraints in the earlier section, it also contains an ap-
parently isotropic contribution. The limited energy resolution
of past gamma-ray detectors could not distinguish monochro-
matic line emission from the Galactic halo from a truly cos-
mological signal. The intensity of the isotropic halo contri-
bution, Iiso, can be estimated from a line of sight integration
in the anti-GC direction, Jiso = J (180◦) ∼ 1, as this is the
minimum contribution from the dark matter halo of the Milky
Way. Regardless of the underlying halo profile, this number
is relatively robust, being mostly dependent on the dark mat-
ter density at the solar circle. The intensity of this isotropic
We present a representative spectrum for this isotropic sig-
nal in Fig. 3 (dotted line), after convolution with a Gaus-
sian of ∼ 10% width to simulate the energy resolution of
a typical detector. We have chosen ε = 1 MeV, with
mχτ = 7 × 1024 GeV s, the maximum value allowed by the
the line emission bounds from the GC region (Fig. 1). For
these parameters, the isotropic contribution of the dark mat-
ter decays in the Galactic halo alone to the iDPB is less than
2% (in a bin of logarithmic width 0.4 dex centered around
ε = 1 MeV). Note that the average flux expected from the
decays in the Galactic halo (which is more directional, peak-
ing toward the GC region) can be at most several times larger
than this isotropic component since we are dealing with de-
caying dark matter particles (contrary to self-annihilating dark
matter, which is highly sensitive to the details of dark matter
clustering).
B. Cosmological Dark Matter Decays
We now evaluate the contribution of truly cosmological
dark matter decays to the iDPB. For late decaying particles
(τ ≫ τ0), the comoving dark matter density has remained
nearly constant since the early universe. The comoving de-
cay rate is then simply proportional to the dark matter fraction
]
r
s
1
-
1
-
s
2
-
m
c
V
e
G
[
E
d
/
Φ
d
2
E
10-4
10-5
10-6
10-7
10-8
0.1
Isotropic
Diffuse Photon Background
0.3
E [MeV]
1
3
FIG. 3: Photon spectrum from isotropic Galactic halo decays (dotted
line) for ε = 1 MeV, with mχτ ≃ 7 × 1024 GeV s chosen from
Fig. 1 such that the line emission bounds from the Galactic Center
region are saturated. Also displayed are the spectra from cosmolog-
ical decays (dashed line) and the total spectrum (solid line), which
falls well below the isotropic diffuse photon background (thick solid
line).
(Ωχ ≃ 0.25) of the critical density of the universe, ρc, and
is given as ρcΩχ/(mχτ ). The diffuse gamma-ray flux (per
solid angle per unit energy) arising from the decays can be
calculated by considering the contributions from all redshifts
(analogous to [27]),
dΦ
dE
=
1
4π
c
H0 Z
dz
h(z)
ρcΩχ
mχτ
δ(E(1 + z) − ε) ,
(9)
where h(z) = [(1 + z)3ΩM + ΩΛ]1/2, ΩM ≃ 0.3, ΩΛ ≃ 0.7,
H0 = 70 km s−1Mpc−1, and c = 3 × 1010 cm s−1 (so that
c/H0 ≃ 1.3 × 1028cm and ρc = 5.3 × 10−6 GeV cm−3).
The integration can be eliminated after using the δ-function
identity; δ(ax − b) = δ(x − b/a)/a, simplifying the result to
dΦ
dE
=
1
4π
c
H0
ρcΩχ
mχτ
1
E
,
(10)
Θ(ε − E)
p(ε/E)3ΩM + ΩΛ
where h is substituted and Θ is a step function. We show this,
using the same parameters as in the preceding subsection and
again smoothing with a ∼ 10% Gaussian, in Fig. 3 (dashed
line). As seen in the figure, this cosmological flux is slightly
lower than the isotropic contribution from the Galactic halo,
and their sum (solid line) still falls well short of the observed
signal, restricting their combined contribution to the iDPB to
be less than 4%.
To quantify and generalize our observations, we calculate
the expected total (cosmological plus isotropic Galactic halo)
spectrum for all values of mχτ and compare to the iDPB (as
denoted by the thick trend curve in Fig. 2), integrating both
in a bin of logarithmic width 0.4 dex centered around ε. This
choice encompasses most of the expected signal where the
decay spectrum peaks, and both exceeds the experimental en-
ergy resolution and the uncertainties on the determination of
the iDPB. In Fig. 1, the region in which dark matter decays
overproduce the iDPB is shown (triangular region).
Above this region, decaying dark matter alone cannot fully
account for the iDPB. In fact, since there should be addi-
tional contributions from AGN at both low and high ener-
gies [25, 26], the actual bound on the parameter mχτ will
be even more stringent than the one presented. Combining
the iDPB overproduction constraint and the gamma-ray line
emission limit from the GC region model-independently ex-
cludes a sizable region in the parameter space of mχτ versus
ε, with the latter picking up when the former is exhausted at
ε ≃ 8 MeV.
IV. DECAYING DARK MATTER MODELS
While we derive our constraints for a decay scenario that
is dominated by monochromatic photon emission, there may
be additional modes of decay or self-annihilations producing
other signals. Our constraints on the lifetime of the dark mat-
ter candidate via monochromatic photon emission could be
generalized to the total lifetime including other decay chan-
nels, as long as the latter is long enough to justify the assump-
tion that the cosmological abundance of the parent particle has
not changed significantly.
4
For the generic decay we are considering, the energy of the
emitted photon is dictated by the splitting, ∆m, as follows.
When ∆m ≪ mχ′ (or equivalently mχ ≃ mχ′), the recoil
of the daughter can be neglected, so that ε → ∆m. For
∆m ≫ mχ′ (or mχ ≫ mχ′), two relativistic particles are
produced, so that ε → ∆m/2 ≃ mχ/2. Generally, models lie
in one of these two regimes. To emphasize the generality of
our constraints, now we discuss particular scenarios.
For example, WIMPs with weak-scale masses and cross
sections may have monochromatic decays. The decay process
between the two lightest Kaluza-Klein (KK) particles, the KK
hypercharge gauge boson, B1 and KK graviton, G1 in mUED
models, and the decay between the two lightest particles in
SUSY theories, the Bino-like neutralino, B and gravitino, G
are well-studied examples. The mass scales of these candi-
dates are ∼800 GeV for the former and ∼80 GeV for the lat-
ter. The decay rates in these theories [9] are highly suppressed
due to the weakness of gravity and is given by
τ ≃
4.7 × 1022 s
b
MeV(cid:19)−3
(cid:18) ∆m
.
(11)
where the parameter b is identified as (2, 10/3, 1, 2) for each of
the decay reactions (G1 → B1+γ, B1 → G1+γ, G → B+γ,
B → G + γ). The lifetime requirement of τ ≫ τ0 translates
into ∆m < 30 MeV. Since ∆m ≪ mχ, the energy carried
away by the emitted photon is ε ≃ ∆m. Eq. (11) can be
rearranged as
mχ τ ≃ 4.7 × 1022 s(cid:16) mχ
b (cid:17)(cid:16) ε
MeV(cid:17)−3
,
(12)
which relates mχτ to ε in terms of a single parameter: mχ/b.
We plot mχτ versus ε in Fig. 4 for mχ/b ≃ 300 GeV to repre-
sent mUED. One sees that the Milky Way constraint requires
ε ≤ 1.5 MeV, which is a very strict limit as the lifetime is
proportional to ε−3, i.e., the decay rate increases by almost an
order of magnitude from 1 MeV to 2 MeV. This translates
to the restriction of ∆m . 1.5 MeV, which is far stricter
than the necessary condition to have a long-lived candidate,
∆m < 30 MeV.
In the R-parity violating supersymmetric extension of the
standard model, the lightest supersymmetric particle is again a
gravitino that might not be stable on cosmological time scales
against decay into a photon and neutrino ( G → ν +γ) through
a small photino-neutrino mixing Uγν. The lifetime of the
gravitino in this model [15] is
τ ≃ 3.8 × 1027 s(cid:18) Uγν
10−8(cid:19)−2
(cid:16) mχ
10 GeV(cid:17)−3
,
(13)
with the resulting photon and neutrino each carrying an energy
of ε = mχ/2. We can rewrite this equation in terms of mχτ
versus ε as
mχτ ≃ 1014 GeV s (Uγν)−2(cid:16) ε
GeV(cid:17)−2
.
(14)
We plot this relation for Uγν = 10−8 in Fig. 1, which
shows that the contribution of this decay model to the iDPB
m
U
E
D
S
U
S
Y
Excluded by Milky Way
γ-ray Line Search
1026
1025
1024
]
s
V
e
G
[
τ
m
χ
1023
0.1
the Isotropic Diffuse Photon Background
Over 10 % Contribution to
Overproduction of
the Isotropic Diffuse
Photon Background
1
ε [MeV]
10
FIG. 4: Similar to Fig. 1, focusing upon the MeV range of ε. The
contribution of late dark matter decays to the isotropic diffuse photon
background is 10% or more in the diagonal band. SUSY and mUED
inspired decaying dark matter models of Ref. [9] cannot make signif-
icant contribution to the iDPB while abiding by the gamma-ray line
emission limits from the Galactic Center region.
will be significant around ε ∼ 5 GeV (corresponding to
mχ ∼ 10 GeV) agreeing with Ref. [15]. Slightly above/below
mχ ∼ 10 GeV, either its contribution is negligible or vastly
overproduces the iDPB.
Dark matter models involving keV-mass sterile neutrinos,
in their simplest description, require only two parameters, the
sterile neutrino's mass and mixing with active neutrinos. The
decay chain for sterile neutrinos is νs → νe,µ,τ + γ, with a
radiative lifetime [28] (for Dirac neutrinos) of
τ = 1.5 × 1022 s(cid:0)sin−2 2θ(cid:1)(cid:16) ms
keV(cid:17)−5
.
(15)
This can similarly be rearranged, keeping in mind that the en-
ergy of the parent sterile neutrino is split equally between the
photon and the daughter neutrino (ε = ms/2), as
msτ ≃ 1015 GeV s (cid:0)sin−2 2θ(cid:1) (cid:16) ε
keV(cid:17)−4
,
(16)
which has only one free parameter, sin2 2θ. For illustration,
we plot Eq. (16) in Fig. 1 for sin2 2θ = 10−18. As seen in the
figure, and has been established in Ref. [5, 29], the gamma-
ray line emission limit from the Galactic Center region pro-
vides quite stringent restrictions on sterile neutrino dark mat-
ter, which can be several orders of magnitude stronger than
constraints from overproduction of the iDPB. Interestingly, all
three models we have discussed have the form mχτ ∝ ε−α,
where α = 3, 2, 4 respectively, as can also be noticed from the
varying slopes of the lines representing the models in Fig. 1.
V. CONCLUSIONS
5
Predictions of photon fluxes from dark matter decays are
considerably more robust than those from annihilations, due
to a lesser dependence upon theoretical uncertainties in the
distribution and clustering of dark matter. We have shown
that the gamma-ray line emission limits from the Galactic
Center region, along the isotropic diffuse photon background,
allow for stringent constraints to be placed on late decaying
dark matter scenarios that produce monoenergetic photons.
We emphasize that the Galactic and cosmic constraints are
not independent of each other, with the GC region provid-
ing stronger limits in its range of applicability due to new
spectroscopic data. Rather than attempting to explain vari-
ous gamma-ray phenomena with a specific model, we report
model-independent bounds on the decaying dark matter pa-
rameter space (as defined by mχτ versus ε). Our general
constraints are applicable to a number of models, and can
be used as a guide for future model building. Upcoming
gamma-ray telescopes with improved energy resolution, such
as GLAST [30] or ACT [31], can improve upon these bounds,
particularly by making use of the unique spectral shape and
directionality of decays from the Galactic halo.
One interesting application of our study is to assess the
recent suggestion that cosmological late dark matter decays
can explain the isotropic diffuse photon background in the
1-5 MeV range, whose origin remains a mystery [9]. We
plot mχτ versus ε in Fig. 4 for mχ/b ≃ 300 GeV and
mχ/b ≃ 50 GeV to represent the aforementioned mUED and
SUSY models of Ref. [9], respectively. We also show the
range of parameters, mχτ versus ε, that can lead to a substan-
tial (> 10%) contribution to the iDPB (shaded diagonal band)
or overproduce them (triangular region) through the sum of
the local decays (Galactic halo) or decays from truly cosmo-
logical sources (all distant dark matter halos). The region ex-
cluded by the gamma-ray line emission limits from the GC
region is below the jagged line. As seen in the figure, even the
combined emission from the Galactic halo and cosmological
sources due to either the mUED or SUSY inspired decaying
dark matter models of Ref. [9] cannot make a significant con-
tribution to the iDPB while abiding by the gamma-ray line
emission limits from the GC region. The mUED model can
contribute to the iDPB only for ∆m . 1.5 MeV with a contri-
bution of . 5%, while the SUSY model is even more severly
constrained. Even relaxing our assumptions on the distribu-
tion of dark matter in the halo does not increase these frac-
tions dramatically, thus, dark matter cannot decay in the late
universe at a high enough rate to make a prominent contribu-
tion to the iDPB in the MeV range.
We thank Shmuel Nussinov, Gary Steigman, Louie Stri-
gari, Akın Wingerter and especially John Beacom for fruit-
ful discussions. HY is supported by National Science Foun-
dation under CAREER Grant PHY-0547102 to JB; MDK
is supported by the Department of Energy grant DE-FG02-
91ER40690; both by CCAPP at OSU.
[1] F. Zwicky, Helv. Phys. Acta 6, 110 (1933).
[2] J. E. Gunn, B. W. Lee, I. Lerche, D. N. Schramm and
G. Steigman, Astrophys. J. 223, 1015 (1978); K. A. Olive
and J. Silk, Phys. Rev. Lett. 55, 2362 (1985); R. Barbieri
and V. Berezinsky, Phys. Lett. B 205, 559 (1988); M. T. Res-
sell and M. S. Turner, Comments Astrophys. 14, 323 (1990);
M. Kamionkowski,
[arXiv:astro-ph/9404079]; P. Nath and
R. Arnowitt, Phys. Lett. B 336, 395 (1994); G. D. Kribs and
I. Z. Rothstein, Phys. Rev. D 55, 4435 (1997) [Erratum-ibid.
D 56, 1822 (1997)]; R. A. Daly, Astrophys. J. L. 324, L47
(1988); A. Masiero, D. Montanino and M. Peloso, Astropart.
Phys. 12, 351 (2000); L. J. Hall and D. R. Smith, Phys. Rev.
D 60, 085008 (1999); H. B. Kim and J. E. Kim, Phys. Lett.
B 527, 18 (2002); K. Sigurdson and M. Kamionkowski, Phys.
Rev. Lett. 92, 171302 (2004); F. J. Sanchez-Salcedo, Astrophys.
J. 591, L107 (2003); S. H. Hansen and Z. Haiman, Astrophys. J.
600, 26 (2004); X. L. Chen and M. Kamionkowski, Phys. Rev.
D 70, 043502 (2004); K. Ichiki, M. Oguri and K. Takahashi,
Phys. Rev. Lett. 93, 071302 (2004); M. Kaplinghat, Phys. Rev.
D 72, 063510 (2005); K. Jedamzik, M. Lemoine and G. Moul-
taka, JCAP 0607, 010 (2006); L. E. Strigari, M. Kaplinghat and
J. S. Bullock, Phys. Rev. D 75, 061303 (2007); M. Mapelli,
A. Ferrara and E. Pierpaoli, Mon. Not. Roy. Astron. Soc. 369,
1719 (2006); L. Zhang, X. Chen, M. Kamionkowski, Z. g. Si
and Z. Zheng, Phys. Rev. D 76, 061301 (2007); D. Cum-
berbatch, K. Ichikawa, M. Kawasaki, K. Kohri, J. Silk and
G. D. Starkman, arXiv:0708.0095 [astro-ph]; K. Jedamzik,
arXiv:0710.5153 [hep-ph]; D. P. Finkbeiner and N. Weiner,
Phys. Rev. D 76, 083519 (2007); M. Pospelov and A. Ritz,
Phys. Lett. B 651, 208 (2007); L. M. Krauss and R. J. Scherrer,
Phys. Rev. D 75, 083524 (2007); J. P. Conlon and F. Quevedo,
JCAP 0708, 019 (2007).
[3] S. Dodelson and L. M. Widrow, Phys. Rev. Lett. 72, 17
(1994); K. Abazajian, G. M. Fuller and M. Patel, Phys. Rev.
D 64, 023501 (2001); A. D. Dolgov and S. H. Hansen, As-
tropart. Phys. 16, 339 (2002); T. Asaka, M. Shaposhnikov
and A. Kusenko, Phys. Lett. B 638, 401 (2006); A. Boyarsky,
A. Neronov, O. Ruchayskiy and M. Shaposhnikov, Mon. Not.
Roy. Astron. Soc. 370, 213 (2006).
[4] C. R. Watson, J. F. Beacom, H. Yuksel and T. P. Walker, Phys.
Rev. D 74, 033009 (2006);
[5] H. Yuksel, J. F. Beacom and C. R. Watson, arXiv:0706.4084
[astro-ph].
[6] T. Asaka, J. Hashiba, M. Kawasaki and T. Yanagida, Phys. Rev.
D 58, 023507 (1998); S. Kasuya and M. Kawasaki, Phys. Rev.
D 73, 063007 (2006).
[7] J. A. R. Cembranos, J. L. Feng and L. E. Strigari, Phys. Rev. D
75, 036004 (2007).
[8] T. Appelquist, H. C. Cheng and B. A. Dobrescu, Phys. Rev.
D 64, 035002 (2001); H. C. Cheng, K. T. Matchev and
M. Schmaltz, Phys. Rev. D 66, 036005 (2002); H. Baer, C. Bal-
azs, A. Belyaev, J. K. Mizukoshi, X. Tata and Y. Wang, JHEP
0207, 050 (2002); J. L. Feng, A. Rajaraman and F. Takayama,
Phys. Rev. D 68, 085018 (2003); N. R. Shah and C. E. M. Wag-
ner, Phys. Rev. D 74, 104008 (2006); S. Matsumoto, J. Sato,
M. Senami and M. Yamanaka, Phys. Lett. B 647, 466 (2007);
M. Kakizaki, S. Matsumoto and M. Senami, Phys. Rev. D 74,
023504 (2006); J. R. Ellis, S. Heinemeyer, K. A. Olive and
G. Weiglein, JHEP 0605, 005 (2006); D. Hooper and S. Pro-
fumo, Phys. Rept. 453, 29 (2007).
[9] J. A. R. Cembranos, J. L. Feng and L. E. Strigari, Phys. Rev.
Lett. 99, 191301 (2007).
[10] K. Watanabe, D. H. Hartmann, M. D. Leising and L. S. The,
Astrophys. J. 516, 285 (1999); P. Ruiz-Lapuente, M. Casse
6
and E. Vangioni-Flam, Astrophys. J. 549, 483 (2001); K. Ahn,
E. Komatsu and P. Hoflich, Phys. Rev. D 71, 121301 (2005);
L. E. Strigari, J. F. Beacom, T. P. Walker and P. Zhang, JCAP
0504, 017 (2005).
[11] Y. Ueda, M. Akiyama, K. Ohta and T. Miyaji, Astrophys. J. 598,
886 (2003); Y. Inoue, T. Totani and Y. Ueda, arXiv:0709.3877
[astro-ph].
[12] K. Ahn and E. Komatsu, Phys. Rev. D 72 061301 (2005);
Y. Rasera, R. Teyssier, P. Sizun, B. Cordier, J. Paul, M. Casse
and P. Fayet, Phys. Rev. D 73, 103518 (2006); K. Lawson and
A. R. Zhitnitsky, arXiv:0704.3064 [astro-ph].
[13] A. Bouquet and P. Salati, Nucl. Phys. B 284, 557 (1987);
F. Takayama and M. Yamaguchi, Phys. Lett. B 485, 388 (2000);
G. Moreau and M. Chemtob, Phys. Rev. D 65, 024033 (2002);
S. Lola, P. Osland and A. R. Raklev, arXiv:0707.2510 [hep-
ph]; W. Buchmuller, L. Covi, K. Hamaguchi, A. Ibarra and
T. Yanagida, JHEP 0703, 037 (2007).
[14] A. Ibarra and D. Tran, arXiv:0709.4593.
[15] G. Bertone, W. Buchmuller, L. Covi and A.
Ibarra,
arXiv:0709.2299 [astro-ph].
[16] B. J. Teegarden and K. Watanabe, Astrophys. J. 646, 965
(2006).
[17] E. Churazov et al., Astron. Astrophys. 467, 529, (2007).
[18] G. Weidenspointner et al., AIP Conf. Proc. 510, 467 (2000).
[19] D. E. Gruber, J. L. Matteson, L. E. Peterson and G. V. Jung,
Astrophys. J. 520, 124 (1999).
[20] P. Sreekumar et al. [EGRET Collaboration], Astrophys. J. 494,
523 (1998).
613, 956 (2004).
[21] A. W. Strong, I. V. Moskalenko and O. Reimer, Astrophys. J.
[22] J. Knodlseder et al., Astron. Astrophys. 441, 513 (2005);
J. F. Beacom and H. Yuksel, Phys. Rev. Lett. 97, 071102
(2006).
[23] J. F. Navarro, C. S. Frenk and S. D. M. White, Astrophys. J.
462, 563 (1996); A. V. Kravtsov, A. A. Klypin, J. S. Bullock
and J. R. Primack, Astrophys. J. 502, 48 (1998); B. Moore,
T. Quinn, F. Governato, J. Stadel and G. Lake, Mon. Not. Roy.
Astron. Soc. 310, 1147 (1999); C. G. Austin, L. L. R. Williams,
E. I. Barnes, A. Babul and J. J. Dalcanton, Astrophys. J. 634,
756 (2005); W. Dehnen and D. McLaughlin, Mon. Not. Roy.
Astron. Soc. 363, 1057 (2005); S. H. Hansen and J. Stadel,
JCAP 0605, 014 (2006).
[24] A. W. Strong,
I. V. Moskalenko
and O. Reimer,
[astro-ph/0506359].
[25] K. Nandra and K. A. Pounds, Mon. Not. Roy. Astron. Soc. 268,
405 (1994); P. Magdziarz and A. A. Zdziarski, Mon. Not. Roy.
Astron. Soc. 273, 837 (1995); D. E. Gruber, J. L. Matteson,
L. E. Peterson and G. V. Jung, Astrophys. J. 520, 124, (1999).
[26] V. Pavlidou and B. D. Fields, Astrophys. J. 575, L5 (2002);
T. Narumoto and T. Totani, Astrophys. J. 643, 81 (2006).
[27] H. Yuksel and M. D. Kistler, Phys. Rev. D 75, 083004
(2007); H. Yuksel and J. F. Beacom, Phys. Rev. D 76, 083007
(2007); H. Yuksel, S. Horiuchi, J. F. Beacom and S. Ando,
arXiv:0707.0196 [astro-ph].
[28] P. B. Pal and L. Wolfenstein, Phys. Rev. D 25, 766 (1982);
V. D. Barger, R. J. N. Phillips and S. Sarkar, Phys. Lett. B 352,
365 (1995) [Erratum-ibid. B 356, 617 (1995)]
[29] A. Boyarsky, D. Malyshev, A. Neronov and O. Ruchayskiy,
arXiv:0710.4922 [astro-ph].
[30] N. Gehrels and P. Michelson, Astropart. Phys. 11, 277 (1999).
[31] S. E. Boggs
[Larger ACT Collaboration],
al.
et
[astro-ph/0608532].
|
astro-ph/0103339 | 1 | 0103 | 2001-03-21T14:31:47 | Massive zero-metal stars: Energy production and mixing | [
"astro-ph"
] | Time-dependent nuclear network calculations at constant temperature show that for zero-metal stars >= 20 Msun (i) beta-decay reactions and (ii) the 13N(p,gamma)14O reaction must be included. It is also shown that the nuclear timescale in these zero-metal stars is shorter than the mixing timescale and therefore the assumption of instantaneous mixing across convective regions is not fulfilled. We conclude that proper modeling of these processes may alter the evolution of massive zero-metal stars. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no.
(will be inserted by hand later)
1
0
0
2
r
a
M
1
2
1
v
9
3
3
3
0
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Massive zero-metal stars: Energy production and mixing
C. W. Straka and W. M. Tscharnuter
Institut fur Theoretische Astrophysik, Universitat Heidelberg, Tiergartenstrasse 15, 69121 Heidelberg, Germany
Received; accepted
Abstract. Time-dependent nuclear network calculations at constant temperature show that for zero-metal stars
& 20M⊙ (i) β-decay reactions and (ii) the 13N(p,γ)14O reaction must be included. It is also shown that the
nuclear timescale in these zero-metal stars is shorter than the mixing timescale and therefore the assumption
of instantaneous mixing across convective regions is not fulfilled. We conclude that proper modeling of these
processes may alter the evolution of massive zero-metal stars.
Key words. nuclear reactions -- convection -- stars: evolution
1. Introduction
The evolution of zero-metal stars in the mass range ∼ 20 --
100M⊙ , i.e., massive Population III stars (Pop-III for
short), has been studied since the pioneering work of Ezer
et al. (1971) who followed the evolution of stars in the
mass range 5 -- 100M⊙ from the pre-main sequence con-
traction phase until the exhaustion of hydrogen on the
main sequence. This opened the field for many studies
that were mainly concerned with later evolutionary stages
(Cary, 1974; Castellani, 1983; El Eid et al., 1983; Ober et
al., 1983; Klapp, 1983, 1984). Motivated by the still ongo-
ing debate about the initial mass function of Pop-III stars,
Marigo et al. (2001) present the most comprehensive study
of zero-metal evolutionary models (0.7 -- 100M⊙) starting
from the ZAMS until the AGB in the case of low- and
intermediate-mass stars, or to the onset of carbon burn-
ing in massive stars.
In earlier studies the authors were forced to make nu-
merous assumptions about equilibria between the chemi-
cal species involved. The most recent studies by Marigo et
al. (2001), however, incorporate nuclear networks without
making any assumptions about equilibria. Nevertheless,
even these recent state-of-the-art studies rely on some sim-
plifications that are certainly fulfilled during the evolution
of normal stars but are questionable in the case of massive
zero-metal stars.
It is the aim of this letter to show that the following
two assumptions widely used, namely
1. β-decay negligible against proton-capture
2. instantaneous mixing across convective regions
Send offprint requests to: C. W. Straka,
e-mail: [email protected]
are not met during the evolution of massive Pop-III stars
and that proper modeling may alter the evolution of these
objects.
2. Nuclear Network
It was first noted by Ezer (1961) that Pop-III stars more
massive than about 20M⊙ are supplied by a different
mode of energy generation. Due to the lack of the elements
C, N and O the CNO-cycle cannot generate the luminosity
needed to halt contraction. This raises the temperature to
the point where the 3α-reaction produces enough carbon
to initiate the CNO-cycle. As a result, the CNO-cycle op-
erates at a much higher temperature, typically at 108 K.
Since the energy generation in the CNO-cycle exhibits a
strong temperature dependence the central temperatures
of the more massive models up to 100M⊙ do not exceed
Tc = 1.3 108 K (see e.g., El Eid et al., 1983; Marigo et al.,
2001). At these temperatures, the CNO-cycle turn-over
rate is governed by the beta-decay half-lives (Ezer et al.,
1971).
This finding has one important consequence for mod-
eling stellar evolution: one has to include the elements 13N
and 15O (mono-cycle), 17F (bi-cycle) and 18F (tri-cycle)
explicitely in time-dependent network calculations. In ad-
dition, for temperatures exceeding 108 K the proton cap-
ture 13N(p,γ)14O (followed by 14O(β+,ν)14N) supersedes
the 13N(β+,ν)13C rate. Contrary to the findings of Klapp
(1983) our calculations show a non-negligible effect on the
energy generation for T > 108 K.
In order to estimate the described effects, we have per-
formed time-dependent network calculations at constant
temperature. Our reference calculation includes 15 chem-
ical species: 1H, 4He, 12C, 13C, 13N, 14N, 15N, 14O, 15O,
2
C. W. Straka and W. M. Tscharnuter: Massive zero-metal stars: Energy production and mixing
Fig. 1. Energy generation per ρX 2
H at constant temperature for a) T = 1.0 108K and b) T = 1.3 108K. The solid curve
is our reference calculation including all 15 chemical species. The dashed curve neglects the β-decay and the dotted-
dashed curve neglects the reaction 13N(p,γ)14O. Differences between the reference curve and the curve neglecting
13N(p,γ)14O are too small to show up in a).
16O, 17O, 18O, 17F, 18F and 19F with initial abundances
of XH = 0.77, XHe = 0.23 and Xother = 10−15, and the
following reaction chains:
-- 3α-process
-- CNO tri-cycle
-- 13N(p,γ)14O and 14O(β+,ν)14N .
Reaction rates are taken from NACRE 1 (see Angulo et al.,
1999). The calculations are performed utilizing the DAE
solver LIMEX that is maintained by Ehrig et al. (1998).
Our results can be viewed in Fig. 1 where we have
plotted the energy generation per ρX 2
H against the time in
years for the two cases T = 1.0 108 K (see Fig. 1a) and T =
1.3 108 K (see Fig. 1b). Neglecting the beta-decay half-lives
(dashed curves) has an impact on the energy generation
of ∼ 10% for 1.0 108 K and ∼ 65% for 1.3 108 K. Whereas
the proton capture on 13N alters the energy generation at
1.0 108 K only slightly (< 1%), it has a ∼ 30% effect at
1.3 108 K (dotted-dashed curve). It is worth noting that
the energy generation is only marginally (< 1%) effected
by both the second (named 17ON) and the third cycle
(named 19FO). These cycles need only be included if one
is interested in the relative abundances of the elements O
and F. At constant temperature, the energy production
never attains a constant value due to the ongoing feeding
of freshly produced carbon via the 3α-process.
3. Mixing
Massive stars contain large convective cores in which mix-
ing of chemical species occurs due to the turbulent con-
vective motion. In normal stars, this process is very rapid
compared to the slow changes of the chemical composition
produced by nuclear reactions. Under these circumstances
1 http://pntpm.ulb.ac.be/Nacre/
one can safely assume that the composition in a convec-
tive region always remains homogeneous, i.e., elements are
instantaneously mixed over the convective region. The fol-
lowing arguments are put forward to show that the turn-
over time of the CNO-cycle at the high temperatures of
massive Pop-III stars can be comparable to the timescale
of turbulent mixing -- or even shorter.
3.1. Mixing timescale
We show that the mixing timescale, τconv, is of the or-
der of 10 days in the case of both normal and zero-metal
stars. An estimate of this timescale is easily derived from
the standard mixing length theory (MLT, Bohm-Vitense,
1958) and the formulas we present resemble those given
in Kippenhahn (1990, chap. 7). Let us assume a standard
composition of XH = 0.77 and XHe = 0.23, i.e., µ = 0.58,
and opacity due to electron scattering κ = 0.2(1 + XH),
i.e., κ = 0.35 cm2 g−1. Even in the case of hot Pop-III
stars, neglecting radiation pressure does not change this
order of magnitude estimate and it suffices to assume a
monoatomic ideal gas: δ = 1, cP = 5R/2µ. The mix-
ing timescale is approximately given by the pressure scale
height HP divided by the velocity of convective motion
vconv. With this:
τconv ∼ 108(cid:18) r
R⊙(cid:19)2(cid:18) M⊙m (cid:19)r T
108 K(cid:18) 10−4
√∇ − ∇e(cid:19) s (1)
where r, T and m are the radius, temperature and mass,
respectively, at one particular locus in a star. The quantity
∇ − ∇e is the difference between the actual temperature
gradient, ∇, and ∇e which describes the variation of T in
a mass element during its motion. Introducing two dimen-
sionless variables:
U ∼ 2 10−11(cid:18) m
M⊙(cid:19)(cid:18) T
108 K(cid:19)3/2
×
C. W. Straka and W. M. Tscharnuter: Massive zero-metal stars: Energy production and mixing
3
,
ρ
(2)
(cid:19)2(cid:18) R⊙r (cid:19)2
×(cid:18) 100 g cm−3
W = ∇rad − ∇ad ,
(3)
and provided that U ≪ (∇−∇e)1/2 ≪ W (see below) the
cubic equation of MLT can be simplified yielding
p∇ − ∇e ∼(cid:18) 8
U W(cid:19)1/3
For an estimate of τconv at the center of a star the first
two equations can be even further reduced, since m =
4/3πρcr3:
(4)
9
.
τconv ∼ 106(cid:18) M⊙m (cid:19)1/3(cid:18) 100 g cm−3
ρc
and
(cid:19)2/3(cid:18) Tc
108 K(cid:19)1/2
(5)
Fig. 2. Lines of constant τnuc at XH = 0.77. In the region
above the dashed line the mixing time is slow against the
nuclear timescale.
×(cid:18) 7 10−4
√∇ − ∇e(cid:19) s
M⊙(cid:19)1/3(cid:18) 100 g cm−3
ρc
(6)
(cid:19)4/3(cid:18) Tc
108 K(cid:19)3/2
U ∼ 4 10−10(cid:18) m
For a typical example, consider a 20M⊙ Pop-III star.
There we have Tc ∼ 108K, ρc ∼ 100 g cm−3. Let us choose
a mass coordinate: m = 1M⊙, hence U ∼ 4 10−10 from
Eq. (6). For reasonable values of W , W ∼ 1 . . . 100 it fol-
lows that (∇−∇e)1/2 ∼ 7 10−4. Note that the assumption
U ≪ (∇−∇e)1/2 ≪ W is fulfilled. With Eq. (5) we finally
arrive at
(7)
τconv ∼ 10 days .
For a star containing metals the central temperature is
by a factor of three smaller but as normal stars are less
compact also the density goes down by a factor of ten.
Thus τconv changes by a factor of two or so. Similar results
are obtained at other loci of the convective core and for
stellar masses in the range 20 -- 100 M⊙.
3.2. Nuclear timescale
A characteristic timescale for nuclear reactions in massive
stars is the turn-over time of the CNO-cycle. This time is
primarily given by the slowest reaction of the cycle which
is 14N(p,γ)15O. Even in the high temperature regime (T =
1.3 108K) where proton-captures become comparable to β-
decays this estimate holds within a factor of two. Hence,
AH
,
(8)
XHρλn14pg(T )
τnuc ∼ τn14pg =
where AH, XH and ρ are the atomic mass of hydrogen,
the mass abundance of hydrogen and the density, respec-
tively (see Clayton, 1983). The reaction rate λn14pg(T )
measured in [cm3 Mol−1 s−1] is only a function of temper-
ature. Taking this reaction rate again from NACRE (see
Angulo et al., 1999):
λn14pg(T ) = 4.83 107 T −2/3
9
×
× exp −15.231 T −1/3
×(1 − 2.00 T9 + 3.41 T 2
9
0.8(cid:19)2!
−(cid:18) T9
9 − 2.43 T 3
9 )
exp(−3.010 T −1
9 )
exp(−9.530 T −1
9 ) ,
+ 2.36 103 T −3/2
+ 6.72 103 T 0.380
9
9
(9)
the nuclear timescale is easily calculated for a given set of
density ρ and temperature T . To explore the parameter
space we plot lines of constant time τnuc in a ρ−T diagram
(see Fig. 2). The dashed line marks the typical timescale
of convective mixing. Thus, the region above (below) this
line is characterized by slow (fast) mixing compared to
the nuclear timescale. The assumption of instantaneous
mixing is only justified in the lower region where mixing
is fast. It can be nicely seen that all normal stars lie in the
lower region but that massive Pop-III stars lie in the upper
region where the assumption of instantaneous mixing does
not hold.
4. Conclusions
Performing time-dependent nuclear network calculations
at constant temperature we show that in the case of
massive Pop-III stars (& 20 M⊙) neglecting the β-decay
against proton-capture leads to a considerable error in
the energy generation rate. In addition, the reaction
13N(p,γ)14O cannot be omitted for temperatures exceed-
ing 108 K.
Moreover, the nuclear timescale of massive Pop-III
stars can be very short (order of hours) compared to the
timescale of convective mixing which is of the order of 10
days. Therefore, instantaneous mixing which is well justi-
fied in normal stars may introduce large errors in evolu-
tionary calculations of zero-metal stars.
Contrary to normal stars massive Pop-III stars have
not forgotten their nuclear history since equilibrium
4
C. W. Straka and W. M. Tscharnuter: Massive zero-metal stars: Energy production and mixing
abundances between chemical species are never attained.
Hence, evolutionary calculation must start on the pre-
main sequence well before the onset of nuclear reactions.
Acknowledgements. We thank W. J. Duschl for helpful discus-
sions and for improving the manuscript. We are particularly
grateful to R. Ehrig for providing and supporting LIMEX.
Part of this work was supported by the German Deutsche For-
schungsgemeinschaft, DFG (SFB 439 Galaxies in the young
universe).
References
Angulo, C., et al. 1999, Nucl. Phys., A656, 3
Bohm-Vitense, E. 1958, Z. Astrophys., 46, 108
Cary, N. 1974, Ap&SS, 31, 3
Castellani, V., Chieffi, A. & Tornamb´e, A. 1983, ApJ, 272,
249
Clayton, D. D. 1983, Principles of Stellar Evolution
and Nucleosynthesis (The University of Chicago Press,
Chicago)
Ehrig, R. & Nowak, U. 1998, LIMEX Version 4.1A1
(Konrad-Zuse-Zentrum fur Informationstechnik, Ber-
lin)
El Eid, M. F., Fricke, K. J. & Ober, W. W. 1983, A&A,
119, 54
Ezer, D. 1961, ApJ, 133, 159
Ezer, D. & Cameron, A. G. W. 1971, Ap&SS, 14, 399
Kippenhahn, R. & Weigert, A. 1990, Stellar Structure and
Evolution (Springer-Verlag, Berlin Heidelberg)
Klapp, J. 1983, Ap&SS, 93, 313
Klapp, J. 1984, Ap&SS, 106, 215
Marigo, P., Girardi, L., Chiosi, C. & Wood, P. R. 2001,
A&A, in press (astro-ph/0102253)
Ober, W. W., El Eid, M. F. & Fricke, K. J. 1983, A&A,
119, 61
|
astro-ph/0506622 | 2 | 0506 | 2005-10-02T04:58:41 | The Ital-FLAMES survey of the Sagittarius dwarf Spheroidal galaxy. I. Chemical abundances of bright RGB stars | [
"astro-ph"
] | We present iron and $\alpha$ element (Mg, Ca, Ti) abundances for a sample of 15 Red Giant Branch stars belonging to the main body of the Sagittarius dwarf Spheroidal galaxy. Abundances have been obtained from spectra collected using the high resolution spectrograph FLAMES-UVES mounted at the VLT. Stars of our sample have a mean metallicity of [Fe/H]=-0.41$\pm$0.20 with a metal poor tail extending to [Fe/H]=-1.52. The $\alpha$ element abundance ratios are slightly subsolar for metallicities higher than [Fe/H]\gtsima-1, suggesting a slow star formation rate. The [$\alpha$/Fe] of stars having [Fe/H]$<$-1 are compatible to what observed in Milky Way stars of comparable metallicity. | astro-ph | astro-ph | (DOI: will be inserted by hand later)
The Ital-FLAMES survey of the Sagittarius dwarf Spheroidal
galaxy. I. Chemical abundances of bright RGB stars ⋆
L. Monaco1,2, M. Bellazzini2, P. Bonifacio1, F.R. Ferraro3, G. Marconi4, E. Pancino2, L. Sbordone5,6, and S.
Zaggia1
1 Istituto Nazionale di Astrofisica -- Osservatorio Astronomico di Trieste, Via Tiepolo 11, I-34131 Trieste, Italy
2 Istituto Nazionale di Astrofisica -- Osservatorio Astronomico di Bologna, Italy I-40127 Bologna, Italy
3 Universit`a di Bologna -- Dipartimento di Astronomia, I-40127 Bologna, Italy
4 European Southern Observatory, Casilla 19001, Santiago, Chile
5 Istituto Nazionale di Astrofisica -- Osservatorio Astronomico di Roma, Italy, Via Frascati 33, 00040 Monteporzio Catone,
Roma
6 Universit`a Tor Vergata, Roma
Received / Accepted
Abstract.
We present iron and α element (Mg, Ca, Ti) abundances for a sample of 15 Red Giant Branch stars belonging to the main body
of the Sagittarius dwarf Spheroidal galaxy. Abundances have been obtained from spectra collected using the high resolution
spectrograph FLAMES-UVES mounted at the VLT. Stars of our sample have a mean metallicity of [Fe/H]=-0.41±0.20 with a
metal poor tail extending to [Fe/H]=-1.52. The α element abundance ratios are slightly subsolar for metallicities higher than
[Fe/H]>
∼-1, suggesting a slow star formation rate. The [α/Fe] of stars having [Fe/H]<-1 are compatible to what observed in
Milky Way stars of comparable metallicity.
Key words. Stars: abundances -- Stars: atmospheres -- Galaxies: abundances -- Galaxies: evolution -- Galaxies: dwarf -- Galaxies:
individual: Sgr dSph
1. Introduction
The Local Group (LG) is a heterogeneous environment.
Galaxies in the LG show a variety of characteristics (e.g.
mass, morphology, gas content) and are evolving under dif-
ferent conditions (e.g. in isolation, on strong dynamical in-
teraction). Thererore, in principle, they could teach us about
galaxy evolution as much as globular clusters did concerning
stellar evolution. Chemical abundances and abundance ratios
are key ingredients to study the star formation histories of stel-
lar systems. The modern generation of spectrographs mounted
on 8-10 m class telescopes allows to investigate the chemical
composition and dynamics of bright stars in LG galaxies but
only a handful of stars have been studied so far (Tolstoy et al.
2004, 2003; Shetrone et al. 2003; Bonifacio et al. 2000, 2004;
Fulbright et al. 2004; Geisler et al. 2005; Shetrone et al. 1998,
2001).
The commonly accepted paradigm (White & Rees 1978)
predicts the formation of large galaxies from the hierarchi-
Send offprint requests to: L. Monaco
⋆ Based on observations obtained with FLAMES at VLT Kueyen
8.2m telescope in the program 71.B-0146.
Correspondence to: [email protected]
cal assembly of small fragments similar to the LG dwarf
spheroidals (dSphs). In this framework, the comparison be-
tween the chemical composition of the Milky Way (MW) and
LG dSph stars is a first local testbed for the hierarchical merg-
ing model. The chemical composition of LG stars turned out to
be remarkably different from that of MW stars of comparable
metallicities. In particular, LG stars show α element abundance
ratios systematically under-abundant with respect to MW stars
(see, for instance, Venn et al. 2004; Bonifacio et al. 2004). The
interpretation of this empirical evidence is controversial. Is the
hierarchical merging a minor process in the assembly of the
MW? Or were the fragments from which the MW formed at
early times different from the nowadays recognizable dSphs?
The chemical difference between MW and LG stars may re-
flect an environmental difference between dwarfs accreted at
early times (galaxies near the bottom of the pre-MW potential
well -- dense environment) and the surviving dwarfs (galaxies
far from the bottom of the pre-MW potential well -- loose en-
vironment, but see Robertson et al. 2005; Bullock & Johnston
2004).
The
Sagittarius
dSph
Ibata, Gilmore, & Irwin 1994)
experiencing strong and disruptive tidal
(hereafter
Sgr,
is a LG galaxy currently
interactions with
5
0
0
2
t
c
O
2
2
v
2
2
6
6
0
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Table 1. Coordinates and atmospheric parameters for the program stars
Star
α (J2000.0) δ (J2000.0)
V
I
(V − I)
vhelio(km/s)
Ta
eff
log g
ξ
[M/H]
2300127
2300196
2300215
2409744
3600230
3600262
3600302
3800199
3800204
3800318
3800319
4303773
4304445
4402285
4408968
3600073⋆
3700178⋆
3800336⋆
4207953⋆
2300168⋆
3600181⋆
3700055⋆
4207391⋆
18 55 46.703
18 55 30.778
18 55 19.146
18 54 55.854
18 53 45.818
18 53 22.340
18 53 45.209
18 55 13.453
18 55 5.7440
18 54 58.264
18 54 58.088
18 54 02.120
18 53 40.606
18 53 19.765
18 53 12.886
18 53 56.477
18 54 18.068
18 55 11.635
18 54 14.546
18 55 20.010
18 53 57.440
18 54 2.6250
18 54 24.849
3600127
18 53 22.441
−30 35 24.683
−30 28 19.635
−30 30 27.978
−30 32 43.106
−30 25 49.419
−30 23 47.172
−30 30 55.702
−30 26 42.249
−30 27 56.602
−30 28 20.165
−30 28 58.481
−30 36 21.665
−30 35 42.879
−30 37 40.099
−30 32 03.565
−30 27 20.337
−30 29 31.259
−30 28 00.544
−30 32 34.502
−30 26 45.824
−30 25 09.207
−30 26 48.807
−30 33 02.291
−30 23 59.172
16.09
16.26
16.52
16.34
16.43
16.63
16.65
15.35
15.81
16.20
16.18
15.97
16.17
16.48
16.67
15.53
16.34
16.21
15.94
16.25
16.09
15.57
15.93
14.08
14.54
14.76
14.54
14.74
14.88
14.86
13.90
14.25
14.42
14.80
14.23
14.51
14.97
15.07
13.59
14.46
14.29
14.11
13.68
13.90
13.45
13.79
15.94
14.42
2.01
1.72
1.76
1.80
1.69
1.75
1.79
1.45
1.56
1.78
1.38
1.74
1.66
1.51
1.60
1.94
1.88
1.92
1.83
2.57
2.19
2.12
2.14
1.52
+147.2±0.58
+148.0±0.94
+154.9±0.79
+131.7±0.80
+153.8±0.56
+156.0±0.61
+143.8±0.67
+138.8±0.63
+153.2±0.61
+151.8±0.84
+141.1±0.88
+143.1±0.63
+119.9±0.59
+159.2±0.61
+144.0±0.68
+156.2±0.67
+149.1±0.81
+131.1±0.74
+129.0±0.89
+136.0±2.81
+147.3±1.23
+121.9±1.34
+133.7±2.01
-127.6±0.51
3687
3908
3877
3837
3947
3882
3848
4245
4101
3856
4364
3895
3976
4156
4047
3731
3770
3741
3815
3599
3610
3634
3624
4152
0.72
0.97
1.08
0.97
1.08
1.16
1.15
0.72
0.93
0.90
1.23
0.80
0.95
1.20
1.26
0.41
0.92
0.83
0.74
2.0
2.3
1.9
1.8
1.6
1.9
1.6
1.9
2.4
1.9
1.9
1.9
1.7
1.5
2.0
2.0
2.3
2.0
1.9
−1.0
−0.5
−0.5
−0.5
−0.5
−0.5
−0.5
−1.0
−1.5
−0.5
−1.5
−0.5
−0.5
−0.5
−0.5
−0.5
−0.5
−0.5
−0.5
a We adopted a reddening of E(V − I) = 0.18
⋆ Star showing TiO molecular bands in the spectra
the MW (Ibata, Gilmore, & Irwin 1995;
Ibata et al. 1997;
Majewski et al. 2003). Therefore, it may provide clues on the
influence of dynamical interactions on the chemical evolution
of dwarf galaxies.
It is well-known that the complex stellar content of Sgr
(see Monaco et al. 2002; Monaco et al. 2003, 2005, and ref-
erences therein) is largely dominated by a population of old-
intermediate age stars (∼6 Gyr, see, e.g. Bellazzini et al. 1999;
Layden & Sarajedini 2000; Monaco et al. 2002). However,
some concerns have been raised on mean metallicity estimates
obtained for this population from spectroscopic and photomet-
ric works (see, e.g. Mateo et al. 1995; Bonifacio et al. 2000;
Cole 2001; Monaco et al. 2002; Bonifacio et al. 2004).
The paper is devoted to the assessment of the mean chem-
ical properties of the Sgr dominant population. We present Fe,
Mg, Ca and Ti abundances for a selected sample of stars be-
longing to this population. In a companion paper (Bonifacio et
al., in preparation) we deal with the issue of the Sgr metallicity
distribution.
[Fe/H] and [α/Fe] abundances as well as the trends in
the [Fe/H] vs [α/Fe] plane constrain the chemical evolution
which led to the formation of the Sgr dominant population
(Lanfranchi & Matteucci 2003). Moreover, mean [Fe/H] and
[α/Fe] values are key ingredients to derive reliable age esti-
mates from the color-magnitude diagrams.
The paper is organized as follows. In §2 we describe the
target selection and the obtained data. In §3 we describe the
procedures followed to fix the atmospheric parameters and the
chemical analysis. In §4 we compare the results obtained with
previous works and in §5 we discuss our findings.
2. Observations
2.1. Targetselection,DataandEquivalentWidths
As part of the guaranteed time awarded to the Ital-FLAMES
consortium, more than 400 stars were observed in the Sgr dSph
(Bonifacio et al. 2005; Zaggia et al. 2004) from May the 23th
to 27th, 2003, using the FLAMES facility mounted on the VLT
(Pasquini et al. 2000). Details on the observations are given in
Zaggia et al. (2004). FLAMES allows to observe 132 targets
in one shot using the intermediate-low resolution spectrograph
GIRAFFE, plus 8 additional targets using the red arm of the
Table 2. Mean chemical abundances for the program stars. The signal to noise ratio of the coadded spectra and the number of
lines used are also reported
Stara
Sun
2300127
2300196
2300215
2409744
3600230
3600262
3600302
3800199
3800204
3800318
3800319
4303773
4304445
4402285
4408968
3600073⋆
3700178⋆
3800336⋆
4207953⋆
S/N
@653nm
A(Fe)
n
A(Mg)
n A(Ca)
n A(Ti)
20
20
14
22
21
21
24
32
31
23
21
18
33
22
18
43
19
25
30
7.51
6.70±0.24
7.02±0.19
7.28±0.18
7.25±0.06
7.34±0.18
7.14±0.18
7.20±0.18
6.41±0.17
5.99±0.08
6.98±0.17
6.14±0.26
6.78±0.15
7.16±0.14
7.22±0.13
7.09±0.16
6.73±0.18
7.06±0.19
6.88±0.16
7.10±0.13
15
15
13
10
16
15
15
15
13
16
32
14
16
15
17
17
16
14
15
7.58
6.82 ±0.15
7.03 ±0.18
7.04 ±0.18
7.09 ±0.03
7.24 ±0.17
7.27 ±0.13
7.03 ±0.14
6.52 ±0.08
6.28 ±0.04
6.83 ±0.15
6.56 ±0.09
6.66 ±0.16
7.21 ±0.05
7.11 ±0.17
6.81 ±0.04
6.67 ±0.14
6.83 ±0.17
7.01 ±0.10
6.97 ±0.10
2
4
4
2
4
3
4
4
2
4
2
4
2
4
3
4
3
3
4
6.35
5.35 ±0.10
5.67 ±0.16
5.98 ±0.22
5.82 ±0.08
5.81 ±0.14
5.58 ±0.19
5.85 ±0.11
5.60 ±0.10
4.97 ±0.07
5.88 ±0.18
5.38 ±0.15
5.47 ±0.17
5.76 ±0.21
5.99 ±0.25
5.82 ±0.11
5.23 ±0.15
5.47 ±0.32
5.63 ±0.12
5.63 ±0.19
9
9
9
8
9
9
9
8
5
9
7
9
9
9
9
9
9
8
9
4.94
3.99±0.06
4.49±0.16
4.84±0.25
4.67±0.13
4.63±0.14
4.34±0.18
4.53±0.17
4.32±0.11
3.65±0.10
4.69±0.17
4.14±0.14
3.98±0.05
4.45±0.13
4.79±0.13
4.56±0.12
3.87±0.20
4.16±0.24
4.40±0.16
4.51±0.17
n
7
9
9
9
9
9
9
8
8
9
7
7
9
9
9
9
9
9
8
⋆ Star showing TiO molecular bands in the spectra
A(X)=log( X
H )+12.00
high resolution specrograph UVES. In this paper we present
the results obtained from the UVES spectra.
It is important to recall that a large number of Milky Way
foreground stars are present along the Sgr line of sight. In order
to optimize the Sgr star detection rate, the target selection for
the UVES fibres was performed using the infrared 2 MASS1
color magnitude diagram (CMD). In fact, in the infrared plane,
the upper Sgr red giant branch (RGB) stands out very clearly
from the contaminating MW field (see, e.g., Cole 2001). This
also allows a thorough sampling of the Sgr dominant popu-
lation. In Fig. 1 we plotted the 2 MASS (K; J-KS ) CMD for
a 1 square degree area centred on the globular cluster M 54.
The heavy continuous line is the selection box. Target stars are
plotted as large filled circles. A similar target selection already
proven to be very effective in detecting stars belonging to the
Sgr Stream (Majewski et al. 2004).
Target stars are marked as large symbols in the optical
CMD plotted in Fig. 2 (Monaco et al. 2002). In Table 1 we re-
port equatorial (J 2000.0) coordinates and V and I magnitudes
for the target stars.
The coordinates in the J2000.0 absolute astrometric system
for both UVES and GIRAFFE samples were obtained with a
procedure already described in other papers (see, for example,
1 See http://www.ipac.caltech.edu/2mass
Ferraro et al. 2001). The new astrometric Guide Star Catalogue
(GSC II) recently released and now available on the web2 was
used as reference. In order to derive an astrometric solution we
used a program specifically developed at Bologna Observatory
(Montegriffo et al., in preparation). As a result of the entire
procedure, rms residuals of ∼0.15 arcsec, both in RA and Dec,
were obtained. The quality of the astrometry was confirmed by
the successful centreing of the fibres.
We performed the analysis on the spectra reduced with the
UVES ESO-MIDAS3 pipeline. For each pointing, 7 fibres were
centred on the target stars while one fibre was used to measure
the sky spectrum. Different spectra of the same star were coad-
ded and the resulting signal to noise ratio (S/N) ranges from 14
to 43 at 653 nm (see Table 2). UVES spectra have a resolution
of R≃43000 and cover the range between 480 nm and 680 nm.
Equivalent widths (EW) were measured on the spectra us-
ing the standard IRAF4 task splot. The Fe, Ca, Mg and Ti line
2 See http://www-gsss.stsci.edu/gsc/gsc2/GSC2home.htm
3 ESO-MIDAS is
the European Southern
Observatory Munich Image Data Analysis System which is de-
veloped and maintained by the European Southern Observatory.
http://www.eso.org/projects/esomidas/
the acronym for
4 IRAF is distributed by the National Optical Astronomy
Observatories, which is operated by the association of Universities for
Table 3. Mean abundance ratios for the program star
Star
[Fe/H]
[Mg/Fe]
[Ca/Fe]
[Ti/Fe]
2300127
2300196
2300215
2409744
3600230
3600262
3600302
3800199
3800204
3800318
3800319
4303773
4304445
4402285
4408968
3600073⋆
3700178⋆
3800336⋆
4207953⋆
-0.81±0.24
-0.49±0.19
-0.23±0.18
-0.26±0.06
-0.17±0.18
-0.37±0.18
-0.31±0.18
-1.10±0.17
-1.52±0.08
-0.53±0.17
-1.37±0.26
-0.73±0.15
-0.35±0.14
-0.29±0.13
-0.42±0.16
-0.78±0.18
-0.45±0.19
-0.63±0.16
-0.41±0.13
+0.05
-0.06
-0.31
-0.23
-0.17
+0.06
-0.24
+0.04
+0.22
-0.22
+0.35
-0.19
-0.02
-0.18
-0.35
-0.13
-0.30
+0.06
-0.20
-0.19
-0.19
-0.14
-0.27
-0.37
-0.40
-0.19
+0.35
+0.14
+0.06
+0.40
-0.15
-0.24
-0.07
-0.11
-0.34
-0.43
-0.09
-0.31
-0.14
+0.04
+0.13
-0.01
-0.14
-0.23
-0.10
+0.48
+0.23
+0.28
+0.57
-0.23
-0.14
+0.14
+0.04
-0.29
-0.33
+0.09
-0.02
⋆ Star showing TiO molecular bands in the spectra
[X/Y]=log( X
Y )-log( X
Y )
⊙
Table 4. Errors in the abundances of star # 3800318 due to
uncertainties in the atmospheric parameters
∆A(Fe)
∆A(Mg)
∆A(Ca)
∆A(Ti)
∆ξ = ±0.2 kms−1
∆Teff = ±100 K
∆ log g = ±0.50
−0.08
+0.10
−0.04
+0.07
+0.16
−0.14
∓0.04
−0.02
+0.05
+0.08
−0.07
∓0.12
±0.10
−0.04
+0.01
−0.15
+0.17
+0.13
−0.12
±0.05
lists as well as the adopted atomic parameters and the mea-
sured EW are reported in Table A.1. A different iron line list
(see Table A.2) was adopted for star #3800319 due to the rel-
atively high temperature and gravity of this star in comparison
with the other stars in the sample. We analysed interactively the
spectral lines. For each line the fit has been visually inspected
and adjusted until reaching a satisfying solution.
Research in Astronomy, Inc., under contract with the National Science
Foundation.
Fig. 1. The K vs J-KS 2 MASS color -- magnitude diagram for
a one square degree region around the globular cluster M 54.
Target stars are plotted as large filled circles.
2.2. RadialvelocitiesandthecontaminatingMilkyWay
field
Radial velocities (see Table 1) were obtained by cross-
correlating the observed spectra with a rest frame labora-
tory line list using the recently released software DAOSPEC5
(Stetson and Pancino, in preparation). The final radial veloci-
ties and relative errors were computed using about 150 lines for
each star. Geocentric observed radial velocities were corrected
to heliocentric velocities using the IRAF task rvcorrect.
The DAOSPEC code has the capability to measure the line
EWs. In our case, however, we used DAOSPEC only to mea-
sure the radial velocities of the target stars while we used the
IRAF task splot to measure EWs for homogeneity with our pre-
vious works on Sgr stars (Bonifacio et al. 2000, 2004). As a
check, the radial velocities of a few stars have also been mea-
sured using the fxcor IRAF task for Fourier cross correlation.
The radial velocities obtained using DAOSPEC and fxcor are
identical, within the errors.
All but one (#3600127, vhelio=-127.6 km/s, open circle in
Fig. 2) of the 24 observed stars are indeed Sgr radial velocity
members lying within ∼2σ of the systemic velocity as mea-
sured by Ibata et al. (1997). In Fig. 3 we plotted the velocity
distribution of the 23 Sgr radial velocity members. The mean
velocity (<vr >=143.08±3.2 km s−1)6 and the velocity disper-
sion of the sample (σ=11.17 km s−1) are in good agreement
with the values derived by Ibata, Gilmore, & Irwin (1995) and
Ibata et al. (1997).
5 See http://cadcwww.hia.nrc.ca/stetson/daospec
6 The quoted 3.2 km s−1 error has been estimated employing a boot-
strap technique.
Fig. 2. I vs V-I color -- magnitude diagram for a one square de-
gree region around the globular cluster M 54. Target stars
are marked with large symbols. Stars showing TiO molecular
bands in the spectra are plotted as plus symbols. Star 3600217
has a radial velocity not compatible with the membership to Sgr
and is plotted as a large empty circle. Theoretical isochrones
from which the surface gravities for the programme stars were
obtained are also plotted as continuous lines.
The MW model of Robin et al. (2003,
hereafter R03)
predicts that in the M 54 line of sight 2% of stars have
vr >100 km s−1, if we consider only stars lying in the same
(V, V-K)7 selection box of the UVES sample. However, the
model predicts only a ∼4% of giant stars (log g<4) in the selec-
tion box and none of them with vr >100 km s−1. We checked
carefully the 24 stars in the sample and we are confident that all
of them are indeed red giant. Therefore, even if the R03 model
provides only an approximate description of the MW, there is
no reason to expect any MW star among the 23 Sgr radial ve-
locity members in the sample.
2.3. M-giantsshowingTiOmolecularbandsinthe
spectra
The coolest (i.e. the reddest) four stars (#2300168, #3600181,
#3700055, #4207391, plus symbols at V-I>2.0 in Fig. 2) have
effective temperatures around 3600◦ K and very strong titanium
oxide bands (TiO, see Selvelli & Bonifacio 2000; Valenti et al.
1998) in the spectra (see Fig. 4). The presence of the TiO
bands confirm these stars as M-giants. Such strong molec-
ular bands prevent from a safe derivation of the equivalent
widths. Therefore, we do not present the chemical analysis for
Fig. 3. Heliocentric radial velocity distribution for the 23 pro-
gramme stars Sgr radial velocity members.
these stars. In addition, stars #3600073, #3700178, #3800366,
#4207953 (plus symbols at V-I<2.0 in Fig. 2) show weak but
clearly recognizable TiO bands. For these stars we provide only
a tentative analysis and the derived abundances will not be dis-
cussed. We plan to provide a detailed chemical analysis for
these 8 stars by performing spectral synthesis including also
the TiO molecular bands.
In the CMD in Fig. 2 star #2300127 lies exactly in the re-
gion occupied by stars with TiO bands in their spectra. Yet this
star does not present any band. The lack of the TiO molecular
bands may be due to the relatively weak Fe and Ti content of
this star ([Fe/H]=-0.81, [Ti/Fe]=-0.17, see Table 3).
3. Atmospheric Parameters and Chemical analysis
The UVES spectra of the 19 stars for which the chemical
analysis was performed (including also stars having weak TiO
bands) are plotted in Fig. 5.
3.1. Effectivetemperaturesandsurfacegravities
The effective temperatures for the target stars (see Table 1)
were derived from the (V-I) color assuming a reddening of E(V-
I)=0.18 (Layden & Sarajedini 2000) and using the calibration
of Alonso, Arribas, & Mart´ınez-Roger (1999).
We
used
the Girardi et al.
theoretical
isochrones, along with E(V-I)=0.18 and (m-M)0=17.10
(Layden & Sarajedini 2000; Monaco et al. 2004) as redden-
ing8 and distance modulus, in order to estimate the gravity
(2002)
7 The R03 model does not provide the (J-K) color, therefore we
define as selection box in the (V, V-K) plane the region which encloses
all the target stars.
8 We assumed the same reddening for all the stars in the sample.
Inspection of the Schlegel et al. (1998) reddening maps provide strong
indications that there is no serious variability of extinction in the con-
Fig. 4. Coadded UVES spectra of the 4 Sgr M-giants having strong TiO bands.
of the program stars. In particular, we used a (Z=0.001;
Age =14.13 Gyr) isochrone for stars #3800199 #3800204
#3800319 and a (Z=0.008; Age =6.31 Gyr) isochrone for
all the other stars (continuous lines in Fig. 2). These two
isochrones fit into the range covered by the target stars on the
CMD and the age and metallicity used are also compatible to
what expected from previous works (see Monaco et al. 2002;
Layden & Sarajedini 2000; Brown, Wallerstein, & Gonzalez
1999; Bonifacio et al. 2004).
sidered field (standard deviation of the reddening value: σE(B−V)=0.03,
Monaco et al. 2004).
3.2. ModelatmosphereandMicroturbulentvelocities
For each star we computed a plane parallel model atmosphere
using version 9 of the ATLAS code (Kurucz 1993) with the
above atmospheric parameters. Abundances were derived from
EWs using the WIDTH code (Kurucz 1993).
Microturbulent velocities (ξ) were determined minimizing
the dependence of the iron abundance from the EW, among the
set of iron lines measured for each star.
In Fig. 6 we plotted ξ as a function of the adopted gravity
for the stars studied by Ivans et al. (2001, bottom panel, here-
after I01), Shetrone et al. (2003, middle panel, hereafter S03)
Fig. 5. Coadded UVES spectra of the 19 Sgr giants analyzed in this paper. Labels on the right denote the star number, those on
the left the [Fe/H].
and for stars in our sample (top panel). A clear trend is present
in both the I01 and S03 samples. The same trend is present also
in our sample, albeit with a larger scatter. Continuous lines are
least square fits to the data points. In the case of our sample the
fit was obtained excluding the points having the highest and
lowest ξ (2.7 and 0.6 km s−1) and the point with the lowest
surface gravity (log g=0.41). As can be seen, the three fitting
lines are very similar to each other. A weak dependence of the
ξ from the effective temperature was found and it can be safely
neglected as a first order approximation. For stars #2300127,
#2300215 and #3800319 (filled circles in fig. 6) the ξ is 2.7,
2.5 and 0.6 km s−1, respectively, i.e. more than 2-σ far from the
fitting relation. When working with low S/N, highly crowded
spectra, it is difficult to measure weak Fe lines accurately. This
may lead to incorrect ξ. Thus, for stars #2300127, #2300215
and #3800319 we adopted the value obtained from the fitting
relation ξ=-0.35×log g+2.29, i.e. ξ=2.0, 1.9 and 1.9 km s−1,
respectively.
Fig. 6. Microturbulent velocities as a function of the surface
gravity for the programme stars (top panel) and stars in the S03
and I01 samples (middle and bottom panel respectively). Least
square fits to the data are plotted as continuous lines. The filled
circles in the upper panel mark stars more than 2-σ far from
the fitting relation.
3.3. ChemicalAbundances
The atmospheric parameters (Te f f , log g, ξ and the assumed
global metallicity [M/H]) adopted for the program stars are re-
ported in Table 1. The chemical abundances obtained for each
line are reported in Tables A.1 and A.2. The mean and standard
deviation of such abundances are reported in Tables 2 and 3 (as
[X/H] abundances in the latter case) for each chemical species.
In Table 2 we also reported the number of lines used to obtain
the mean abundance for each species. The line scatter reported
in Table 2 should be representative of the statistical error aris-
ing from the noise in the spectra and from uncertainties in the
measurement of the equivalent widths.
Under the assumption that each line provides an indepen-
dent measure of the abundance, the error in the mean abun-
dances should be obtained by dividing the line scatter by √n
(where n is the number of measured lines) and by adding to
this figure the errors arising from the uncertainties in the atmo-
spheric parameters. In Table 4 we report these latter errors in
the case of star #3800318, taken as representative of the whole
sample.
In Fig. 7 we plotted the metallicity distribution ob-
tained. Our sample spans a rather large metallicity range (-
1.52≤[Fe/H]≤-0.17). The distribution peaks around [Fe/H]≃-
0.4 and presents an extended metal poor tail9. In particular,
considering only stars more metal rich than [Fe/H]≃-1, which
9 Preliminary results obtained from the GIRAFFE sample show that
such tail extends at least down to [Fe/H]<-2.5 (Zaggia et al. 2004;
Bonifacio et al. 2005)
Fig. 7. Metallicity distribution of the program stars.
should be representative of the Sgr dominant population, we
obtain a mean value of <[Fe/H]>=-0.41±0.20.
In Fig. 8 we plotted the [Ti/Fe], [Ca/Fe] and [Mg/Fe] ra-
tios (from top to bottom panel) for the program stars as a
function of the [Fe/H] abundance. The 5 M 54 stars stud-
ied by Brown, Wallerstein, & Gonzalez (1999, hereafter B99)
are plotted as large open stars. Assuming < [α/Fe] >=
[Mg/Fe]+[Ca/Fe]
, we also obtain a mean value of <[α/Fe]>=-
0.17±0.07 for the dominant population.
Following Salaris, Chieffi & Straniero (1993), these values
correspond to a global metallicity10 of [M/H]=-0.51, which is
in good agreement with the recent photometric estimate by
Monaco et al. (2002).
2
3.4. NotesonMetalpoorstars:#3800199,
#3800204,#3800319
The three most metal poor stars (#3800199 #3800204 and
#3800319) occupy in the optical CMD (see Fig. 2) positions
compatible with the M 54 RGB (which is roughly represented
by the bluer isochrone in the plot).
The most metal poor star (#3800204, [Fe/H]=-1.52) lies
very near to the M 54 center (∼1′) and its chemical abundances
(Fig. 8) are identical to those of the M 54 stars studied by B99.
Therefore, it seems quite likely that this star does indeed belong
to M 54.
Star #3800319 ([Fe/H]=-1.37) is also quite near (∼1′.4)
to the cluster center but its chemical composition is only
marginally compatible with M 54 and it will be considered
a Sgr field star. However, we note that Layden & Sarajedini
10 The
[M/H]=[Fe/H]+log(0.638×10[α/Fe]+0.362)
metallicity"
"global
is
defined
as:
Fig. 8. α element abundance ratios ([Ti/Fe], [Ca/Fe] and
[Mg/Fe] from top to bottom panel) as a function of the iron
abundance for the program stars. Large open stars mark the 5
M 54 stars studied by B99.
(2000) claimed a metallicity dispersion of ∼0.16 dex for M 54
from the width of the red giant branch.
Star #3800199 ([Fe/H]=-1.10) is placed at 3′.2 from the
cluster center (which corresponds to ∼7 half light radii,
Trager, King, & Djorgovski 1995) and is significantly more
metal rich than M 54 ([Fe/H]∼-1.55, B99). Therefore, we con-
sider star #3800199 part of the Sgr galaxy field.
4. Comparison with previous works
Beside the present work, chemical abundances have been pre-
sented for Sgr RGB stars by Bonifacio et al. (2000, 2004, 2 and
10 stars, respectively) and by Smecker -- Hane & McWilliam
(2002,
(2004,
hereafter B04) also considered the two stars studied in
Bonifacio et al. (2000) obtaining a final sample of 12 stars.
hereafter S02, 14 stars). Bonifacio et al.
In Fig. 9 we plotted the [α/Fe] as a function of the iron
abundance for the stars in the 3 samples. Stars in our sample are
plotted as filled circles, while stars in the B04 and S02 sample
are plotted as empty squares and empty triangles, respectively.
The 5 M 54 stars studied by B99 are marked as large open
stars. The α element abundance ratio is defined as [α/Fe] =
[Mg/Fe]+[Ca/Fe]
for stars in our sample and in the B04 and B99
[S i/Fe]+[Ca/Fe]+[T i/Fe]
2
samples, while it is defined as [α/Fe] =
for stars in the S02 sample11.
3
Stars in the S02 sample range from [Fe/H]≃-1.6 to
[Fe/H]≃0. In particular, 3 stars in their sample have [Fe/H]<-1
and 11 stars are in the range -0.7÷0.0. This latter sub-sample
11 S02 do not provide abundances for each species but only mean
values.
2
Fig. 9. α element abundance ratio -- defined as: [α/Fe] =
[Mg/Fe]+[Ca/Fe]
-- as a function of the iron abundance for the
program stars (filled circles). Large open stars mark the 5 M 54
stars studied by B99. Open squares and open triangles mark
stars in the B04 and S02 samples, respectively. For stars in the
S02 sample: [α/Fe] =
[S i/Fe]+[Ca/Fe]+[T i/Fe]
.
3
has a mean metallicity and α element abundance ratio of:
<[Fe/H]>=-0.36±0.19 and <[α/Fe]>=+0.01±0.04. However,
it is important to note that these values should not be considered
as representative of the dominant population, since their target
selection has been biased toward stars with metallicities within
0.5 dex of the solar value based on previously obtained approx-
imate metallicities (McWilliam & Smecker-Hane 2004).
The metallicity range of stars in the B04 sample, on the
other hand, is -0.83≤[Fe/H]<+0.09. Therefore, it extends to-
ward slightly higher metallicity with respect to the S02 sam-
ple, but it lacks of metal poor stars. The mean metallicity and
α element abundance ratio of the B04 sample are: <[Fe/H]>=-
0.23±0.26 and <[α/Fe]>=-0.20±0.06.
The mean iron abundance obtained in this paper ([Fe/H]=-
0.41) is similar to that of the S02 and B04 samples. The
0.18 dex difference between the B04 mean iron abundance and
our figure would be also a little bit lowered (by 0.06÷0.09 dex)
by taking into account the different assumption about the red-
dening (B04 adopted E(V-I)=0.22 from Marconi et al. 1998).
The different target selection criterion adopted by B04 may
also be responsible for the residual difference in the mean iron
abundance (∼0.1 dex), which is, nevertheless, well inside the
involved errors.
The <[α/Fe]> ratio obtained by B04 is very similar to our
value (< [α/Fe] >=
=-0.17). S02 evaluate the
[S i/Fe]+[Ca/Fe]+[T i/Fe]
.
α element abundance ratio as [α/Fe] =
, we obtain a < [α/Fe] >
Considering [α/Fe] =
fairly similar to the S02 figure. The small residual difference
[Mg/Fe]+[Ca/Fe]
[Ca/Fe]+[T i/Fe]
3
2
2
(∼0.1 dex higher in the S02 sample) may be partly ascribed
to the [Si/Fe] abundances and, possibly, to a different set of
lines and atomic parameters adopted in the chemical analysis.
Unfortunately, S02 neither provide abundances for each species
nor the atomic data and the adopted line list and this hypothesis
cannot be checked further.
Finally, as already stressed in section 3.4, we remark that
the Fe, Mg, Ca and Ti abundances of star #3800204 are consis-
tent with the results obtained by B99 for M 54 stars.
5. Discussion and Conclusions
The main purpose of this paper was to study the chemical com-
position of the dominant population of the Sgr dSph galaxy. We
selected 24 target stars using the 2 MASS infrared CMD, where
the upper RGB of Sgr is well separated from the MW field.
Target stars have been observed using the red arm of the high
resolution spectrograph FLAMES-UVES. We reported radial
velocities for these 24 stars and all but one are Sgr radial ve-
locity members. Eight stars show strong or visible TiO bands.
For stars with weak TiO bands we present a tentative chemi-
cal analysis while we do not present any chemical analysis for
stars presenting strong TiO bands in the spectra.
For the remaining 15 stars, we reported Fe, Mg, Ca and Ti
chemical abundances. This is the largest sample of high resolu-
tion spectra analyzed so far for stars in the Sgr dSph galaxy, and
the only sample thoroughly representative of the Sgr dominant
population.
The metallicity ranges from [Fe/H]=-1.52 to [Fe/H]=-0.17.
Three stars have [Fe/H]<-1 and the most metal poor of them
(#3800204) can be reasonably considered M 54 member.
The mean iron content of stars with [Fe/H]>-1 (i.e. the Sgr
dominant population) is <[Fe/H]>=-0.41±0.20, with a mean
α element abundance ratio <[α/Fe]>=-0.17±0.07. These fig-
ures lead to a global metallicity [M/H]=-0.51 which is in close
agreement with the most recent photometric estimates obtained
for the Sgr dominant population (Monaco et al. 2002).
In order to obtain a more statistically significant sample,
we now join the B04 and our samples. In Fig. 10 we plot-
ted in the [α/Fe] vs [Fe/H] plane the mean points obtained for
Sgr from this larger sample of Sgr stars as filled circles. For
−0.65 <[Fe/H]< 0.1, filled circles represent running means
with 0.20 dex as bin and 0.1 dex as step. For stars having
−1.0 <[Fe/H]< −0.65 and −1.5 <[Fe/H]< −1 (i.e. excluding
star #3800204 which has been tagged as M 54 member) filled
circles are straight means of the [Fe/H] and [α/Fe] with the
corresponding standard deviations as errorbars. A weak, but
clearly recognizable trend between the α element abundance
ratio and the mean iron abundance exists at high metallicity.
Such a trend waits to be confirmed from a much more extended
sample such as that obtained using the FLAMES-GIRAFFE
multifibre spectrograph which is currently under analysis. For
[Fe/H]<-1, a sudden increase of the [α/Fe] is apparent.
The mean [α/Fe] at low metallicities is consistent with
the values observed in MW stars (crosses in Fig. 10, from
Venn et al. 2004) of comparable metallicities and somewhat
higher with respect to stars in the LG galaxies (asterisks in the
figure, from Venn et al. 2004). Therefore, metal poor stars lost
[Mg/Fe]+[Ca/Fe]
2
Fig. 10. [α/Fe]=
as a function of [Fe/H] for stars
in the MW and in Local Group dwarf galaxies (crosses and
asterisks, respectively, from Venn et al. 2004). Filled circles are
mean points for Sgr obtained joining the B04 sample and our
data.
in early passages which now are not recognizable as Sgr tidal
debris (Helmi 2004), would be part of the typical content of
the MW Halo and impossible to tag as an accreted component
from the chemical composition.
The three metal poor stars in the S02 sample are compati-
ble with MW stars as well. This occurrence led the authors to
suggest that the upper mass end of the Sgr initial mass function
(IMF) should not be significantly different from the MW one.
The level of [α/Fe] which characterizes a galaxy at low metal-
licities may indeed give information on the IMF of the galaxy
at that time (see McWilliam 1997, and references therein),
since the amount of α elements and iron produced by a Type
II SN is a function of the mass of the SN progenitor. Although
this is true in principle, in practice this information may not be
presently extracted. In fact the ratio of the α elements and iron
produced by a Type II SN is also a sensitive function of the
"mass cut", i.e. the mass coordinate which separates the ma-
terial of the SN which "falls back" on the SN remnant from
the material which is ejected. The deeper the mass cut, the
more iron-peak elements are ejected, thus lowering the overall
[α/Fe]. Current SN models are unable to determine the mass
cut in a self consistent way or from first principles: the mass
cut is always assumed. We do not have either any indication on
whether the mass cut is in any sense "Universal" or if it may
vary e.g. depending on the mass of the star or on its metallicity.
With this state of affairs, any inference on the IMF from the
level of the [α/Fe] ratio of a galaxy would be highly uncertain.
The metal-rich Sgr stars lie on the extension to high metal-
licity of the pattern followed by stars in LG galaxies and be-
low MW stars. A low [α/Fe] at high metallicity is traditionally
interpreted as evidence for a slow or bursting star formation
rate (S02, B04, Marconi et al. 1994). On the contrary, in or-
der to reproduce the Sgr [α/Fe] ratios, Lanfranchi & Matteucci
(2003) required a high star formation rate. However, they con-
strained their model using preliminary abundances presented
by Smecker-Hane & Mc William (1999). The somewhat lower
[α/Fe] values obtained here and in S02 and B04 should be in
better agreement with a lower star formation rate.
Our data suggest that Sgr had a different chemical evolu-
tion from both the MW and the LG galaxies (see Fig. 10). A
different chemical evolution for Sgr with respect to the other
LG galaxies is expected, since Sgr experienced strong and dis-
ruptive dynamical interactions with the MW. Such interactions
are witnessed by the Sgr tidal tails studied by Majewski et al.
(2003) and are expected to trigger star formation activity
(see, for instance, Kravtsov et al. 2004; Mayer et al. 2001;
Zaritsky & Harris 2004).
Finally, we note that the Large Magellanic Cloud (LMC)
metallicity distribution strongly resambles the Sgr one. In fact,
Cole et al. (2005) approximated the metallicity distribution of
the LMC bar by two Gaussians having [Fe/H]=-0.37±0.15 and
[Fe/H]=-1.08±0.46 and containing 89% and 11% of the stars,
respectively. The same results hold also for the LMC disk (see
Cole et al. 2005). Clearly, Sgr has the same mean metallicity
of the LMC dominant population as well as the same fraction
of metal poor stars (see Monaco et al. 2003). Such occurrence
may suggest a similarity of the Sgr progenitor with the LMC.
Acknowledgements. Part of the data analysis has been performed us-
ing software developed by P. Montegriffo at the INAF - Osservatorio
Astronomico di Bologna. This research was done with support from
the Italian MIUR COFIN/PRIN grants 2002028935 and 2004025729.
We are grateful to L. Girardi for useful comments and to G. Schiulaz
for a careful reading of the manuscript.
References
Ferraro F. R., D'Amico N., Possenti A., Mignani R. P.,
Paltrinieri B., 2001, ApJ, 561, 337
Fulbright, J. P., Rich, R. M., & Castro, S. 2004, ApJ, 612, 447
1988.
Fuhr,
J.R., Martin, G.A.,
and Wiese, W.L.
J.Phys.Chem.Ref.Data 17, Suppl. 4. (FMW)
Geisler, D., Smith, V. V., Wallerstein, G., Gonzalez, G., &
Charbonnel, C. 2005, AJ, 129, 1428
Girardi, L., Bertelli, G., Bressan, A., Chiosi, C., Groenewegen,
M. A. T., Marigo, P., Salasnich, B., & Weiss, A. 2002, A&A,
391, 195
Gratton, R. G., Carretta, E., Claudi, R., Lucatello, S., &
Barbieri, M. 2003, A&A, 404, 187 (G03)
Helmi, A. 2004, MNRAS, 351, 643
Ibata, R. A., Gilmore, G., & Irwin, M. J. 1994, Nature, 370,
194
Ibata, R. A., Gilmore, G., & Irwin, M. J. 1995, MNRAS, 277,
781
Ibata, R.A., Wyse, R.F.G., Gilmore, G., Irwin, M.J., & Suntzeff,
N.B., 1997, AJ, 113, 634
Ivans, I. I., Kraft, R. P., Sneden, C., Smith, G. H., Rich, R. M.,
& Shetrone, M. 2001, AJ, 122, 1438
Kravtsov, A. V., Gnedin, O. Y., & Klypin, A. A. 2004, ApJ,
609, 482
Kurucz, R. L. 1993, CD-ROM 13, 18 http://kurucz.harvard.edu
Layden, A.C., & Sarajedini, A., 2000, AJ, 119, 1760 (LS00)
Lanfranchi, G. A., & Matteucci, F. 2003, MNRAS, 345, 71
Majewski S.R., Skrutskie M.F., Weinberg M.D., Ostheimer
J.C., 2003, ApJ, 599, 1082
Majewski, S. R., et al. 2004, AJ, 128, 245
Marconi, G., Matteucci, F., & Tosi, M. 1994, MNRAS, 270, 35
Marconi, G., Buonanno, R., Castellani, M., Iannicola, G.,
Molaro, P., Pasquini, L., & Pulone, L. 1998, A&A, 330, 453
1988.
Martin, G.A., Fuhr,
and Wiese, W.L.
J.R.,
J.Phys.Chem.Ref.Data 17, Suppl. 3. (MFW)
Mayer, L., Governato, F., Colpi, M., Moore, B., Quinn, T.,
Wadsley, J., Stadel, J., & Lake, G. 2001, ApJ, 559, 754
Mateo, M., Udalski, A., Szymansky, M., Kaluzny, J., Kubiak,
Alonso, A., Arribas, S., & Mart´ınez-Roger, C. 1999, A&AS,
M., & Krzeminski, W., 1995, AJ, 110, 1141
140, 261
Bellazzini, M., Ferraro, F.R., Buonanno, R., 1999b, MNRAS,
307, 619
Bonifacio P. et al., Proceedings of the ESO/Arcetri Conference
on Chemical Abundances and Mixing in Stars in the Milky
Way and its Satellites held in Castiglione della Pescaia, Italy,
13-17 September 2004
McWilliam, A. 1997, ARA&A, 35, 503
Mc William, A., & Smecker-Hane, T. 2004, to appear in ASP
conference on "Cosmic Abundances as Records of Stellar
Evolution and Nucleosynthesis", 2005; editors F.N. Bash
and T.G. Barnes (astro-ph/0409083)
Monaco L., Ferraro F.R., Bellazzini M., Pancino E. 2002, ApJ,
578, L47
Bonifacio P., Hill V., Molaro P., Pasquini L., Di Marcantonio
Monaco L., Bellazzini M., Ferraro F.R., Pancino E., 2003, ApJ,
P., Santin P. 2000, A&A, 359, 663
597, L25
Bonifacio, P., Sbordone, L., Marconi, G., Pasquini, L., & Hill,
Monaco L., Bellazzini, M., Ferraro, F.R., Pancino, E., 2004,
V. 2004, A&A, 414, 503
MNRAS, 353, 874
Brown, J. A., Wallerstein, G., & Gonzalez, G. 1999, AJ, 118,
Monaco, L., Bellazzini, M., Ferraro, F. R., & Pancino, E. 2005,
1245
Bullock J. S., & Johnston K. V. 2004, to appear in the proceed-
ings of the conference "Satellites and Tidal Streams", held in
La Palma, Spain, 26-30 May 2003, eds. Prada F., Martinez-
Delgado D., Mahoney T. (astro-ph/0401625)
Cole, A. A., Tolstoy, E., Gallagher, J. S., & Smecker-Hane,
T. A. 2005, AJ, 129, 1465
Cole, A. A. 2001, ApJ, 559, L17
MNRAS, 356, 1396
O'Brian, T. R., Wickliffe, M. E., Lawler, J. E., Whaling, J. W.,
& Brault, W. 1991, Optical Society of America Journal B
Optical Physics, 8, 1185 (O)
Pasquini, L., et al. 2000, Proc. SPIE, 4008, 129
Robertson B., Bullock J. S., Font A. S., Johnston K. V., &
Hernquist L. 2005, submitted to ApJ(astro-ph/0501398)
Table A.2. Iron line list and adopted atomic parameters for star
#3800319. The measured equivalent width and the correspond-
ing abundance obtained for each line are also reported
Ion
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
log gf
λ
(nm)
source of
log gf
(see notes)
EW
(pm)
3800319
487.1318
491.8994
492.0502
495.7298
495.7596
500.6119
501.2068
505.1635
511.0413
517.1596
519.1454
519.2344
519.4941
522.7189
523.2940
526.6555
527.0356
532.4179
532.8039
532.8531
537.1489
539.7128
540.5775
542.9696
543.4524
544.6916
545.5609
561.5644
595.2718
602.7051
616.5360
670.3566
-0.410
-0.370
0.060
-0.342
0.127
-0.615
-2.642
-2.795
-3.760
-1.793
-0.551
-0.421
-2.090
-1.228
-0.190
-0.490
-1.510
-0.240
-1.466
-1.850
-1.645
-1.993
-1.844
-1.879
-2.122
-1.930
-2.091
-0.140
-1.440
-1.210
-1.550
-3.160
FMW
FMW
FMW
WBW2
WBW2
FMW
FMW
FMW
FMW
FMW
O
O
FMW
O
FMW
FMW
FMW
FMW
FMW
O
FMW
FMW
FMW
FMW
FMW
FMW
O
FMW
FMW
FMW
FMW
FMW
17.32
17.84
30.56
17.69
28.12
22.38
23.42
22.54
18.78
20.64
18.57
14.89
18.75
23.52
19.94
19.96
25.96
17.49
43.04
22.24
29.90
26.15
26.33
32.89
22.59
28.62
36.09
19.44
5.11
6.22
5.95
4.69
ǫ
5.949
5.958
6.368
5.878
6.162
6.571
6.398
6.528
5.715
6.045
6.362
5.643
6.144
5.867
6.012
6.379
6.378
6.123
6.118
6.302
5.86
5.884
5.852
6.209
5.806
6.095
6.604
6.333
6.135
6.195
6.57
6.127
FMW -- Fuhr et al. (1988)
WBW2 -- Wolniket et al. (1971)
O -- O'Brian et al. (1991)
Robin, A. C., Reyl´e, C., Derri`ere, S., & Picaud, S. 2003, A&A,
409, 523
Salaris, M., Chieffi, A., Straniero O., 1993, ApJ, 414, 580
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500,
525
Selvelli, P. L., & Bonifacio, P. 2000, A&A, 364, L1
Shetrone, M. D., Bolte, M., & Stetson, P. B. 1998, AJ, 115,
1888
Shetrone, M. D., Cot´e, P., & Sargent, W. L. W. 2001, ApJ, 548,
592
Shetrone, M., Venn, K. A., Tolstoy, E., Primas, F., Hill, V., &
Kaufer, A. 2003, AJ, 125, 684
Smecker-Hane, T., & Mc William, A. 1999, Astronomical
Society of the Pacific Conference Series, 192, 150
Smecker -- Hane T.A., McWilliam A., 2002, ApJ submitted,
astro-ph/0205411
Smith, G., & Raggett, D. S. J. 1981, Journal of Physics B
Atomic Molecular Physics, 14, 4015 (SR)
Tolstoy, E., et al. 2004, ApJ, 617, L119
Tolstoy, E., Venn, K. A., Shetrone, M., Primas, F., Hill, V.,
Kaufer, A., & Szeifert, T. 2003, AJ, 125, 707
Trager, S. C., King, I. R., & Djorgovski, S. 1995, AJ, 109, 218
Valenti, J. A., Piskunov, N., & Johns-Krull, C. M. 1998, ApJ,
498, 851
Venn, K. A., Irwin, M., Shetrone, M. D., Tout, C. A., Hill, V.,
& Tolstoy, E. 2004, AJ, 128, 1177
White, S. D. M., & Rees, M. J. 1978, MNRAS, 183, 341
Wiese, W.L., Smith, M.W., and Miles, B.M. 1969, NSRDS-
NBS 22. (NBS)
Wolnik, S.J., Berthel, R.O., and Wares, G.W. 1971, ApJ, 166,
L31 (WBW2)
Zaggia, S., et al. 2004, Memorie della Societa Astronomica
Italiana Supplement, 5, 291
Zaritsky, D., & Harris, J. 2004, ApJ, 604, 167
Appendix A: Individual line data
The following tables report the line list and adopted atomic pa-
rameters for the program stars. The measured equivalent width
and the corresponding abundance obtained for each line are
also reported.
Table A.1. Line list and adopted atomic parameters for the program stars. The measured equivalent width and the corresponding
abundance obtained for each line are also reported
Ion
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Mg I
Mg I
Mg I
Mg I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
λ
(nm)
585.5076
588.3817
595.2718
602.4058
602.7051
605.6005
609.6664
615.1617
616.5360
618.7989
622.6734
651.8366
659.7559
670.3566
673.9521
674.6954
679.3258
552.8405
571.1088
631.8717
631.9237
585.7451
586.7562
612.2217
616.9042
643.9075
645.5558
649.3781
649.9650
650.8850
588.0269
590.3315
593.7809
595.3160
597.8541
601.6995
606.4626
609.1171
609.2792
log gf
source of
log gf
(see notes)
EW
(pm)
2300127
ǫ
EW
(pm)
2300196
ǫ
EW (pm)
(pm)
2300215
ǫ
EW
(pm)
2409744
ǫ
EW
(pm)
3600073⋆
-1.76
-1.36
-1.44
-0.12
-1.21
-0.46
-1.93
-3.30
-1.55
-1.72
-2.22
-2.75
-1.07
-3.16
-4.95
-4.35
-2.47
-0.522
-1.729
-1.945
-2.165
0.240
-1.490
-0.315
-0.797
0.390
-1.290
-0.109
-0.818
-2.110
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
G03
G03
G03
G03
SR
G03
SR
SR
SR
SR
SR
SR
NBS
-2.045 MFW
-2.145 MFW
-1.890 MFW
-0.329 MFW
-0.496 MFW
-3.630 MFW
-1.944 MFW
-0.423 MFW
-1.379 MFW
--
9.19
6.51
11.52
12.20
8.91
5.98
15.14
7.46
8.18
2.34
13.18
5.97
11.49
11.75
--
2.41
23.92
14.75
--
--
19.82
6.94
30.67
17.36
26.02
14.86
21.22
16.09
7.53
13.16
12.61
13.52
15.25
14.21
--
--
11.49
7.80
--
6.607
6.269
6.654
7.15
6.825
6.666
6.671
6.772
6.755
6.116
6.811
7.017
6.823
6.714
--
6.658
6.67
6.98
--
--
5.381
5.263
5.327
5.494
5.338
5.535
5.289
5.243
5.276
4.002
4.029
3.903
4.023
3.981
--
--
4.081
3.89
5.89
10.70
9.42
15.93
14.34
9.66
--
17.38
9.78
10.88
7.57
14.61
6.07
13.50
11.36
6.23
--
25.91
16.63
7.78
3.67
23.93
8.57
30.52
19.67
29.40
15.11
21.69
20.09
7.99
14.99
16.09
15.31
17.71
15.55
3.21
13.98
13.57
10.40
7.375
6.763
6.684
7.153
7.335
6.808
--
6.982
7.066
7.108
7.045
6.979
6.968
7.107
6.785
7.124
--
6.811
7.125
7.279
6.913
5.848
5.643
5.576
5.846
5.736
5.632
5.351
5.843
5.550
4.512
4.789
4.403
4.516
4.325
4.615
4.219
4.553
4.511
--
11.67
--
13.97
11.75
9.65
9.26
17.52
9.20
9.67
7.73
17.20
7.34
14.97
--
--
4.19
24.36
16.58
5.24
3.73
21.37
12.08
41.66
19.24
28.96
14.67
22.97
21.65
9.06
16.08
12.98
15.86
17.94
17.77
4.47
16.58
16.34
8.53
--
7.195
--
7.173
7.214
7.031
7.352
7.339
7.19
7.149
7.229
7.677
7.32
7.632
--
--
7.151
6.827
7.326
7.012
6.981
5.782
6.253
6.086
6.089
5.854
5.795
5.812
6.364
5.744
4.985
4.544
4.761
4.935
5.039
4.783
4.896
5.236
4.35
4.42
--
--
14.75
--
--
--
16.63
8.86
10.19
7.63
13.50
6.72
12.19
--
6.66
--
--
14.16
--
4.49
20.52
9.27
--
17.68
28.92
13.82
22.43
18.07
9.75
14.31
13.38
15.38
17.02
15.03
3.78
16.49
12.30
9.65
7.253
--
--
7.367
--
--
--
7.263
7.172
7.291
7.238
7.162
7.248
7.225
--
7.284
--
--
7.052
--
7.12
5.719
5.811
--
5.908
5.861
5.687
5.795
5.908
5.836
4.684
4.624
4.705
4.826
4.567
4.611
4.925
4.571
4.487
4.44
10.66
7.29
11.14
9.30
7.34
6.18
15.05
6.80
9.04
5.37
13.83
4.54
11.57
9.83
5.01
2.73
20.64
12.13
4.74
2.68
17.66
6.25
26.93
14.76
24.25
13.84
18.13
16.61
5.49
9.68
10.91
13.95
11.69
12.59
1.69
13.91
10.83
7.01
ǫ
7.149
6.861
6.398
6.569
6.637
6.535
6.702
6.701
6.657
6.895
6.706
6.937
6.747
6.856
6.486
6.843
6.742
6.485
6.612
6.868
6.709
5.192
5.231
5.298
5.151
5.322
5.439
4.933
5.4
5.082
3.591
3.884
4.088
3.516
3.8
3.956
4.08
4.063
3.89
⋆ Star showing TiO molecular bands in the spectra
FMW -- Fuhr et al. (1988)
G03 -- Gratton et al. (2003)
SR -- Smith et al. (1981)
NBS -- Wiese et al. (1969)
MFW -- Martin et al. (1988)
Table A.1. Line list and adopted atomic parameters for the program stars. The measured equivalent width and the corresponding
abundance obtained for each line are also reported (continued)
Ion
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Mg I
Mg I
Mg I
Mg I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
λ
(nm)
585.5076
588.3817
595.2718
602.4058
602.7051
605.6005
609.6664
615.1617
616.5360
618.7989
622.6734
651.8366
659.7559
670.3566
673.9521
674.6954
679.3258
552.8405
571.1088
631.8717
631.9237
585.7451
586.7562
612.2217
616.9042
643.9075
645.5558
649.3781
649.9650
650.8850
588.0269
590.3315
593.7809
595.3160
597.8541
601.6995
606.4626
609.1171
609.2792
log gf
source of
log gf
(see notes)
EW
(pm)
3600230
ǫ
EW
(pm)
3600262
ǫ
EW (pm)
(pm)
3600302
ǫ
EW
(pm)
3700178⋆
ǫ
EW
(pm)
3800199
-1.76
-1.36
-1.44
-0.12
-1.21
-0.46
-1.93
-3.30
-1.55
-1.72
-2.22
-2.75
-1.07
-3.16
-4.95
-4.35
-2.47
-0.522
-1.729
-1.945
-2.165
0.240
-1.490
-0.315
-0.797
0.390
-1.290
-0.109
-0.818
-2.110
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
G03
G03
G03
G03
SR
G03
SR
SR
SR
SR
SR
SR
NBS
-2.045 MFW
-2.145 MFW
-1.890 MFW
-0.329 MFW
-0.496 MFW
-3.630 MFW
-1.944 MFW
-0.423 MFW
-1.379 MFW
6.73
11.04
10.48
14.35
10.62
--
9.35
15.16
9.38
10.02
7.26
15.48
6.60
10.61
12.39
5.59
6.63
26.90
16.13
5.97
6.41
19.10
7.53
30.57
16.63
24.07
14.45
20.98
16.19
7.32
12.47
11.11
13.78
15.62
13.19
3.09
13.25
10.83
7.39
7.701
7.257
7.249
7.397
7.166
--
7.493
7.221
7.341
7.352
7.221
7.638
7.23
7.083
7.381
7.18
7.608
7.019
7.383
7.137
7.428
5.772
5.712
5.856
6.011
5.723
6.063
5.653
5.876
5.657
4.671
4.527
4.754
4.902
4.531
4.703
4.655
4.578
4.374
--
9.78
9.60
14.91
10.34
--
7.49
15.05
7.81
10.75
8.91
13.44
5.85
12.58
10.84
7.14
4.56
--
14.74
7.10
6.33
17.91
9.25
26.93
17.27
23.66
15.14
19.90
16.61
6.72
11.71
11.98
12.89
13.83
12.35
3.59
14.44
11.86
7.72
--
6.87
6.952
7.335
6.98
--
7.071
6.991
6.974
7.36
7.448
7.125
7.087
7.271
6.947
7.384
7.238
--
7.102
7.309
7.413
5.4
5.814
5.432
5.785
5.415
5.88
5.418
5.629
5.43
4.235
4.395
4.274
4.184
4.068
4.661
4.532
4.5
4.253
--
8.95
10.97
11.72
9.77
9.99
8.41
14.02
--
8.67
8.27
13.62
5.13
11.40
9.69
5.87
4.00
26.42
13.29
3.88
4.91
20.24
9.00
31.74
17.26
28.40
13.90
20.23
17.58
8.86
11.48
10.54
14.18
13.32
14.25
2.81
13.73
12.48
9.62
--
6.902
7.428
7.037
7.077
7.307
7.388
7.062
--
7.163
7.492
7.39
7.053
7.282
6.914
7.262
7.196
7.003
7.044
6.841
7.246
5.78
5.859
5.737
6.016
5.894
5.875
5.679
6.019
5.79
4.349
4.285
4.709
4.318
4.649
4.48
4.628
4.783
4.606
4.42
12.59
--
12.74
11.18
9.79
8.95
18.70
9.53
9.35
5.19
13.99
7.75
15.48
11.79
5.65
4.92
25.88
12.35
5.53
--
18.14
10.79
41.80
20.98
27.02
14.39
20.86
15.73
9.14
13.54
13.31
14.58
12.88
15.84
3.24
14.89
14.67
8.87
7.217
7.134
--
6.777
6.929
6.939
7.174
7.176
7.114
6.959
6.761
6.929
7.328
7.423
6.816
7.053
7.243
6.831
6.621
7.042
--
5.09
5.788
5.91
5.916
5.399
5.392
5.125
5.102
5.527
4.106
4.192
4.102
3.629
4.209
4.382
4.147
4.556
4.141
3.06
8.77
7.28
9.52
7.11
5.76
4.23
11.90
5.69
6.20
3.07
10.83
2.73
8.72
7.40
--
--
20.68
10.03
2.45
1.96
15.44
3.81
24.84
13.52
21.83
10.05
18.14
13.18
--
8.06
7.33
8.04
11.25
8.98
--
10.04
6.40
3.56
ǫ
6.832
6.514
6.376
6.18
6.235
6.147
6.378
6.428
6.435
6.415
6.321
6.586
6.284
6.555
6.47
--
--
6.617
6.425
6.475
6.575
5.501
5.442
5.805
5.657
5.641
5.564
5.604
5.564
--
4.387
4.412
4.24
4.335
4.144
--
4.502
4.262
4.275
⋆ Star showing TiO molecular bands in the spectra
FMW -- Fuhr et al. (1988)
G03 -- Gratton et al. (2003)
SR -- Smith et al. (1981)
NBS -- Wiese et al. (1969)
MFW -- Martin et al. (1988)
Table A.1. Line list and adopted atomic parameters for the program stars. The measured equivalent width and the corresponding
abundance obtained for each line are also reported (continued)
Ion
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Mg I
Mg I
Mg I
Mg I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
λ
(nm)
585.5076
588.3817
595.2718
602.4058
602.7051
605.6005
609.6664
615.1617
616.5360
618.7989
622.6734
651.8366
659.7559
670.3566
673.9521
674.6954
679.3258
552.8405
571.1088
631.8717
631.9237
585.7451
586.7562
612.2217
616.9042
643.9075
645.5558
649.3781
649.9650
650.8850
588.0269
590.3315
593.7809
595.3160
597.8541
601.6995
606.4626
609.1171
609.2792
log gf
source of
log gf
(see notes)
EW
(pm)
3800204
ǫ
EW
(pm)
3800318
ǫ
EW (pm)
(pm)
3800319
-1.76
-1.36
-1.44
-0.12
-1.21
-0.46
-1.93
-3.30
-1.55
-1.72
-2.22
-2.75
-1.07
-3.16
-4.95
-4.35
-2.47
-0.522
-1.729
-1.945
-2.165
0.240
-1.490
-0.315
-0.797
0.390
-1.290
-0.109
-0.818
-2.110
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
G03
G03
G03
G03
SR
G03
SR
SR
SR
SR
SR
SR
NBS
-2.045 MFW
-2.145 MFW
-1.890 MFW
-0.329 MFW
-0.496 MFW
-3.630 MFW
-1.944 MFW
-0.423 MFW
-1.379 MFW
--
7.89
5.21
9.69
5.84
5.00
12.91
1.13
4.00
4.86
2.25
9.51
--
6.42
5.76
--
--
19.95
10.90
--
--
14.48
--
--
11.15
--
8.64
16.22
11.42
--
3.97
4.44
4.94
8.42
6.43
--
6.83
4.41
1.94
--
6.138
5.89
5.95
5.867
5.896
6.007
5.933
6.027
6.05
6.014
6.086
--
6.005
5.998
--
--
6.239
6.313
--
--
4.969
--
--
4.938
--
5.092
4.873
4.958
--
3.56
3.744
3.55
3.599
3.515
--
3.78
3.744
3.711
4.61
9.56
8.47
14.80
9.99
9.71
7.39
15.52
7.74
9.24
7.37
12.81
5.63
11.91
9.47
5.72
--
25.04
14.70
3.54
2.38
23.54
12.35
29.75
17.28
28.72
14.54
23.18
19.68
9.91
15.47
12.62
15.34
17.81
16.98
3.94
16.48
13.42
9.20
7.241
6.771
6.692
7.274
6.858
7.008
6.993
6.987
6.9
7.033
7.128
6.947
6.994
7.083
6.648
7.096
--
6.858
7.06
6.702
6.681
5.949
6.278
5.621
5.77
5.848
5.749
5.826
6.077
5.841
4.822
4.436
4.621
4.875
4.849
4.657
4.829
4.706
4.403
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
18.45
10.81
--
--
13.20
--
18.71
9.69
17.55
7.43
15.61
12.49
--
3.66
3.67
4.30
6.63
7.17
--
6.06
5.51
--
ǫ
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
6.469
6.653
--
--
5.347
--
5.267
5.237
5.28
5.356
5.477
5.687
--
4.069
4.187
4.029
3.917
4.14
--
4.293
4.346
--
EW
(pm)
3800336⋆
--
8.71
8.23
12.01
10.99
7.55
7.17
16.32
7.25
7.53
7.11
12.48
5.52
--
10.34
5.85
--
24.82
14.66
--
4.07
21.15
9.14
--
16.82
25.90
14.79
22.53
18.59
7.59
14.00
12.11
14.55
16.24
14.47
4.33
16.14
13.62
8.93
ǫ
--
6.641
6.675
6.835
7.049
6.689
6.986
7.04
6.853
6.771
7.112
6.862
7.041
--
6.71
7.103
--
6.866
7.101
--
7.067
5.654
5.667
--
5.6
5.529
5.691
5.658
5.828
5.389
4.405
4.19
4.336
4.463
4.261
4.521
4.627
4.626
4.214
⋆ Star showing TiO molecular bands in the spectra
FMW -- Fuhr et al. (1988)
G03 -- Gratton et al. (2003)
SR -- Smith et al. (1981)
NBS -- Wiese et al. (1969)
MFW -- Martin et al. (1988)
Table A.1. Line list and adopted atomic parameters for the program stars. The measured equivalent width and the corresponding
abundance obtained for each line are also reported (continued)
Ion
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Fe I
Mg I
Mg I
Mg I
Mg I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ca I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
Ti I
λ
(nm)
585.5076
588.3817
595.2718
602.4058
602.7051
605.6005
609.6664
615.1617
616.5360
618.7989
622.6734
651.8366
659.7559
670.3566
673.9521
674.6954
679.3258
552.8405
571.1088
631.8717
631.9237
585.7451
586.7562
612.2217
616.9042
643.9075
645.5558
649.3781
649.9650
650.8850
588.0269
590.3315
593.7809
595.3160
597.8541
601.6995
606.4626
609.1171
609.2792
log gf
source of
log gf
(see notes)
EW
(pm)
4207953⋆
ǫ
EW
(pm)
4303773
ǫ
EW (pm)
(pm)
4304445
ǫ
EW
(pm)
4402285
ǫ
EW
(pm)
4408968
-1.76
-1.36
-1.44
-0.12
-1.21
-0.46
-1.93
-3.30
-1.55
-1.72
-2.22
-2.75
-1.07
-3.16
-4.95
-4.35
-2.47
-0.522
-1.729
-1.945
-2.165
0.240
-1.490
-0.315
-0.797
0.390
-1.290
-0.109
-0.818
-2.110
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
FMW
G03
G03
G03
G03
SR
G03
SR
SR
SR
SR
SR
SR
NBS
-2.045 MFW
-2.145 MFW
-1.890 MFW
-0.329 MFW
-0.496 MFW
-3.630 MFW
-1.944 MFW
-0.423 MFW
-1.379 MFW
5.63
11.23
--
12.90
11.81
8.70
7.32
16.99
8.92
9.47
8.24
14.04
6.24
12.82
--
5.90
4.09
26.88
13.37
6.26
3.64
17.97
8.13
34.59
19.00
26.92
13.85
20.17
17.16
8.22
16.14
13.87
14.79
16.82
14.40
3.52
--
12.29
8.93
7.408
7.057
--
6.953
7.169
6.822
6.959
7.166
7.082
7.052
7.249
7.11
7.087
7.193
--
7.078
7.068
6.953
6.858
7.135
6.927
5.36
5.588
5.836
6.02
5.7
5.593
5.405
5.648
5.556
4.879
4.574
4.453
4.627
4.309
4.51
--
4.462
4.3
3.06
10.20
8.60
13.07
--
8.52
6.15
14.48
5.83
7.88
4.73
11.96
--
11.45
9.24
--
4.42
21.07
13.48
2.48
2.94
20.31
6.72
27.12
16.17
23.20
10.57
19.33
16.77
5.68
10.28
8.84
10.84
12.91
11.65
--
--
8.95
5.96
6.896
6.836
6.666
6.936
--
6.741
6.739
6.781
6.531
6.751
6.647
6.765
--
6.965
6.586
--
7.103
6.563
6.856
6.456
6.775
5.702
5.47
5.529
5.619
5.406
5.163
5.347
5.655
5.307
4.003
3.92
3.938
4.008
3.942
--
--
4.05
4.024
4.90
11.99
--
13.48
11.50
9.16
8.17
13.99
8.76
10.19
7.65
14.87
7.29
11.33
10.66
6.41
3.60
--
15.72
6.43
--
19.18
7.18
28.52
17.49
23.37
14.76
20.40
15.94
5.06
11.33
10.86
11.13
15.24
13.06
1.70
12.78
10.64
6.83
7.28
7.326
--
7.139
7.219
6.936
7.164
6.905
7.113
7.27
7.205
7.422
7.264
7.101
6.97
7.25
6.988
--
7.268
7.161
--
5.761
5.663
5.794
6.103
5.637
6.059
5.755
5.77
5.34
4.419
4.457
4.231
4.731
4.444
4.429
4.513
4.515
4.306
4.63
--
9.46
13.24
10.76
--
7.72
13.75
9.79
9.90
6.40
12.52
7.17
10.31
9.42
4.36
5.86
24.18
13.61
7.58
3.80
18.74
6.97
24.80
18.10
24.34
13.22
17.64
15.75
5.71
9.77
10.49
11.23
14.01
12.22
1.85
12.89
10.41
6.76
7.275
--
7.076
7.238
7.235
--
7.191
7.143
7.447
7.366
7.078
7.254
7.293
7.13
7.021
7.058
7.463
6.949
7.108
7.389
6.995
5.982
5.865
5.827
6.549
6.013
6.149
5.704
6.106
5.677
4.594
4.837
4.7
4.95
4.722
4.818
5.011
4.853
4.63
5.11
12.23
8.97
14.22
10.57
10.45
8.38
16.13
9.19
9.18
6.36
15.48
6.75
11.01
10.28
6.19
5.15
23.99
12.82
--
3.20
20.40
8.81
30.80
17.85
25.33
14.53
21.24
18.30
7.25
11.74
12.52
13.69
14.05
13.10
1.95
14.13
11.87
8.42
ǫ
7.309
7.2
6.75
7.087
6.91
7.048
7.137
7.114
7.105
6.993
6.966
7.373
7.126
6.972
6.888
7.253
7.278
6.79
6.774
--
6.862
5.784
5.888
5.847
5.96
5.678
5.887
5.675
5.97
5.676
4.461
4.692
4.597
4.368
4.357
4.63
4.665
4.662
4.568
⋆ Star showing TiO molecular bands in the spectra
FMW -- Fuhr et al. (1988)
G03 -- Gratton et al. (2003)
SR -- Smith et al. (1981)
NBS -- Wiese et al. (1969)
MFW -- Martin et al. (1988)
|
astro-ph/0603062 | 1 | 0603 | 2006-03-02T21:00:18 | A 22 Degree Tidal Tail for Palomar 5 | [
"astro-ph"
] | Using Data Release 4 of the Sloan Digital Sky Survey, we have applied an optimal contrast, matched filter technique to trace the trailing tidal tail of the globular cluster Palomar 5 to a distance of 18.5 degrees from the center of the cluster. This more than doubles the total known length of the tail to some 22 degrees on the sky. Based on a simple model of the Galaxy, we find that the stream's orientation on the sky is consistent at the 1.7 sigma level with existing proper motion measurements. We find that a spherical Galactic halo is adequate to model the stream over its currently known length, and we are able to place new constraints on the current space motion of the cluster. | astro-ph | astro-ph | A 22◦ Tidal Tail for Palomar 5
C. J. Grillmair
Spitzer Science Center, 1200 E. California Blvd., Pasadena, CA 91125
[email protected]
and
O. Dionatos1
INAF - Osservatorio Astronomico di Roma, via de Frascati 33, 00040, Monteporzio
Catone, Italy
[email protected]
ABSTRACT
Using Data Release 4 of the Sloan Digital Sky Survey, we have applied an
optimal contrast, matched filter technique to trace the trailing tidal tail of the
globular cluster Palomar 5 to a distance of 18.5 degrees from the center of the
cluster. This more than doubles the total known length of the tail to some
22 degrees on the sky. Based on a simple model of the Galaxy, we find that
the stream's orientation on the sky is consistent at the 1.7σ level with existing
proper motion measurements. We find that a spherical Galactic halo is adequate
to model the stream over its currently known length, and we are able to place
new constraints on the current space motion of the cluster.
Subject headings: globular clusters: general -- globular clusters: individual(Palomar
5) -- Galaxy: Structure -- Galaxy: Halo
6
0
0
2
r
a
M
2
1
v
2
6
0
3
0
6
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1.
Introduction
The tidal tails of globular clusters are very interesting from a dynamical standpoint as
they are expected to be very cold (Combes, Leon, & Meylan 1999). This should make them
1Department of Physics, University of Athens, Pan/polis, 15771 Athens, Greece
-- 2 --
useful for constraining not only the orbits of the clusters themselves, but for probing both
the global mass distribution of the Galaxy and its lumpiness (Murali & Dubinski 1999).
Tidal tails in globular clusters were first discovered in photographic surveys (Grillmair et al.
(1995); Leon, Meylan, & Combes (2000)) and now number over 30 in both the Galaxy and
in M 31 (Grillmair et al. 1996). Most recently, Grillmair & Johnson (2006) detected a tidal
stream associated with NGC 5466 which subtends some 45◦ on the sky.
Among the first discoveries in the Sloan Digital Sky Survey (SDSS) data were the
remarkably strong tidal tails of Palomar 5 (Odenkirchen et al. 2001; Rockosi et al. 2002;
Odenkirchen et al. 2003), spanning over 10◦ on the sky. Grillmair & Smith (2001) found
evidence for a stellar mass function relatively depleted of lower mass stars, and Koch et
al.
(2004) found that these lower mass stars have indeed made their way into Pal 5's tidal
stream. Pal 5 may well be on its last orbit around the Galaxy before dissolving completely.
In this paper we examine Data Release 4 of the SDSS, which encompasses an area not
(2003). We briefly describe our analysis in Section 2. We
available to Odenkirchen et al.
discuss our findings in Section 3, and make concluding remarks in Section 4.
2. Data Analysis
Data comprising u′, g′, r′, i′, and z′ photometric measurements and their errors for 3.7 ×
106 stars in the region 224◦ < R.A. < 247◦ and −3◦ < δ < +10◦ were extracted from the
SDSS DR4 database using the SDSS CasJobs query system. The data were analyzed using
the matched-filter technique described by Rockosi et al.
(2002), and the reader is referred
to this paper for a complete description of the method. Briefly, we constructed an observed
color-magnitude density or Hess diagram for Pal 5 using stars within 5′ of the cluster center.
An optimum star count weighting function was created by dividing this color-magnitude
distribution function by a similarly binned color-magnitude distribution of the field stars.
The Hess diagram for field stars was created using several regions to the east and west of
Pal 5 and the stream, together subtending ∼ 147 square degrees. The resulting weighting
function was then applied to all stars in the field and the weighted star counts were summed
by location in a two dimensional array.
We used all stars with 16 < g′ < 22, dereddened as a function of position on the sky
using the DIRBE/IRAS dust maps of Schlegel, Finkbeiner, & Davis (1998). The distribution
of E(B − V ) across the field is shown in Figure 1 and ranges from 0.02 along the northern
edge of the field to 0.35 along the southern edge.
We optimally filtered the g′ − u′, g′ − r′, g′ − i′, and g′ − z′ star counts independently and
-- 3 --
then co-added the resulting weight images. As expected, most of the filter weight is given
to main sequence turn-off and horizontal branch stars, as these populations lie blueward
of the vast majority of field stars. In Figure 2 we show the co-added, filtered star count
distribution for stars on the giant branch and the main sequence. The image has been
smoothed with a Gaussian kernel of width σ = 0.15◦. A low-order, polynomial surface was
fitted and subtracted from the image to remove large scale gradients due to the Galactic disk
and bulge. The blank area running eastward from just north of Pal 5 and the several blank
areas on the eastern end of the field are regions not included in the SDSS DR4. We note
that the region just north of the cluster (containing the globular cluster M 5) is included in
the dataset considered by Odenkirchen et al. (2003), and the reader is referred to this work
for a glimpse of Pal 5's trailing tail in the region 1.3◦ < r < 3.2◦ from the cluster.
3. Discussion
Figure 2 clearly shows a long stream of stars with Pal 5's color-luminosity distribution,
extending from R.A., dec = (226.3◦, −2.9◦) to (246◦, +7.9◦). The southern, leading tail has
already been described in detail by Odenkirchen et al.
(2003), as has the portion of the
northern tail westward of R.A. = 234◦. The new result here is the additional 12◦ of trailing
tidal tail extending to R.A. = 246◦. The feature is quite strong, and appears as a continuous
extension of the trailing tail identified by Odenkirchen et al.
(2003). The feature vanishes
if the weighting filter is shifted 0.5 magnitudes blueward or redward of the Pal 5's observed
main sequence.
Comparing Figures 1 and 2, there are clear correlations between regions of high color
excess and areas of reduced star count density. (e.g. 2◦ SE of the cluster, and at R.A., dec =
239◦, 3.3◦). This is to be expected, given that SDSS sample completeness will be a function
of apparent magnitude. The distribution of foreground dust will affect both the apparent
shape and the number counts along Pal 5's tidal tails. The effect of the dust can be reduced
by counting only the brighter stars, but at some considerable cost to signal to noise ratio.
This is particularly true at the eastern end of the tail, where contamination by foreground
bulge stars is almost twice what it is in the vicinity of the cluster itself.
To investigate whether the additional arc of Pal 5's tail could be due to confusion with
background galaxies we have reanalyzed the same survey area using as input those objects
classified in DR4 as galaxies. We find no indication of significant enhancements in the galaxy
counts, either extended along the feature, or as discrete components of it. We conclude that
star-galaxy confusion cannot be held to account for the extended northern feature.
-- 4 --
For all these reasons, we conclude that the feature extending from R.A., dec = (234◦, 3.2◦)
to (246◦, 7.9◦) is a bona fide portion of Pal 5's trailing tidal tail. As found by Odenkirchen et
al. (2003) for the inner portion of the tails, the linear density of stars continues to fluctuate
along the extended tail, rising to more than 200 stars per linear degree at R.A. ≈ 240.6◦, and
dropping almost to indetectablity at R.A. ≈ 236.8◦ and 242.8◦. We do not expect significant
diminution in the filter output due to changes in the average distance between the Sun and
the stream. Based on the best-fit orbit model described below, the distance between us and
the stream increases from 23.2 kpc at the cluster itself to 23.9 kpc at apogalacticon (which
occurs at R.A. ≈ 235◦), dropping back down to 23.2 kpc at R.A. = 246◦. Given the pho-
tometric uncertainties, this variation in distance modulus (0.06 mag) is too small to have
any noticeable effect on the output of a matched filter which is optimized for stars at the
distance of Pal 5. The density fluctuations along the stream must therefore be real and are
presumably a natural consequence of the episodic nature of tidal stripping of a cluster on an
eccentric and disk-crossing orbit.
In Figure 3 we show the distribution of 78 candidate blue horizontal branch stars. There
is no obvious tendency for these stars to be distributed along the tidal tails - only four stars
outside the cluster itself could be said to lie along the tails. This is probably not surprising,
however, given that Pal 5 has very few blue HB stars; most of the points in Figure 3 are
likely to be foreground contaminants.
Assuming that tidal streams closely parallel the orbits of their parent clusters (Odenkirchen
et al. 2003), we can use the observed orientation of the extended stream to better constrain
the current velocity and global orbit of Pal 5. Using the analytic Galactic model of Allen &
Santillan (1991) (which includes a disk as well as a spherical bulge and halo), we integrate
along the orbit of Pal 5 both forwards and backwards and project this path onto the sky. We
adopt R⊙ = 8.5 kpc and vc = 220 km s−1, and use Odenkirchen et al.
(2002)'s heliocentric
radial velocity measurement for Pal 5 of −58.7 ± 0.2 km s−1
Using Cudworth's 1998 (unpublished, but listed in Dinescu, Girard, & van Altena
(1999)) proper motion measurement of µα cos(δ), µδ = (−2.55 ± 0.17, −1.93 ± 0.17) mas
yr−1, we arrive at an expected orbit projection as shown by the solid line in Figure 3. Other
proper motion measurements for Pal 5 (e.g. Schweizer, Cudworth, & Majewski
(1993);
Scholz et al.
(1998)) result in orbit projections which are almost perpendicular to the
stream in Figure 3. With due allowance for our incomplete knowledge of the Galactic po-
tential and for the limitations of Allen & Santillan (1991)'s model, the projected orbits
implied by the existing proper motion measurements do not agree particularly well with the
orientation of Pal 5's tidal tails.
Based on the width of the "S" curvature of the tails in the immediate vicinity of Pal
-- 5 --
5, we adopt a mean offset between Pal 5's projected orbit and the centerlines of the tails of
0.2◦. We then define a set of normal points which trace the highest surface densities along
the tails in Figure 2. Finally, we project orbits for a grid of possible proper motions and use
χ2 minimization to obtain a "best fit". The minimum χ2 solution is shown by the dotted
line in Figure 3.
Over the length of the stream for which we have good signal to noise ratio, our best fit
orbit parallels the stream contours quite well. Departures appear to be of high order rather
than systematic, indicating that a spherical halo potential is adequate to fit the data. There
is a suggestion that the southern tail and the model may be parting company near the limits
of the field, and it will be interesting to see whether the simple potential used here can be
made to fit the continuation of the southern tail when additional data become available.
The proper motion which corresponds to our best fit orbit has µα cos(δ), µδ = (−2.27 ±
0.08, −2.19 ± 0.03) mas yr−1, where the uncertainties correspond to the 99% confidence
interval. We note that, while the uncertainties are influenced by measurement errors, random
motions of stream stars, confusion with the foreground population, variable extinction across
the stream, they do not incorporate uncertainties in our adopted model for the Galaxy.
Our modeled proper motions are within ∼ 1.7σ of Cudworth's measured values. For the
orbit shown in Figure 3, the corresponding space velocities of Pal 5 are U,V,W = (48.3 ±
2.0, −334 ± 5, −14.2 ± 2.1) km s−1. The (radial) period of the orbit is 2.9 × 108 yrs, with
peri and apogalacticon of 7.9 and 18.8 kpc, respectively. These values agree very well with
similar estimates by Odenkirchen et al.
(2003) based on their less extensive data set.
The best fit orbit predicts a total length of ≃ 8.3 kpc for the northern arm. The width
of the tail appears roughly constant along its length, though the signal to noise ratio is
low enough that we cannot rule out a possible widening of the tail beyond R.A. = 241.6.
Comparing the contours in Figure 3 with the best fit orbit, there is also a suggestion of
"wobbling" of the stream eastwards of R.A. = 239◦, and perhaps a 0.5◦ northward jog of
the tail at 241.5◦ < R.A. < 244.5.
If borne out by future kinematic data, this could be
evidence for irregularities in the Galactic potential or for a weak encounter between stream
stars and a substantial mass concentration (e.g. a large molecular cloud) during a recent
passage through the disk. A more extensive analysis of the stream and its consequences will
be presented in a forthcoming paper.
-- 6 --
4. Conclusions
Applying optimal contrast filtering techniques to SDSS DR4 data, we have detected a
continuation of Pal 5's trailing tidal stream out to almost 19◦ from the cluster. Combining
this with the already known southern tail of Pal 5 yields a stream some 22.5◦ long on the sky.
The extension of the stream shows marked fluctuations in surface density along its length
which are presumably a natural consequence of the episodic nature of tidal stripping. The
orientation of the stream is in accord (at the 1.7σ level) with one of the existing measurements
of Pal 5's proper motion. The stream can be modeled reasonably well using a model Galactic
potential with a spherical halo, and we are able to place much better constraints on the
current space motion of the cluster.
We can use the current data set to assign to each star a probability that it is associated
with Pal 5's tidal stream. These probabilities can be taken to the telescope, where we will
need to measure radial velocities sufficiently accurately to unambiguously tie the stars to
the stream. Ultimately, the vetted stream stars will become prime targets for the Space
Interferometry Mission, whose proper motion measurements will enable very much stronger
constraints to be placed on both the orbit of the cluster and on the potential field of the
Galaxy.
We are grateful to an anonymous referee for observations which improved the clarity of
the manuscript.
Funding for the creation and distribution of the SDSS Archive has been provided by
the Alfred P. Sloan Foundation, the Participating Institutions, the National Aeronautics and
Space Administration, the National Science Foundation, the U.S. Department of Energy, the
Japanese Monbukagakusho, and the Max Planck Society.
Facilities: Sloan.
REFERENCES
Allen, C., & Santillan, A. 1991, Rev. Mex. Astron. Astrofis., 22, 255
Combes, F., Leon, S., & Meylan, G. 1999, A&A, 352, 149
Dinescu, D. I., Girard, T. M., & van Altena, W. F. 1999, AJ, 117, 1792
Grillmair, C. J., Freeman, K. C., Irwin, M., & Quinn, P. J. 1995, AJ, 109, 2553
-- 7 --
Grillmair, C. J., Ajhar, E., Faber, S. M., Baum, W. A., Lauer, T. R., Lynds, C. R., & O'Neil,
Jr., E. 1996, AJ, 111, 2293
Grillmair, C. J., & Smith, G. H. 2001, AJ, 122, 3231
Grillmair, C. J., & Johnson, R., 2006, ApJ, in press.
Koch, A., Grebel, E. K., Odenkirchen, M., Martinez-Delgado, D., & Caldwell, John A. R.
2004, AJ, 128, 2274
Leon, S., Meylan, G., & Combes, F. 2000, A&A, 359, 907
Murali, C., & Dubinski, J. 1999, AJ, 118, 911
Odenkirchen, M., Brosche, P., Geffert, M., Tucholke, H.-J. 1997, New Astronomy, 2, 477
Odenkirchen, M., Grebel, E. K., Rockosi, C. M., Dehnen, W., Ibata, R., Rix, H., Solte, A.,
Wolf, C., Anderson, J. E. Jr., Bahcall, N. A, Brinkmann, J., Csabai, I., Hennessy, G.,
Hindsley, R. B., Ivezic, Z., Lupton, R., Munn, J. A., Pier, J. R., Stoughton, C., York,
D. G. 2001, ApJ, 548, 1650
Odenkirchen, M., Grebel, Dehnen, Rix, H.-W, & Cudworth, K. M. 2002, AJ, 124 1497
Odenkirchen, M., Grebel, E. K., Dehnen, W., Rix, H., Yanny, B., Newberg, H., Rockosi, C.
M., Martinez-Delgado, D., Brinkmann, J., Pier, J. R. 2003, AJ, 126, 2385
Odenkirchen, M., & Grebel, E. K. 2004, in Satellites and Tidal Streams, ASP Conference
Series, Vol. 327, eds. F. Prada, D. Martinez Delgado, & T. J. Mahoney
Rockosi, C. M., Odenkirchen, M., Grebel, E. K., Dehnen, W., Cudworth, K. M., Gunn, J.
E., York, D. G., Brinkmann, J., Hennessy, G. S., & Ivezic, Z. 2002, AJ, 124, 349
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525
Scholz, R, -D., Irwin, M., Odenkirchen, M., & Meusinger, H. 1998, a, 333, 531
Schweizer, A. E., Cudworth, K. M., & Majewski, S. R. 1993, in ASP Conf. Ser. 48, The
Globular Cluster-Galaxy Connection, ed G. H. Smith & J. P. Brodie (San Francisco:
ASP), 113
This preprint was prepared with the AAS LATEX macros v5.2.
-- 8 --
Fig. 1. -- Distribution of E(B-V) over the DR4 field surrounding Pal 5. The position of Pal
5 is indicated by the open square at R.A., dec = (229,-0.11). The color excess ranges from
0.02 along the northern border of the image, to 0.3 along the southern edge.
-- 9 --
Fig. 2. -- Smoothed, summed weight image of the SDSS field after subtraction of a low-order
surface fit. Darker areas indicate higher surface densities. The location of Pal 5 is indicated
by the open square at R.A., dec = (229,-0.11). The weight image has been smoothed with a
Gaussian kernel of width of 0.15◦. The irregular borders and the missing stripe are defined
by the limits of SDSS Data Release 4. The newly discovered extension of Pal 5's tidal stream
extends from R.A., dec = (234◦, 3.2◦) to (246◦, 7.9◦).
-- 10 --
Fig. 3. -- Surface density contours and the orbit of Pal 5. Contours are taken from the
weight image in Figure 2 and represent the 2, 3, 4, and 5σ levels. The filled circles show the
positions of candidate blue horizontal branch stars. The solid line shows an orbit integration
based on the Cudworth's 1998 measurement of the cluster's proper motion (Dinescu, Girard,
& van Altena 1999). The dotted line shows an orbit with µα cos(δ), µδ = (−2.27, −2.19)
mas yr−1.
|
astro-ph/0401184 | 1 | 0401 | 2004-01-12T10:31:06 | Chandra Observations of the X-ray Environs of SN 1998bw/GRB 980425 | [
"astro-ph"
] | (Abrigded) We report X-ray studies of the environs of SN 1998bw and GRB 980425 using the Chandra X-Ray Observatory 1281 days after the GRB. Combining our observation of the supernova with others of the GRB afterglow, a smooth X-ray light curve, spanning ~1300 days, is obtained by assuming the burst and supernova were coincident at 35.6 Mpc. When this X-ray light curve is compared with those of the X-ray ``afterglows'' of ordinary GRBs, X-ray Flashes, and ordinary supernovae, evidence emerges for at least two classes of lightcurves, perhaps bounding a continuum. By three to ten years, all these phenomena seem to converge on a common X-ray luminosity, possibly indicative of the supernova underlying them all. This convergence strengthens the conclusion that SN 1998bw and GRB 980425 took place in the same object. One possible explanation for the two classes is a (nearly) standard GRB observed at different angles, in which case X-ray afterglows with intermediate luminosities should eventually be discovered. Finally, we comment on the contribution of GRB afterglows to the ULX source population. | astro-ph | astro-ph |
Submitted to ApJ, June 2, 2018
Chandra Observations of the X-ray Environs of
SN 1998bw/GRB 980425
C. Kouveliotou1,2, S. E. Woosley3, S. K. Patel2, A. Levan4, R. Blandford5, E.
Ramirez-Ruiz6, R.A.M.J. Wijers7, M. C. Weisskopf1, A. Tennant1, E. Pian8, P. Giommi9
[email protected]
ABSTRACT
We report X-ray studies of the environs of SN 1998bw and GRB 980425 using
the Chandra X-Ray Observatory 1281 days after the GRB. Eight X-ray point
sources were localized, three and five each in the original error boxes - S1 and S2
- assigned for variable X-ray counterparts to the GRB by BeppoSAX. The sum
of the discrete X-ray sources plus continuous emission in S2 observed by CXO on
day 1281 is within a factor of 1.5 of the maximum and the upper limits seen by
BeppoSAX. We conclude that S2 is the sum of several variable sources that have
not disappeared, and therefore is not associated with the GRB. Within S1, clear
evidence is seen for a decline of approximately a factor of 12 between day 200 and
day 1281. One of the sources in S1, S1a, is coincident with the well-determined
radio location of SN 1998bw, and is certainly the remnant of that explosion. The
1NASA/Marshall Space Flight Center, NSSTC, SD-50, 320 Sparkman Dr., Huntsville, AL 35805, USA
2Universities Space Research Association, NSSTC, SD-50, 320 Sparkman Dr., Huntsville, AL 35805, USA
3Department of Astronomy & Astrophysics, University of California at Santa Cruz, Santa Cruz, CA
95064, USA
4Department of Physics and Astronomy, University of Leicester,University Road, Leicester, LE1 7RH,
UK
5Kavli Institute for Particle Astrophysics and Cosmology, Stanford, CA 94305, USA
6Institute for Advanced Study, Olden Lane, Princeton, NJ 08540, USA; Chandra Fellow
7Astronomical Institute "Anton Pannekoek"and Center for High Energy Astrophysics, University of Am-
sterdam, Kruislaan 403, 1098 SJ Amsterdam, NL
8INAF, Osservatorio Astronomico di Trieste, Via G.B. Tiepolo 11, I-34131 Trieste, I
9ASI Science Data Center, c/o ESRIN, Via G. Galilei, I-00044 Frascati, I
-- 2 --
nature of the other sources is also discussed. Combining our observation of the su-
pernova with others of the GRB afterglow, a smooth X-ray light curve, spanning
∼ 1300 days, is obtained by assuming the burst and supernova were coincident at
35.6 Mpc. When this X-ray light curve is compared with those of the X-ray "af-
terglows" of ordinary GRBs, X-ray Flashes, and ordinary supernovae, evidence
emerges for at least two classes of lightcurves, perhaps bounding a continuum.
By three to ten years, all these phenomena seem to converge on a common X-ray
luminosity, possibly indicative of the supernova underlying them all. This con-
vergence strengthens the conclusion that SN 1998bw and GRB 980425 took place
in the same object. One possible explanation for the two classes is a (nearly)
standard GRB observed at different angles, in which case X-ray afterglows with
intermediate luminosities should eventually be discovered. Finally, we comment
on the contribution of GRBs to the ULX source population.
Subject headings: gamma-ray bursts
1.
Introduction
One of the most exciting developments in the study of gamma-ray bursts (GRBs) was
the discovery, in 1998, of a GRB apparently in coincidence with a very unusual supernova
of Type Ic (Galama et al. 1998). This coincidence of SN 1998bw and GRB 980425 offered
compelling evidence that GRBs are indeed associated with the deaths of massive stars,
and that, at least in some cases, GRBs go hand in hand with stellar explosions (Woosley
1993; MacFadyen & Woosley 1999). The large energy release inferred for the supernova also
suggested a novel class of explosions, called by some "hypernovae" (Paczy´nski 1998), having
unusual properties of energy, asymmetry, and relativistic ejecta.
However, this identification was challenged on two grounds. First, there were two vari-
able X-ray sources identified with BeppoSAX in the initial 8.0′ radius GRB error box; one
was not the supernova. Second, if it were associated with the nearby supernova, GRB 980425
was a most unusual burst with gamma-ray energy per solid angle roughly four orders of mag-
nitude less than typical. Further, the BeppoSAX decay of the SN 1998bw associated X-ray
source was much slower than the typical GRB X-ray afterglow decay (Pian et al. 2000).
Interestingly, observations (Pian et al. 2003) of the two sources in March 2002 with the
X-ray Multi-Mirror (XMM) telescope revealed that the X-ray emission of the SN 1998bw
associated BeppoSAX error box had decreased at a faster pace than expected by a simple
extrapolation of the earlier measurements. Moreover, the second XMM source was found
to consist of a number of faint point-like sources, whose integrated emission was consistent
-- 3 --
with the average brightness measured with BeppoSAX (Pian et al. 2000).
Strong support for the GRB-SN association came from the very recent spectroscopic
detection (Hjorth et al. 2003; Stanek et al. 2003) of a supernova (SN 2003dh) in the optical
light curve of GRB 030329. In this case, the SN detection was obscured by the extreme
optical brightness of the GRB afterglow. While the models of MacFadyen and Woosley
(1999) suggest that there can be adequate 56Ni to make the GRB-supernova bright, it is
now becoming evident that we must also account for the potentially much brighter optical
afterglow of the GRB itself.
Here we report the results of a study of the environs of SN 1998bw and GRB 980425
using the Chandra X-Ray Observatory (CXO) 1281 days after the GRB. This study had
several goals. First, given the intervening three years, has the evidence strengthened for the
GRB-SN association? We believe that it has (§ 4.1). Second, how does the X-ray light curve
of GRB 980425, measured across 1300 days, compare with those of other GRBs and with
other kinds of high energy transients - in particular X-ray flashes (XRFs; Heise 2003) and
supernovae? What does the comparison tell us about the nature and origin of GRBs? We
find that it provides evidence for a common theme underlying all these events - a powerful
asymmetric supernova with relativistic ejecta along its polar axes and an observable event
that varies depending upon the viewer's polar angle (§ 4 and § 5).
Finally, we are interested in the environs of SN 1998bw. Aside from the one supernova,
does this region show evidence for unusual stellar activity as might characterize a vigorously
star forming region (§ 3.1)? SN 1998bw offers the best opportunity to study a GRB site up
close, and one should take every advantage of that.
2. Prompt Observations of GRB 980425 and SN 1998bw
GRB 980425 triggered the Burst And Transient Source Experiment (BATSE ) on board
NASA's Compton Gamma-Ray Observatory (CGRO) on 1998 April 25, 21:49:09 UT; the
event was simultaneously detected by the BeppoSAX Gamma-ray Burst Monitor (GRBM)
and Wide Field Camera (WFC). The burst consisted of a single peak of ∼ 23 s duration, with
peak flux and fluence (24-1820 keV) of (3.0 ± 0.3) × 10−7 erg/cm2 s, and (4.4 ± 0.4) × 10−6
erg/cm2, respectively. Galama et al. (1998) observed the WFC 8.0′ error box with the New
Technology Telescope (NTT) at the European Southern Observatory (ESO) on April 28.4
and May 1.3 UT and, in the error box of GRB 980425, they found supernova SN 1998bw,
located in an HII region in a spiral arm of the face-on barred spiral galaxy ESO 184−G82,
at z=0.0085, corresponding to a distance of 38.5 Mpc (Galama et al. assumed a Hubble
-- 4 --
constant, H0 = 65 km/s Mpc, but in the following we use H0 = 72 ± 8.0 km/s Mpc as
measured by Freedman et al. (2003), placing SN 1998bw at 35.6 Mpc).
On 1998 April 26-28, ten hours after the GRB, the BeppoSAX Narrow Field Instruments
(NFI) observed (Pian et al. 2000) the WFC error box and revealed two previously unknown,
weak X-ray sources, 1SAX J1935.0 −5248 (hereafter S1) and 1SAX J1935.3 −5252 (hereafter
S2) with an uncertainty radius of 1.5′ each. S1 included the SN 1998bw, but S2 was 4.5′
away. Both sources were observed two more times with the NFI resulting, in the case of S2,
in two detections and two upper limits. In contrast, during the six month interval spanned
by all NFI observations, the flux, F , of S1 followed a power-law temporal decay (Pian et
al. 2000): F2−10keV = 4.3(±0.5) × 10−13(t/1day)−0.2 erg/cm2 s, a much flatter trend than the
one observed for other GRB X-ray afterglows, but a decaying trend nevertheless.
At 35.6 Mpc the apparent isotropic energy of GRB 980425 (7 × 1047 erg) was about four
orders of magnitude smaller than that of 'normal' GRBs (Bloom et al. 2003). Moreover,
independent of its connection with a GRB, SN 1998bw was extraordinary in many ways. Its
light curve resembled a Type Ia supernova in brilliance, but the spectrum was more like Type
Ic (H, He, and Si lines absent, but the spectrum was peculiar even for Ic). Thus, the two
phenomena together presented a very interesting scientific puzzle whose solution required
the combined superb resolution of the Hubble Space Telescope (HST) and CXO.
3. CXO Observations at Day 1281
CXO observed S1 and S2 on 2001 October 27, for a total time on source of 47.7 ks. S1
fell completely on ACIS-S3 (a back-illuminated CCD) and S2 only partially, with most of
the error region falling on ACIS-S2 (a front-illuminated CCD), both operating in Time Ex-
posure (TE) mode. The data were processed using the CIAO (v3.0.1)1 and CALDB(v2.34)2
software. More specifically, we used the CIAO tool acis-process-events to ensure that the
latest gain corrections were applied (those corresponding to our observation date). Further,
we removed the standard pixel randomization, applied CTI corrections, filtered the data to
include events with ASCA grades = 0,2,3,4, and 6, and applied standard GTIs. We cor-
rected the systematic offset in the aspect using the fix-offset thread3; this results in an offset
of δRA = 0.16′′ and δDEC = 0.43′′. To improve the ACIS-S3 spatial resolution we use
1http://asc.harvard.edu/ciao/
2http://asc.harvard.edu/caldb/
3http://cxc.harvard.edu/cal/ASPECT
-- 5 --
the method described by Mori et al. (2001) and Tsunemi et al. (2001) to adjust the event
locations. These studies conclude that by defining the event location on a pixel as a function
of the event grade (rather than placing the event at the pixel center) one can achieve ∼ 10%
improvements in spatial resolution. During our observation the X-ray background increased
by ∼ 50% with variations on time scales of a few kiloseconds. Since the point source emission
is not significantly masked by such slow background variations, we chose to include all the
data in our analysis to preserve our (limited) source counts.
3.1. S1 and S2: Source Identification, Locations and Energetics
We used the source-finding method originally described in Swartz et al. (2002), accepting
as detections all sources with a minimum signal-to-noise ratio (S/N) of 2.6. For source
detection purposes we searched images consisting of data between 0.3 − 8.0 keV to avoid the
ACIS high energy background. We discuss below the sources within the 1.5′ radius (1σ) NFI
error circles of S1 and S2 only.
3.1.1. S2
S2 was resolved into five sources (S2abcde; Figure 1c and Table 1). Given the very
limited count number per source, we combined four of them in one spectral fit with a power
law function and a hydrogen column density, NH = 3.95 × 1020 cm−2 (corresponing to the
Galactic absorption in the line of sight to GRB 980425; Schlegel, Finkbeiner & Davis 1998).
Here we assume a spectral similarity between these sources to mitigate the difficulty of
fitting individual spectra of very few counts each. This order of magnitude approximation is
acceptable within our source statistics (see also Figure 2, where rudimentary count spectra
are presented for S1abc, and S2ce). However, one source (S2d), fell into the gap between
the ACIS CCDs S2 and S3 and its total counts had to be adjusted taking into account
the reduced gap exposure time. Using the spectral index of -1 from the fit, we applied a
conversion factor of 1.89 × 10−11 erg/cm2 s (0.3 − 10.0 keV) per count/sec (0.3 − 8.0 keV)
to the S2d counts and finally estimated the total (combined) flux in S2 (0.3 − 10 keV) to be
3.0(3) × 10−14 erg/cm2 s (here and in Table 1 the numbers in parentheses correspond to the
1σ errors in the last digit).
The BeppoSAX flux for S2 was, however, calculated for a much larger extraction radius
(3′) to account for the extreme faintness of the detection within a 1.5′ radius, which was at
the level of the NFI confusion limit (Pian et al. 2000). Only ∼ 60% of this larger error circle
-- 6 --
is covered with CXO. Using the same source-finding algorithim criterion described above, we
found a total of 18 sources within the enlarged area, with a total flux of S2(3′) = 1.6 × 10−13
erg cm−2 s. This is roughly the flux to be compared to the BeppoSAX S2 value at day 1.
Subsequently we re-calculated the total BeppoSAX flux within S2 at day 1, using the same
spectral function derived with CXO and found it to be 4 × 10−13 erg/cm2 s (with < 2σ
significance). Taking into account the partial coverage with CXO of the NFI error circle of
S2, we estimate that the day 1281 CXO flux value is within a factor of 1.5 of the 3σ detection
limit of BeppoSAX, indicating that at best there was no siginificant variation of the sources
within S2 over the last 3.5 years. This result is consistent with the XMM observations of S2
(Pian et al. 2003). We further discuss the evolution of the light curve of S2 in § 4.1.
3.1.2. S1
We initially identified two sources 36′′ apart within the S1 (1.5′) error region4, one of
which coincided with the location of SN 1998bw. However, further inspection of the latter
resolved this source into two with a radial separation of ∼ 1.5′′ (Figure 1b). We fitted these
sources simultaneously with two 2-D circular Gaussians to better estimate their centroids
and found that they are both consistent with point sources. Hereafter we designate the
X-ray sources detected within the S1 error region as S1a (corresponding to SN 1998bw),
S1b, and S1c. We have fitted a power law to each unbinned non-background subtracted
source spectrum, assuming the same NH as for S2. Our fit parameters are derived using the
C-statistic (Cash 1979), appropriate for low count data (Table 1).
Further, we reanalyzed all archival HST and ESO/Very Large Telescope (VLT) obser-
vations of ESO 184−G82, the host galaxy of SN 1998bw, concentrating on the immediate
environment of the supernova. The small field of view of HST/STIS (∼ 50 ′′) contains only
the sources S1a and S1b. In order to obtain accurate astrometry we, therefore, registered
both the CXO and HST images independently to an R-band observation obtained at the VLT
on 1999 April 18. The 3.4′ × 3.4′ field of the VLT image contained three additional X-ray
sources with apparent optical counterparts which we used to align the two fields. Finally,
we aligned the HST and VLT images using eight, non-saturated point sources present in
each image (both alignments were performed using IRAF and the tasks geomap & geoxy-
tran). We were then able to project the relative position of S1a and S1b onto the HST field
4Here we are only considering the area corresponding to the NFI half power radius of 1.5′, as most of the
BeppoSAX signal for S1 at day 1 was within this area (Pian et al. 2000). An enlarged radius region (3′) for
S1 results in a total flux of 2.7 × 10−13 erg cm−2 s, almost half the BeppoSAX value at day 1
-- 7 --
with a positional accuracy of ∼ 0.′′3. This is shown schematically in Figures 1e,f, where we
zoom in the region around S1a and S1b for both the CXO and HST data. It is obvious in
the HST image that while there is a clearly identified optical counterpart to the SN 1998bw,
there is no variable counterpart within the ∼ 0.′′3 CXO error circle (1σ radius) of S1b; we do
however, see a variable source just outside the error region. Levan et al.
(in preparation)
present a detailed study of this source together with three more transients within a radius of
6′′ of the SN 1998bw as well as narrow field spectroscopy of the SN 1998bw environment. A
counterpart search for the other six sources in S1 and S2 in all available catalog data5 failed
to identify any known objects at their positions.
Table 1: CXO Sources Within S1 and S2
Countsa S/N Indexb
DEC
(◦ ′ ′′)
RA
(hms)
19 35 3.31
19 35 3.23
19 35 0.56
19 35 25.75
19 35 25.22
19 35 24.65
19 35 23.15
19 35 14.03
-52 50 44.8
-52 50 43.4
-52 50 21.2
-52 54 19.0
-52 54 21.2
-52 54 40.7
-52 53 5.6
-52 53 51.9
24
52
27
10
7
22
9
23
4.53
6.47
4.69
2.75
2.66
4.46
2.82
4.19
1.0(3)
1.6(2)
0.5(3)
--
--
1.7(4)
--
0.7(3)
ID
S1a
S1b
S1c
S2ae
S2be
S2c
S2df
S2e
Fluxc
Luminosityc,d
erg cm−2 s−1
8(2) × 10−15
9(1) × 10−15
1.2(2) × 10−14
4(1) × 10−15
3(1) × 10−15
5(1) × 10−15
8(3) × 10−15
1.0(2) × 10−14
erg s−1
1.2 × 1039
1.4 × 1039
--
--
--
--
--
--
abetween 0.3 − 8.0 keV, bfor an energy spectrum of ∝ E−γ, cbetween 0.3 − 10.0 keV, dassuming the source
is in ESO 184−G82 at 35.6 Mpc, efluxes for sources with less than 20 counts are derived using a PL index
of -1.0, and a conversion factor of 1.89 × 10−11 ergs s−1 cm−2 (0.3 − 10.0 keV) per 1 count/s (0.3 − 8.0 keV),
f source flux corrected for decreased exposure
From Table 1, we notice that the X-ray luminosities of S1a and S1b marginally exceed
the Ultra Luminous X-ray source (ULX) threshold luminosity of L = 1×1039 erg/s (Fabbiano
1989).6 Further, inspection of the S1b light curve in Figure 2 reveals that the source was
'on' during the first 20 ks of our CXO observation, remained 'off' for the following ∼ 28 ks
and possibly turned on again during the last 2 ks of the observation. Counting the GRB-SN
source also as a ULX, we then have two of these sources in ESO 184−G82. What is the
5http://heasarc.gsfc.nasa.gov/db-perl/W3Browse/w3browse.pl
6Here we are not discussing the nature of S1c as there is no clear evidence for the association of this
source with ESO 184−G82
-- 8 --
probability that S1b resides in ESO 184−G82, how common are ULXs in galaxies and what
is the nature of S1b, this ultraluminous X-ray transient?
(2001): N(> S) = 1200 × (
To address the first question we use the log N-log S relation for the hard X-ray sources
2×10−15 )−1±0.2 sources/deg2 (the one for
by Giacconi et al.
the soft sources is smaller, so we use this as a conservative upper limit). Substituting for
S = 9 × 10−15 erg/cm2 s, we find N(> S) = 400 sources/deg2. Correspondingly, we expect
0.01 source within a radius of 10′′. We conclude that S1b is most likely associated with the
spiral arm of ESO 184−G82 and is not a background source.
S
38
There are two extensive studies of ULX populations in normal galaxies (Roberts and
Warwick 2000; Humphrey et al. 2003) based on data from Rosat and CXO, respectively.
While both studies find substantial evidence for a correlation between the ULX count number
with star forming mass, the Rosat survey tends to underestimate their count numbers by
a factor of 5 − 10 with respect to the CXO. We have, therefore, used the Rosat luminosity
distribution DN/dL38 = (1.0 ± 0.2)L−1.8
, normalized to a B-band luminosity of 1010L⊙ to
calculate the expected number of ULXs in ESO 184−G82, multiplied by a factor of 10 as
indicated by the CXO survey. We derive a total corrected expected number of 0.28 ULXs in
ESO 184−G82, a factor of ∼ 7 less than the actual observed number of 2. We conclude that
if S1a and S1b are both in ESO 184−G82, there is a somewhat unusual concentration of
ULXs in this galaxy. Finally, ESO 184−G82 is one of 5 members of a galaxy group. In the
CXO data, which covers all members, we do not detect any other X-ray sources from this
galaxy cluster. This practically means that, assuming that all members are at the distance
of ESO 184−G82, we detect only 2 ULXs from the whole cluster (one of which is a GRB/SN
afterglow), both within a radius of 6′′. We discuss the results of narrow band photometry
of ESO 184−G82 together with the properties of other unusual transients found in the HST
data in the SN 1998bw environment in detail in Levan et al. (in preparation).
To test S1b for variability we extracted the source lightcurve in 1000 s bins and per-
formed a Kolmogorov-Smirnov test against a constant flux distribution; this is the prefered
method for the low number of counts (< 100) which are seen in the source. We derive
PKS = 0.05 (where PKS is the probability that the source is constant); thus S1b is consistent
with being variable at the 95% confidence level. The transient nature and the energetics
of S1b indicate a possible microquasar similar to e.g., GRO J1915+105 in our own Galaxy
(Mirabel & Rodriguez 1994; Greiner, Morgan & Remillard 1996). This superluminal source
exhibits transient behavior in a wide variety of time scales and intensities, and has an ap-
parent (isotropic) X-ray luminosity of ∼ 7 × 1039 erg s−1, well above its Eddington limit
(Greiner, Cuby & McCaughrean 2001). Fabbiano et al. (2003) have found in a study with
CXO of the ULX sources in the Antennae galaxies that seven out of the nine ULXs are time
-- 9 --
variable, most likely accreting compact X-ray binaries. Interestingly, the average co-added
spectrum of the Antennae ULXs resembles that of Galactic microquasars (Zezas et al. 2002).
We conclude that most likely S1b is a microquasar in ESO 184−G82.
4. Systematics of the X-ray light curves of High Energy Transients
Figure 3 shows, on a common scale, all GRB afterglows with measurements covering
several tens of days. There are, unfortunately, only a few curves because such measurements
can only be made on GRBs that are relatively nearby, but their light curves should be
illustrative. For GRB 030329, we used data points from both the Rossi X-Ray Timing
Explorer (RXTE) and XMM reported by Tiengo et al. (2003). The original four points of
GRB 980425 were reported by Pian et al.
(2000); the fifth point is the CXO observation
of this paper. For GRBs 021004, 010222, 000926, and 970228, we used data from Sako &
Harrison (2002ab), Bjornsson et al. (2002), Harrison et al. (2001), and Costa et al. (1997),
respectively. We reanalyzed all data, calculated the fluxes in the same energy interval (0.3−10
keV) and then converted them into (equivalent isotropic) luminosities using a cosmology of
ΩM = 0.27, ΩΛ = 0.73, and H0 = 72 km s−1 Mpc−1 to eliminate the distance dependence. As
has been noted many times, GRB 980425 and its early afterglow falls orders of magnitude
below the "ordinary" GRBs.
We then compared these curves with both those of supernovae of all types and with
XRFs. We collected, reanalyzed and converted flux data to luminosities as described above
for the Type IIb SN 1993J (Uno et al. 2002; Swartz et al. 2002), the Type Ic supernovae
1994I and 2002ap (Immler et al. 2002; Sutaria et al. 2003) and the Type II supernovae
1998S and 1999em (Pooley et al. 2002). Each source is indicated with a different symbol in
Figure 3. Finally we added the three XRF afterglow light curves available (011030, 020427,
and 030723; Bloom et al. 2003; Butler et al. 2003) assuming a redshift of z = 1 for each of
them (there are no redshift measurements for any of these sources). The slopes of the XRF
afterglows agree well with those of the typical GRBs and their luminosities are comparable
for the distances assumed. Recently, Soderberg et al. (2003b) have reported the counterpart
identification of XRF 020903 at a redshift z = 0.251. To reflect this lower distance scale for
XRFs, we plot on Figure 3 another set of light curves (dashed lines) corresponding to the
luminosities all XRFs mentioned above would have, when placed at a distance of z = 0.251.
They still fall well within the 'typical' GRB range; thus XRFs would have to be extremely
nearby for their X-ray light curves to be distinct from the generic GRB X-ray afterglow.
The resulting plot is striking in several ways. Despite the huge disparity in initial appear-
ance, there are compelling indications of a common convergence of all classes of phenomena
-- 10 --
plotted - GRBs, GRB 980425, XRFs, and the most energetic supernovae - to a common
resting place, L ∼ 1039 − 1040 erg s−1 about three to ten years after the explosive event.
GRB 980425, being the closest by far of any GRB ever studied, has the virtue of being fol-
lowed all the way to the "burial ground", but a simple logarithmic extrapolation of the GRB
and XRF light curves places them squarely in this region as well. As the collapsar model
has predicted (MacFadyen & Woosley 1999) and observations of SN 2003dh/GRB 030329
have unambiguously confirmed for one case (Hjorth et al. 2003; Stanek et al. 2003), an
energetic supernova is expected to underly all GRBs of the long soft variety (Kouveliotou et
al. 1993). Zhang, Woosley & Heger (2003) have also predicted that a similar supernova will
underly all XRFs.
We proceed now to a wider GRB X-ray afterglow comparison with those plotted on
Figure 3. We have often observed GRB afterglows with temporal decays which cannot be
described by a single power law. Rather, these afterglows initially decay as t−1 or a bit
steeper, and at later times the decay index becomes approximately −2. This steepening is
attributed to the fact that the ultrarelativistic outflow is initially collimated within some
angle θ (typically thought to be of order 10 degrees, Frail et al. 2001; see also Van Paradijs
et al. 2000, and references therein). The transition to a steeper power law marks the time
when the outflow begins to expand laterally, which occurs around the time when θ ∼ 1/Γ,
where Γ is the Lorentz factor of the blast wave. Typically, the 'break time' when this
happens is around 0.3 − 3 days after the burst, but a few cases have much later breaks, if
any (e.g., GRB 970508, 000418; Frail et al. 2001). If this is also true for the X-ray afterglows
shown in Fig. 3, then they should be extrapolated to later times with a steeper power law
than the average of the data shown. In that case, they would reach the luminosity level of
SNe sooner, and after that evolve like the SNe, because the blast wave will have become
non-relativistic. Thereafter, the system may simply evolve like an X-ray supernova with an
energy that has been augmented with that of the initially relativistic blast wave. However,
it is tantalizing that a few systems, e.g., GRB 030329 do not yet show such a steep decay
even at 30-40 days, possibly indicating that other mechanisms than the standard collimated
afterglow contribute to the emission. For example, it has been proposed that the outflows
of GRBs are structured (e.g., M´esz´aros et al. 1998, Rossi et al. 2002, Ramirez-Ruiz et al.
2002), i.e., they eject material with ever lower Lorentz factors at ever larger angles from the
jet axis, and this slower material affects the afterglow at ever later times, making it decay
more slowly. In this context, it is interesting to note that in GRB 030329 there is evidence
for two jet breaks, one at 0.5 days and one at about one week (Berger et al. 2003) from radio
to X-ray data. Moreover, WSRT radio data at 1.4 − 5 GHz indicate that there is probably
an even wider outflow of yet slower material (Rol et al.
It is not clear
however, that this material produces enough X-ray emission to explain the slow late decays,
in preparation).
-- 11 --
and so the cause of these may be altogether different.
From Figure 3 alone, it is not clear whether ordinary GRBs and GRB 980425 are two
distinct classes of events with different X-ray light curves, or they form the boundaries of
a continuum of high energy transients that will eventually fill in the entire left side of the
figure. A simple theoretical interpretation discussed in § 5 favors a continuum of events. As
we discuss, at early times an off-axis observer sees a rising light curve, peaking when the jet
Lorentz factor is ∼ 1/θobs, and approaching that seen by an on-axis observer at late times.
This leads us to believe that the low X-ray luminosity of the recently discovered GRB 031203
may have also been due to its having being viewed substantially off-axis.
4.1. The X-Ray Light Curve of Sources S1 and S2
Figure 4 summarizes the X-ray observations of sources S1 and S2. The first four data
points (or upper limits) up to day 200 are from BeppoSAX (Pian et al. 2000) and do not
resolve individual sources within S1 and S2; our CXO observations on day 1281 do resolve
the sources. We consider two hypotheses: 1) that SN 1998bw and GRB 980425 were the
same event, both happening within S1, and 2) that GRB 980425 occurred within the error
box of S2 at a cosmological distance and was thus a more ordinary GRB. Hence the S2
observations are plotted at distance, z = 1.34, such that the afterglow luminosity on day
1 is comparable to ordinary GRBs (see also § 4). The S1 observations are plotted with an
assumed distance of 35.6 Mpc, the distance to the supernova. The subtraction of the known
fluxes of sources S1b and S1c on day 1281 from the four BeppoSAX points reduces their
values by ∼ 4%, which is less than the size of the symbols used in the plot. We chose instead
to plot the sum of all three S1 sources in Figure 4 (S1sum) to indicate the "expected" flux
from extrapolation of the BeppoSAX X-ray light curve of S1 and compare it with the flux of
the S1a (the SN 1998bw) point only.
In contrast to the curve for S1 or ordinary GRBs (Figure 3), the curve for S2 shows
random variability and is nearly flat (assuming S(3′) for the CXO flux of S2 on day 1281).
It is consistent with a collection of variable X-ray sources, probably distant AGNs, whose
sum sometimes exceeds and at other times falls below the BeppoSAX threshold. The total
flux of the CXO observations of the sources within the larger (3′ radius) error circle of S2 is
within a factor of 1.5 of the brightest flux ever detected by BeppoSAX for S2 (day 1) and
the two upper limits given by BeppoSAX on days 2 and 200. If the GRB occurred within
S2 it either declined more rapidly than any other GRB ever studied before, in which case
the observations of S2 offer no supporting evidence for the connection, or it created a most
unusual afterglow that has not declined, in over three years. Moreover, the afterglow on day
-- 12 --
1281 would have a luminosity orders of magnitude greater than other GRBs after day 50
(Figure 3). The simplest conclusion is that S2 did not contain GRB 980425 and that the
BeppoSAX detection was a collection of variable background sources.
The light curve of S1, on the other hand, shows a gentle decline to day 200, followed by
a rapid fading, by factor of about 12 to 1.1 × 1039 erg s−1, by day 1281 (here we compare
the BeppoSAX value at day 1 to S1sum). This last data point is consistent with the light
curves of other particularly luminous supernovae, e.g., SN 1993J, that may have had high
mass loss rates, but one must take care because SN 1998bw was a Type Ic supernova, which
presumably occurred in a Wolf-Rayet star. Such stars are known to have a high wind velocity,
and hence a low circumstellar density. Bregman et al. (2003) have recently discussed the
X-ray emission of young supernovae and find that most have an X-ray luminosity in the
0.5 − 2 keV energy band less than 2 × 1039 erg s−1. They do point out exceptional cases -
SN 1978K, SN 1996J, SN 1998Z, SN 1995N, and SN 1998S - that have X-ray luminosities
from 1039.5 to 1041 erg s−1 even a decade after the event, but these were all Type II. Given
the exceptionally large kinetic energy inferred for SN 1998bw (Iwamoto et al. 1998; Woosley
et al. 1999), perhaps it is not surprising that its luminosity after three years should place,
e.g., about an order of magnitude above common Type Ic supernovae like SN 1994I.
The simplest conclusion here is that the brilliant emission of S1a during the first days
was the X-ray afterglow of the relativistic ejecta that made GRB 980425. By day 1281
however, we were seeing the energetic, but not especially relativistic ejecta of SN 1998bw
colliding with the presupernova mass loss of its progenitor star. This hypothesis is discussed
further in § 5.
5. Theoretical Interpretation
The X-ray afterglow emission of ordinary GRBs is generally attributed to synchrotron
emission from shocks as the blast encounters the interstellar or circumstellar medium. Some
useful scaling relations for blast waves in which each particle emits a fixed fraction, ǫ, of the
energy it gains in the shock, have been given by Cohen, Piran & Sari (1998).
with
L(t) ∝ t−[(m−3)/(m+1)]−1
m =
ǫ2 + 14ǫ + 9
3 − ǫ
.
(1)
(2)
Fundamentally, 0 < ǫ < 1 and 3 < m < 12, so that L ∝ t−1 to t−22/13 = t−1.69. This
expression is for constant density. Chevalier & Li (2000) also give expressions for the power
-- 13 --
law scaling of afterglow light curves and find for a medium with ρ ∝ r−2,
L(t) ∝ t−α
(3)
with α 1.75 to 2.17 for radiative blast waves and 1.38 to 1.75 for adiabatic blast waves if the
index of the electron power distribution, p, is between 2.5 and 3.0. The curve L = 1046t−1.69
erg s−1 is plotted in Figure 3, and provides a reasonable description of the early X-ray
afterglow lightcurve (when most of the energy is emitted at these frequencies). At observer
times longer than a week, the blast wave would, however, be decelerated to a moderate
Lorentz factor, irrespective of the initial value. The beaming and aberration effects are
thereafter less extreme. If the outflow is beamed, a decline in the light curve is expected
at the time when the inverse of the bulk Lorentz factor equals the opening angle of the
outflow (Rhoads 1997). If the critical Lorentz factor is less than 3 or so (i.e. the opening
angle exceeds 20◦) such a transition might be masked by the transition from ultrarelativistic
to mildly relativistic flow, so quite generically it would be difficult to limit the late-time
afterglow opening angle in this way if it exceeds 20◦. For reasonable conditions then, the
power law declines of both XRFs and luminous GRBs, given in Figure 3, are what might
be expected from very relativistic ejecta slowing in either a constant density medium or a
circumstellar wind.
However, the very slow decline of SN 1998bw needs a different explanation. GRB 980425,
or at least that portion directed at us, was very weak. The total energy of its relativistic
ejecta has been estimated as no more than 3 × 1050 erg (Li & Chevalier 1999) and most of
that probably was not directed toward us. Wieringa, Kulkarni & Frail (1999) discuss the
possibility of two shocks associated with SN 1998bw, a relativistic one from the GRB and
a slower moving, more powerful shock from the supernova, but until very late times when
the blast reaches the termination of the presupernova wind, the thermal emission from the
supernova itself is expected to be weak because of the low density of the wind (Chevalier
2000).
Two possibilities can be considered. First, that GRB 980425 was an "ordinary" (or
somewhat subluminous) GRB observed off-axis. This has been suggested many times (e.g.
Woosley, Eastman & Schmidt 1999; Nakamura 1999; Granot, Panaitescu, Kumar & Woosley
2002; Yamazaki, Yonetoku & Nakamura 2003; Zhang, Woosley & Heger 2003) with different
underlying assumptions regarding the angular distribution of the ejecta. Nakamura (1999)
assumes that the jet has sharp edges and the peripheral emission comes from scattering.
Woosley et al.
(1999) assume that there is a distribution of ejecta energies and Lorentz
factors and that, during the burst, we see only the low energy wing moving toward us. The
other possibility is that GRB 980425 was deficient in energetic gamma-rays at all angles. This
is not incompatible with the fact that it may have ejected 3 × 1050 erg of mildly relativistic
-- 14 --
(Γ > 2) material, but concerns only the very relativistic ejecta, Γ > 200, thought responsible
for harder GRBs.
The X-ray light curve can help to distinguish these two possibilities. First, energetic
as it may be, at early times the X-ray emission of the underlying subrelativistic supernova
is probably negligible. Though the mass loss rate of the Wolf-Rayet star progenitor may
have been high, the wind velocity was also large implying a low circumstellar density and
inefficient supernova emission (Chevalier 2000). Type Ib and Ic supernovae are typically
weak X-ray emitters (Figure 3; Bregman 2003). If we attribute the X-ray emission during
the first 200 days to a weak, but on-axis GRB, then the X-ray afterglow should have faded
with a power law not too different from the other GRBs. The fact that its decay is nearly
flat is inconsistent with any relativistic blast wave in which the electrons emit a constant
fraction of the energy gained in the shock (Cohen, Piran & Sari 1998), even in a constant
density medium, and argues against an explosion with a single energy and Lorentz factor
seen pole on.
The data may be more consistent with a powerful burst seen off-axis (Granot et al.
2002). Even along its axis, the burst may have been weaker than most, but the energy per
solid angle could still have been orders of magnitude greater than along our line of sight,
accounting for most of the ∼ 3 × 1050 erg inferred from the radio (Li & Chevalier 1999).
Because of relativistic beaming, initially we see only the low energy material moving toward
us, but as the core of the jet decelerates, its afterglow is beamed to an increasing angle so
that more and more energetic material becomes visible along our line of sight. Depending
on the geometry, the afterglow luminosity could even temporarily increase. Granot et al.
(2002) consider the appearance of GRB 980425 at various angles and conclude that the
viewing angle needs to have been >
∼3θo with θo the half angle of the most energetic part of
the jet, otherwise the optical afterglow would have contaminated the supernova light curve
unacceptably. One expects that the X-ray light curve would look similar to some of the
plots of Granot et al. with the critical addition of low energy wings of ejecta as calculated
by Zhang, Woosley & Heger (2003). This material would raise the luminosity at early times
when almost nothing is seen of the central jet. A calculation of this sort must be a high
priority for the community.
As time passes, beaming becomes less important and the entire decelerating jet becomes
visible, followed a little later by the underlying supernova. The light curve should then
decline, as it did between days 200 and 1281 in Figure 3. One can estimate the time scale
for when beaming becomes unimportant. For a circumstellar density distribution ρ = Ar−2
Waxman, Kulkarni & Frail (1998), assuming typical mass loss and GRB parameters, estimate
-- 15 --
a radius
rNR =
and a corresponding time
E
4πAc2 = 1.8 × 1018 E52
A∗
cm,
tNR =
rNR
c
= 1.9 (1 + z)
E52
A∗
yr
(4)
(5)
when the explosion has swept up a rest mass comparable to the initial relativistic ejecta.
Here, A∗ = A/5 × 1011 g cm−1, corresponding to a fiducial mass loss rate of 1.0 × 10−5 M⊙
yr−1 for a wind velocity of 1000 km/s, and E52 is the relativistic energy in 1052 erg s−1.
One thus expects that for times of order several years the relativistic energy will have been
radiated away and the emission will become isotropic. We can define this as the onset of the
supernova stage.
Evidence of a close association of the early X-ray emission to the overall GRB phe-
nomenon is also presented by Berger, Kulkarni & Frail (2003). They have studied a sample
of 41 GRB X-ray afterglows and they find a strong correlation between their X-ray isotropic
luminosities (LX,iso) (normalized to t = 10 hours after trigger) and their beaming fractions.
We plot LX,iso as a function of the GRB isotropic equivalent γ−ray energy, Eiso (Figure 5).
The data are taken from Berger, Kulkarni & Frail (2003) and we have added data for GRBs
031203, 030329 and 980425, as indicated on the plot. Several points are striking in this
empirical relation: GRBs 031203 and 980425 fall well within the overall correlation and ex-
tend the association by roughly three orders of magnitude. We have fitted the data without
(dashed line) and with (solid line) all the outliers (980425, 031203,000210, 990705), with a
power law of index 0.61 and 0.72, respectively; we conclude that the data are consistent with
a trend extending roughly six orders of magnitude in X-ray luminosity. Parenthetically, the
two outliers in the plot correspond to GRB 000210 (a 'dark' GRB) and GRB 990705 (a very
bright GRB), and indicate that it may be possible to distinguish GRB subclasses by simply
using their X and γ−ray properties, as also pointed out by Berger, Kulkarni & Frail (2003).
Note that the convergence to a "supernova" at three to ten years here does not require
all XRFs and GRBs to have a bright optical supernova following the GRB. Bright optical
emission is a statement about the radioactivity the supernova made. X-ray emission at 3
years is about its kinetic energy. If the variation in GRB energy and X-ray light curve is
simply an effect of viewing angle, then one expects the empty parameter space in Figures 3
and 5 to fill in with future observations. The recent GRB 031203 (Figures 3, 5) may well be
the first of these 'gap' events (Rodriguez-Pascual et al. 2003; Soderberg et al. 2003a) if the
reported redshift of z = 0.105 (Prochaska et al. 2003) is confirmed.
-- 16 --
6. Conclusions
Our study resulted in the detection, on day 1281, of multiple X-ray point sources in the
two BeppoSAX error boxes S1 and S2 originally given (Pian et al. 2000) for the variable
X-ray counterpart to GRB 980425 (§ 3.1). Based upon the accurately known radio location,
the source S1a is definitely the supernova. The sum of all sources in S2 on day 1281 is similar
to the maximum observed for this error box by BeppoSAX by a factor that is smaller than
1.5. This is consistent with the hypothesis that S2 did not contain the GRB, but instead
some variable X-ray sources that have not disappeared. S1, however, has been consistently
observed to decline for ∼ 1300 days. We conclude that the source S1 contained the GRB
and that SN 1998bw and GRB 980425 were the same explosion.
Additional insight comes when we compare the X-ray light curve of GRB 980425 with
the broader family of high energy transients (§ 4). Comparison of published X-ray light
curves for GRB, XRF and SNR supports a unification hypothesis, similar to that proposed
for AGN. Under this hypothesis all of these sources are associated with standard supernova
explosions of massive stars. We can also distinguish a strong unification hypothesis in which
the properties of the external medium are also standardized. There is considerable, though
still inconclusive, evidence that the strong unification hypothesis is false, though it is still
worth testing. These explosions are conjectured to produce an anisotropic, beamed com-
ponent associated with a decelerating, ultrarelativistic outflow and an unbeamed, isotropic
component associated with the slowly expanding stellar debris. The flux associated with
the beamed component depends upon the observer direction and declines rapidly with ob-
server time; the closer to the axis the larger the flux and the more rapid is the onset to
the typical afterglow decline. As the beamed component decelerates to become entirely
non-relativistic, the observed flux will become independent of orientation. Under the strong
unification hypothesis a one-parameter (inclination) family of X-ray light curves will be pro-
duced, converging asymptotically to a single variation when the beamed component becomes
non-relativistic. If only the unification hypothesis is true, we should still be able to observe
these trends and estimate the inclination. Indeed, this is roughly what we observe when we
plot the isotropic luminosities of GRB, XRF and SNR on a common scale (Fig. 3). It appears
that after three years all explosions are subrelativistic, with X-ray luminosity dominated by
the stellar debris, ∼ 1039 erg s−1. We therefore tentatively identify GRB, XRF and SNR as
similar objects observed with small, medium and large inclination respectively. More specific
to this paper, the observation that SN 1998bw and GRB 980425 follow a smooth light curve,
which fits this pattern supports the claim that they are the same source.
In § 5 we discussed two possible interpretations of these light curves based upon either
a standard phenomenon viewed at different angles or explosions that eject variable amounts
-- 17 --
of relativistic ejecta. We conclude that the observations, especially the slow initial decline
rate, are more consistent with the "off-axis model" in which GRB 980425 was a much more
powerful GRB seen at an angle greater than about three times the opening angle of the
central jet. Emission at early times does not come from this central jet but from wings of less
energetic material. After about three years the emission of all these high energy transients
becomes isotropic and we see the sub-relativistic ejecta of the supernova interacting with
the circumstellar wind. Thus all high energy transients have a common luminosity at three
years due to their non-relativistic ejecta, but they follow different decay rates, depending
upon the viewing angle, to reach there.
Further, we discussed the stellar environment in that region of the galaxy ESO 184−G82
where the supernova occurred. The supernova is one member of an X-ray doublet, both of
which seem to be in the galaxy and to have spend part of their life cycles as ULXs. The
projected distance between the two X-ray sources is ∼ 300 pc, which is suggestive of some
sort of a very active star forming region.
Finally, we would like to stress that the existence of a relation between the decline
rate of the X-ray light curve during the first few weeks and its brightness implies that such
measurements might be useful for diagnosing the character and subsequent evolution of a
given high energy transient as well as constraining its distance. However, the sparse coverage
of the current X-ray afterglow data does not allow us to address fundamental questions such
as: Did the rapid decline of GRB 980425 continue? Did it/will it level out above or in
the vicinity of other Type Ib/c supernovae? At what point does SN 1998bw become like
ordinary supernovae? The current data enable us to make a prediction on the unification of
the GRB-SN phenomena, which can only be vindicated with further observations, obviously
of SN 1998bw, but also of the nearest XRFs and GRBs to fill in the missing parameter space.
We strongly encourage, therefore, followup observations of nearby GRBs and XRFs for as
long as the available instrumentation allows; we also encourage the calculation of off-axis
models of the X-ray light curves, especially for a variable distribution of Lorentz factors and
energies.
We are grateful to Re'em Sari for educational conversations regarding the nature of X-ray
afterglows. CK and SP acknowledge support from SAO grant GO1-2055X. SW is supported
by the NASA Theory Program (NAG5-12036), and the DOE Program for Scientific Discovery
through Advanced Computing (SciDAC; DE-FC02-01ER41176). ERR has been supported
by NASA through a Chandra Postdoctoral Fellowship award PF3-40028. Part of this work
was performed while CK, ERR and RB were visiting the University of CA in Santa Cruz.
The authors acknowledge support (RAMJW) and benefits from collaboration within the
Research Training Network "Gamma Ray Bursts: An Enigma and a Tool", funded by the
-- 18 --
EU under contract HPRN-CT-2002-00294.
REFERENCES
Berger, E., Kulkarni, S. R., & Frail, D. A. 2003, ApJ, 590, 379
Berger, E. et al., 2003, Nat 426, 154
Bjornsson, G., Hjorth, J., Pedersen, K., & Fynbo, J. U. 2002, ApJ, 579, L59
Bloom, J. et al. 1999, Nature 401, 453
Bloom, J. et al. 2003, astro-ph/0303514
Bregman, J. N., Houck, J. C., Chevalier, R. A., & Roberts, M. S. 2003, ApJ, 596, 323
Butler, N. et al. 2003, GCN #2347
Cash, W. 1979, ApJ, 228, 939
Chevalier, R. A. 1990, in Supernovae, ed. A. G. Petschek, Springer Verlag, p. 91
Chevalier, R. A. & Li, Z. 2000, ApJ, 536, 195
Cohen, E., Piran, T., & Sari, R. 1998, ApJ, 509, 717
Costa, E. et al. 1997, Nature 387, 783
Frail, D.A., et al. 2001, ApJ 562, L55
Fabbiano, G. 1989, ARA&A, 27, 87
Fabbiano, G. et al. 2003, ApJ, 584, L5
Frail, D. et al. 1997, Nature 389, 261
Freedman, W. L., Kennicutt, R. C., Mould, J. R., & Madore, B. F. 2003, A decade of
Hubble Space Telescope science. Proceedings of the Space Telescope Science Institute
Symposium, held in Baltimore, MD, USA, April 11-14, 2000, edited by Mario Livio,
Keith Noll, Massimo Stiavelli. Space Telescope Science Institute symposium series,
Vol. 14. Cambridge, UK: Cambridge University Press, ISBN 0-521-82459-1, 2003,
p. 214 - 221, 214
Galama, T. et al. 1998, Nature 395, 670
-- 19 --
Galama, T., et al. 2000, preprint
Giacconi, R. et al. 2001, ApJ, 551, 624
Granot, J., Panaitescu, A., Kumar, P., & Woosley, S. E. 2002, ApJ, 570, L61
Greiner, J., Morgan, E. H., & Remillard, R. A. 1996, ApJ, 473, L107
Greiner, J., Cuby, J.G. & McCaughrean, M.J. 2001, Nature, 414, 522
Harrison, F. A. et al. 2001, ApJ, 559, 123
Heise, J. 2003, American Institute of Physics Conference Series, 662, 229
Hjorth, J. et al. 2003, Nature, 423, 847
Humphrey, P. J., Fabbiano, G., Elvis, M., Church, M. J., & Ba luci´nska-Church, M. 2003,
MNRAS, 344, 134
Iwamoto, K., et al. 1998 Nature 395, 672
Kouveliotou, C. et al. 1993, ApJ, 413, L101
Kulkarni, S. et al. 1998, Nature 395, 663
Li, Z. & Chevalier, R. A. 1999, ApJ, 526, 716
MacFadyen, A. & Woosley, S. 1999, ApJ, 524, 262
Meszaros, P., Rees, M.J., Wijers, R.A.M.J., 1998, ApJ, 499, 301
Mirabel, I. F. & Rodriguez, L. F. 1994, Nature, 371, 46
Mori, K., Tsunemi, H., Miyata, E., Baluta, C. J., Burrows, D. N., Garmire, G. P., & Char-
tas, G. 2001, in ASP Conf. Ser. 251, New Century of X-Ray Astronomy, Yokohama
Symp. (San Francisco: ASP), 576
Nakamura, T. 1999, ApJ, 522, L101
Paczy´nski, B. 1998, ApJ, 494, L45
Pian, E. et al. 2000, ApJ, 536, 778
Pian, E. et al. 2003, astro-ph/0304521
Pooley, D. et al. 2002, ApJ, 572, 932
-- 20 --
Portegies Zwart, S.F., Pooley, D. & Lewin, W.H.G. 2002, ApJ, 574, 762
Prochaska, J. et al. 2003, GCN #2482
Ramirez-Ruiz, E., Celotti, A., Rees, M. J. 2002, MNRAS, 337, 1349
Rhoads, J. E. 1997, ApJ, 487, L1
Rodriguez-Pascual, P., Santos-Lleo, M., Gonzalez-Riestra, R., Schartel, N., & Altieri, B.
2003, GCN #2474
Rossi, E., Lazzati, D., Rees, M.J. 2002, MN 332, 945
Sako, M. & Harrison, F.A. 2002a, GCN #1624
Sako, M. & Harrison, F.A. 2002b, GCN #1716
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525
Soderberg, A.M. et al. 2003a, GCN #2483
Soderberg, A.M. et al. 2003b, astro-ph/0311050.
Soffitta, P. et al. 1998, IAUC 6884
Stanek, K. Z. et al. 2003, ApJ, 591, L17
Swartz, D. A., Ghosh, K. K., McCollough, M. L., Pannuti, T. G., Tennant, A. F., & Wu, K.
2003, ApJS, 144, 213
Swartz, D. A., Ghosh, K. K., Suleimanov, V., Tennant, A. F., & Wu, K. 2002, ApJ, 574, 382
Tsunemi, H., Mori, K., Miyata, E., Baluta, C. J., Burrows, D. N., Garmire, G. P., & Char-
tas, G. 2001, ApJ, 554, 496
Tiengo, A., Mereghetti, S., Ghisellini, G., Rossi, E., Ghirlanda, G., & Schartel, N. 2003,
A&A, 409, 983
Townsley, L. et al. 2002, APS Meeting Abstracts, 17061
Uno, S. et al. 2002, ApJ, 565, 419
Wieringa, M. H., Kulkarni, S. R., & Frail, D. A. 1999, A&AS, 138, 467
Woosley, S. E. 1993, ApJ, 405, 273
-- 21 --
Woosley, S., Eastman, R., & Schmidt, B. 1999, ApJ, 516, 788
Waxman, E., Kulkarni, S. R., & Frail, D. A. 1998, ApJ, 497, 288
Yamazaki, R., Yonetoku, D., & Nakamura, T. 2003, ApJ, 594, L79
Zezas, A., Fabbiano, G., Rots, A.H., & Murray, S. 2002, ApJ, 577, 710
Zhang, W., Woosley, S. E., & Heger, A. 2003, ApJ, in press, astro-ph/0308389
This preprint was prepared with the AAS LATEX macros v5.0.
-- 22 --
Fig. 1. -- (a): Digitized Sky Survey (DSS) image of the area around GRB 980425. The outer
(partial) circle represents the 8.0′ error region for GRB 980425 from the BeppoSAX WFC.
The full solid circles represent the 1.5′ BeppoSAX NFI S1 and S2 error regions. The dashed
circle (radius= 15′′) within S1 identifies the area covered by the GRB 980425 host galaxy
ESO 182 − G82. (b):CXO image of S1. The boxed region containing S1ab (lower left) and
S1c (upper right) is shown in (d). (c):CXO image of S2. Almost the entire error region (over
97%) was covered by the CXO CCDs S2 and S3. Source S2d falls into the gap between these
two CCDs. (d): CXO blow up of the square box in (b). The dashed circle (radius= 15′′)
indicates the host galaxy area (also identified in (a)). A close-up view of the solid box shown
here is seen in (e). (e):CXO blow up of the square box in (d). Sources S1a and S1b are
clearly resolved. S1a is the source to the south and coincides with the established radio
location of SN 1998bw. The two solid circles represent the 1σ error regions for these sources
with (r = 0.3′′) derived with comparative astrometry. (f):HST/STIS image of (e) taken in
the 50CCD/Clear mode (no filter). The white circles are the CXO error circles from (e).
-- 23 --
Fig. 2. -- Light curves (left column; binned in 3000 s wide bins) and spectra (right column;
binned in 0.232 keV wide bins) of the five sources with more than 20 counts total within
S1 and S2 collected with CXO (0.3 − 10.0 keV). Notice source S1b is highly variable during
the 50 ks observation. All other sources are either too faint to determine any variability or
consistent with a constant persistent emission.
-- 24 --
Fig. 3. -- Compilation of GRB, XRF, SN I and SN II X-ray light curves (0.3 − 10.0 keV)
presented as (isotropic) luminosity distances using a cosmology with ΩM = 0.27, ΩΛ = 0.73,
and H0 = 72 km s−1 Mpc−1. The XRF luminosities are calculated assuming two redshifts,
z = 1 (solid lines) and z = 0.251 (dashed lines). The solid long line corresponds to a temporal
decay of 1046 t(−1.69)
, discussed in the text.
days
-- 25 --
Fig. 4. -- The X-ray light curves of sources S1 and S2. The upper curve gives the two
BeppoSAX observations and upper limits for the error box of S2 during the first 200 days
after the GRB trigger. The last points (day 1281) on that curve are the sum (S2sum) of all
point sources observed with CXO and the sum (S2(3′)) of all sources observed with CXO
within an error circle of 3′ radius. A distance of z = 1.343 has been assumed for all S2 points
(see text). The arrow on day 1281 reflects the CXO detection limit of a source placed at
z = 1.343 or less. The lower curve is the light curve of source S1 assuming a distance equal to
that of ESO 184−G82 (35.6 Mpc). The first four points are the BeppoSAX observations; the
fifth (upper) point at day 1281 is the sum of all S1 CXO sources (S1sum). The point below
S1sum is the SN 1998bw luminosity after we subtract the flux contribution from sources S1b
and S1c from S1sum.
-- 26 --
Fig. 5. -- Isotropic equivalent luminosity of GRB X-ray afterglows, LX, scaled to t = 10
hours after the burst as a function of their isotropic equivalent γ−ray energy, Eiso. The data
are taken from Berger, Kulkarni & Frail (2003); here we have added data for GRBs 031203,
030329 and 980425, as indicated on the plot. The solid line is a fit including all data; the
dashed line is a fit exluding GRBs 980425, 031203, 000210, and 990705.
|
astro-ph/0406130 | 1 | 0406 | 2004-06-04T21:40:57 | NGC 604, the Scaled OB Association (SOBA) Prototype. I: Spatial Distribution of the Different Gas Phases and Attenuation by Dust | [
"astro-ph"
] | We have analyzed HST and ground-based data to characterize the different gas phases and their interaction with the MYC in NGC 604, a GHR in M33. The warm ionized gas is made out of two components: a high-excitation, high-surface brightness H II surface located at the faces of the molecular clouds directly exposed to the ionizing radiation of the central SOBA; and a low-excitation, low-surface brightness halo that extends to much larger distances from the ionizing stars. The cavities created by the winds and SN explosions are filled with X-ray-emitting coronal gas. The nebular lines emitted by the warm gas experience a variable attenuation as a consequence of the dust distribution, which is patchy in the plane of the sky and with clouds interspersed among emission-line sources in the same line of sight. The optical depth at H alpha as measured from the ratio of the thermal radio continuum to H alpha shows a very good correlation with the total CO (1-0) column, indicating that most of the dust resides in the cold molecular phase. The optical depth at H alpha as measured from the ratio of H alpha to H beta also correlates with the CO emission but not as strongly as in the previous case. We analyze the difference between those two measurements and we find that <=11% of the H II gas is hidden behind large-optical-depth molecular clouds. We detect two candidate compact H II regions embedded inside the molecular cloud; both are within short distance of WR/Of stars and one of them is located within 16 pc of a RSG. We estimate the age of the main stellar generation in NGC 604 to be approx. 3 Myr from the ionization structure of the H II region. The size of the main cavity is smaller than the one predicted by extrapolating from single-star wind-blown bubbles. | astro-ph | astro-ph |
To appear in the September 2004 issue of the Astronomical
Journal
NGC 604, the Scaled OB Association (SOBA) Prototype. I:
Spatial Distribution of the Different Gas Phases and Attenuation
by Dust
J. Ma´ız-Apell´aniz1,2, E. P´erez3, & J. M. Mas-Hesse4
ABSTRACT
We have analyzed HST and ground-based data to characterize the different
gas phases and their interaction with the Massive Young Cluster in NGC 604, a
Giant H II Region in M33. The warm ionized gas is made out of two components:
a high-excitation, high-surface brightness H II surface located at the faces of the
molecular clouds directly exposed to the ionizing radiation of the central Scaled
OB Association; and a low-excitation, low-surface brightness halo that extends
to much larger distances from the ionizing stars. The cavities created by the
winds and SN explosions are filled with X-ray-emitting coronal gas. The nebular
lines emitted by the warm gas experience a variable attenuation as a consequence
of the dust distribution, which is patchy in the plane of the sky and with clouds
interspersed among emission-line sources in the same line of sight. The optical
depth at Hα as measured from the ratio of the thermal radio continuum to Hα
shows a very good correlation with the total CO (1 → 0) column, indicating that
most of the dust resides in the cold molecular phase. The optical depth at Hα as
measured from the ratio of Hα to Hβ also correlates with the CO emission but
not as strongly as in the previous case. We analyze the difference between those
two measurements and we find that . 11% of the H II gas is hidden behind large-
optical-depth molecular clouds; we pinpoint the positions in NGC 604 where that
hidden gas is located. We detect two candidate compact H II regions embedded
1Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, U.S.A.
2Affiliated with the Space Telescope Division of the European Space Agency, ESTEC, Noordwijk, Nether-
lands.
2Instituto de Astrof´ısica de Andaluc´ıa (CSIC), P. O. Box 3004, 18080 Granada, Spain.
3Laboratorio de Astrof´ısica Espacial y F´ısica Fundamental - INTA, P. O. Box 50727, E-28080 Madrid,
Spain.
– 2 –
inside the molecular cloud; both are within short distance of WR/Of stars and
one of them is located within 16 pc of a RSG. We estimate the age of the main
stellar generation in NGC 604 to be ≈ 3 Myr from the ionization structure of the
H II region, a value consistent with previous age measurements. The size of the
main cavity is smaller than the one predicted by extrapolating from single-star
wind-blown bubbles; possible explanations for this effect are presented.
Subject headings: dust, extinction - HII regions - ISM: individual (NGC 604)
- ISM: structure - Galaxies: ISM
1.
INTRODUCTION
Giant H II Regions (GHRs) constitute one of the most conspicuous type of objects in
spiral and irregular galaxies at optical wavelengths. For example, the 30 Doradus nebula
can be detected with the naked eye at a distance of 50 kpc and five GHRs in M101 have
their own NGC number at a distance of 7.2 Mpc. GHRs are powered by Massive Young
Clusters (MYCs) that include 102 − 103 O and WR stars, as opposed to the few stars of this
type encountered in regular H II regions1. The advent of high-spatial resolution data has
revealed that GHRs have complex structures and cannot be accurately modeled as simple
Stromgren spheres centered on the ionizing cluster. The deposition of energy in the form of
ultraviolet light, stellar winds, and supernova explosions and the complex initial distribution
in mass and velocity of the ISM produce a spatial distribution of the warm ionized gas that
is anything but spherically symmetric. The analysis of the best studied GHR, 30 Doradus,
reveals that most of the optical nebular emission originates at the surface of a molecular
cloud in those points where it is directly exposed to the UV flux from the central cluster.
The molecular cloud is what remains from the original material that gave birth to the MYC
and the resulting region of intense optical nebular emission is a relatively thin layer instead
of a classical Stromgren sphere (Walborn et al. 2002; Rubio et al. 1998; Scowen et al. 1998).
This near-bidimensional character of an H II region is also observed in other objects, from the
low-mass Orion nebula (Ferland 2001 and references therein) to much more massive regions
such as N11 in the LMC (Barb´a et al. 2003). Another property shared by these objects is
that a second thin layer, the Photo Dissociation Region (PDR), where the non-ionizing UV
light transforms H2 into atomic hydrogen, can be found sandwiched between the H II region
1For example, the Trapezium cluster in the Orion nebula includes only one O star, θ1 Ori C; the rest of
the massive stars in the cluster are actually early-type B stars that provide only a small contribution to the
total number of ionizing photons (O'Dell 2001).
– 3 –
and the molecular cloud (Hollenbach & Tielens 1997).
The MYCs at the center of GHRs can be divided into two types: Super Star Clusters
(SSCs) are organized around a compact (1−3 pc) core while Scaled OB Associations (SOBAs)
lack such a structure and are more extended objects, with sizes larger than 10 pc (Ma´ız-
Apell´aniz 2001). The cores of some SSCs (e.g. 30 Doradus) are surrounded by extended
halos that are themselves similar to SOBAs in terms of structure and number of stars. This
distinction between SSCs and SOBAs is important for the long-term evolution of the cluster,
since the extended character of the latter makes them highly vulnerable to disruption by tidal
forces on time scales shorter than the age of the Universe. Therefore, only SSCs are likely to
survive to become the globular clusters of the future (Fall & Rees 1977). MYCs are expected
to form in a time scale shorter than 10 Myr but some of the best-studied examples, such as
30 Doradus and N11, reveal that their stars are not completely coeval: several million years
after the central regions have been formed, there is induced star-formation in their periphery,
leading to the so-called two-stage starbursts (Parker et al. 1992; Walborn & Parker 1992;
Walborn et al. 1999).
The analysis of GHRs in galaxies experiencing strong episodes of star formation has been
the basic tool for decades to understand the properties of starbursts and their interaction
with the surrounding medium. In order to characterize GHRs correctly, one has to establish
the balance between the ionizing UV flux produced by the stars and the flux of the Balmer
lines emitted by the GHR. The former can only be obtained from the analysis of the stellar
continua in the UV domain, where the extinction is strongest. The latter fluxes are also
affected by extinction, and it has been known for a long time that a simple correction using
the Hα/Hβ flux ratio might underestimate the intrinsic fluxes, due to the complex geometry
of the dust and the gas within the GHR (Caplan & Deharveng 1986). Furthermore, attempts
to study the extinction in starbursts using integrated data have discovered that, on average,
the color excess E(B − V ) experienced by the stars is about half that experienced by the
gas (Calzetti et al. 2000), therefore complicating even more the calculation of the balance
between the ionizing photons and the observed nebular spectrum, and making it very difficult
to derive the fraction of ionizing photons escaping from the GHR.
The main objective of this work is therefore to characterize the properties, distribu-
tion and interrelation between the stars, gas and dust components of GHRs, in order to
understand how they affect the resulting spectral energy distributions at various wavelength
ranges. To achieve this goal we started some years ago a program to analyze in detail nearby
resolved GHRs. In this work we present the first results obtained on NGC 604, concerning
the properties of its gas and dust phases.
NGC 604 is a Giant H II Region in M33, the third largest galaxy in the Local Group
– 4 –
after M31 and the Milky Way, located at a distance of 840 kpc (Freedman et al. 2001) along
a direction with low foreground extinction (Gonz´alez Delgado & P´erez 2000). 30 Doradus
and NGC 604 are the two largest GHRs in the Local Group. NGC 604 is powered by a MYC
without a central core that contains & 200 O and WR stars (Ma´ız-Apell´aniz 2001; Hunter
et al. 1996; Drissen et al. 1993). The MYC is located inside a cavity that has the kinematic
profile of an incomplete expanding bubble. On the other hand, the brightest knots all have
kinematic profiles that can be well characterized by single gaussians (Sabalisck et al. 1995;
Ma´ız-Apell´aniz 2000), a strong indication that the overall dynamics of the GHR is not yet
dominated by the kinetic energy deposition by stellar winds and SN explosions (Tenorio-
Tagle et al. 1996).
Given its proximity, size, low foreground extinction and structural/kinematical prop-
erties, NGC 604 constitutes an ideal object for our study. It can indeed be considered as
the prototype Scaled OB Association. We have combined HST and ground-based data to
characterize the Massive Young Cluster, the surrounding Interstellar Medium and the in-
teraction between them.
In this first article we present and discuss the properties of the
GHR created by the SOBA and its relationship with the molecular and coronal gas and dust
clouds in the region. We have used WFPC2/HST and ground-based data to characterize
the spatial distribution of the optical nebular emission properties. We have complemented
this information with published data at other wavelength ranges. In a subsequent paper we
will analyze the properties of the individual massive stars in NGC 604 and its extinction
using HST imaging and HST objective-prism spectroscopy, aiming to derive accurately the
properties of the individual stars and their evolutionary state. This will allow to compare
the ionizing flux being emitted in the region with the value derived in this work from the
nebular emission, constraining so the fraction of ionizing photons potentially escaping the
region without contributing to the ionization.
We present the observations and describe the data analysis in section 2. We describe
the attenuation by dust experienced by the warm ionized gas in section 3 and its density
and excitation structure in section 4. Finally, we discuss our results in section 5.
2. OBSERVATIONS AND DATA ANALYSIS
2.1. WHT Long-slit spectra
We obtained optical long-slit spectra of NGC 604 on the 18-19 August 1992 night with
the ISIS spectrograph of the William Herschel Telescope as a part of the GEFE collabora-
– 5 –
tion2. The spectra were obtained at ten different positions, all of them at a P.A. of 90◦, with
an effective width for each one of 1′′ and a spacing of 2′′ − 3′′ between the center of each two
consecutive positions (Fig. 1). Two spectra were taken simultaneously at each position, one
in the range between 6390 A and 6840 A (red arm) and the other one in the range between
4665 A and 5065 A (blue arm), both of them with a dispersion of approximately 0.4 A /
pixel. The slit was approximately 200′′ long, with a spatial sampling along the slit of 0.′′33525
/ pixel in the red arm and 0.′′3576 / pixel in the blue arm. The exposure times ranged be-
tween 900 and 1200 seconds and the airmass varied between 1.21 and 1.01. More details of
the data reduction and analysis are given in Ma´ız-Apell´aniz et al. (1998) and Ma´ız-Apell´aniz
(2000).
With this strategy the direction perpendicular to the slit was under-sampled. However,
taking into account that the seeing was comparable to the 1′′ − 2′′ not sampled between
subsequent slits, linear interpolation over the areas not covered is enough to reproduce the
spatial distribution of the different parameters. This approach allowed us to cover the
whole central region of NGC 604 in less than a single night without losing any significant
amount of information. The reconstructed Hα image (top panel in Fig. 9) shows indeed
all the features obtained with imaging techniques at similar resolution. These data have
been used for a variety of purposes in several previous works (Terlevich et al. 1996; Ma´ız-
Apell´aniz 2000; Tenorio-Tagle et al. 2000). Here we will use them to produce 2-D maps of
the [S II] λ6717/[S II] λ6731 ratio across the face of NGC 604 and to calibrate the nebular
WFPC2 data.
An additional set of spectroscopic data with the WHT were obtained by means of
a spatial scan of the spectrograph perpendicular to the slit length direction, which was
positioned at a PA=60◦. These scanned spectra were centered at RA = 1h 31m 43s, dec =
30◦ 31′52′′, and they cover the core of the region by displacing a 1′′-wide long-slit in steps of
1′′ and taking at each position a 1 minute exposure. The details of this observation and of
its calibration are given in Gonz´alez Delgado & P´erez (2000).
2GEFE (Grupo de Estudios de Formaci´on Estelar) is an international group whose main objective is the
understanding of the parameters that control massive star formation in starbursts. GEFE obtained 5% of
the total observing time of the telescopes of the Observatories of the Instituto de Astrof´ısica de Canarias.
This time is distributed by the Comit´e Cient´ıfico Internacional among international programs.
– 6 –
2.2. WFPC2 imaging
We list in Table 1 all the NGC 604 WFPC2 data currently available in the HST archive.
The data correspond to four different programs, each with a different field of view, but all
images include the region where the nebular line emission caused by the NGC 604 central
cluster can be clearly identified. The images available in each program can be summarized as
follows: 5237, optical broad-band (UV I) + Hα; 5384, FUV broad-band; 5773, "standard"
narrow-band ([O III] λ5007, Hα, and [S II] λ6717+6731) + wide Stromgren y; and 9134,
"extended" narrow-band ([O II] λ3726+3729, Hβ, [N II] λ6584, and [S III] λ9531). Only
some of those images will be used in this paper but the common part of the reduction
process is described here in order to use the data in a subsequent paper.
The standard WFPC2 pipeline process takes care of the basic data reduction (bias, dark,
flat field corrections). Cosmic rays were eliminated in each case using the crrej task and the
correction of the geometrical distortion and montage of the four WFPC2 fields was processed
using the wmosaic task, both of which are included in the standard STSDAS package for
WFPC2 data analysis. Hot pixels were eliminated using a custom-made IDL routine.
The next step in the data reduction, the registering of the images obtained under differ-
ent programs, followed a more elaborated process. The geometric distortion of the WFPC2
detectors is well known (Casertano & Wiggs 2001), allowing for a precise relative astrometry.
However, the absolute astrometry has two problems: First, the Guide Star Catalog, which is
used as a reference, has typical errors of ∼ 1′′ (Russell et al. 1990). Second, if only one point
in one of the four WFPC2 chips can be established as a precise reference using an external
catalog, the average error in the orientation induces an error of ∼ 0.′′03 in a typical position
in the other three chips. In our case, a star that is present in all 13 WFPC2 fields is included
in the Tycho-2 catalog as entry 2 293-642-1. We used the VizieR service (Ochsenbein et al.
2000) to obtain its coordinates and proper motion using the J2000 epoch and equinox and
we then corrected for the different epoch of the observations. The star is saturated (in some
filters only weakly, with only a few pixels affected, but in others quite heavily, with significant
bleeding to pixels in the same column) in all filters except F170W but we were able to fix
its centroid within one quarter of a WF pixel in all cases using the diffraction spikes. Thus,
the procedure followed to establish a uniform coordinate system had three steps. First, we
corrected for the difference in plate scale for each filter with respect to F555W using the
parameters provided by Dolphin (2000). Note that at the present time the STSDAS tasks
wmosaic and metric use the non-wavelength-dependent Holtzmann solution, which can in-
troduce ∼ 0.′′1 errors for FUV data; errors are an order of magnitude or less smaller if only
optical data are involved. Second, we rotated all mosaiced images in order to make them
share a common orientation. Third, using TYC 2 293-642-1 as a reference, we corrected for
– 7 –
the general displacement in RA and declination and compared the positions of the stars in
the central region of NGC 604 (at a distance of ∼ 70′) using images obtained under different
filters. As expected, coordinates differed by a few hundredths of an arcsecond, which we
take to be the precision of our absolute astrometry3.
In this first paper we are interested in obtaining "pure nebular" Hα, Hβ, [O III] λ5007,
[N II] λ6584 and [S II] λ6717+6731 images from the narrow-band F656N, F487N, F502N,
F658N, and F673N WFPC2 data, respectively. In order to do that, we used the broad-band
images to eliminate the continuum contribution. Following this approach one has to be
careful since broad-band images can be contaminated by the nebular emission itself (this is
especially so for F555W data). If one has to use a heavily-contaminated broad-band filter to
subtract the continuum, the full linear system of equations for the reciprocal contributions
has to be solved (see, e.g. MacKenty et al. 2000). Here we used F814W, F547M, and
F336W images, which have much weaker nebular contaminations and can be considered to a
first order approximation as free from contributions from emission lines. The continuum at
Hα, [N II] λ6584, and [S II] λ6717+6731 was interpolated from the F547M and F814W data
while that of Hβ and [O III] λ5007 was interpolated from the F336W and F547M data. In
principle, the continuum subtraction for the first three lines is expected to be more accurate
than for the last two for two reasons: (1) The difference in flux between F547M and F814W
is relatively small for the type of sources that contribute to the continuum (early-type stars
for the central cluster region, nebular continuum4 for the nebulosity, late-type stars for the
background) in comparison to the difference in flux between F336W and F547M for some of
those types. (2) Some of those source types have strong Balmer jumps in their spectra, which
makes interpolation between F336W and F547M more inaccurate. With those caveats in
mind, we tested several interpolation mechanisms and the best results (for both point sources
and background) were obtained using a double (in λ and in flux) logarithmic scale, which
is exact if the flux follows a power law in λ. We should point out that, even though the
remaining flux due to the continuum contribution at the location of a bright star in the final
nebular image typically averages to zero (i.e. the overall continuum subtraction is correct),
there is usually some structure visible at such locations in the resulting image. The main
reason for this effect is the undersampled nature of the WFPC2 PSF, which does not permit
3Recently, Anderson & King (2003) and Kozhurina-Platais et al. (2003) have attacked the geometric
distortion solution of the WFPC2 with success, improving its accuracy significantly, but we do not use those
results here since the precision we attain is sufficient for our needs.
4Note that, in any case, the nebular contribution is rather unimportant and in most cases of the same or
smaller magnitude as the uncertainties derived from the photometric calibration.
– 8 –
a perfect subtraction of one image from another5. For that reason, we decided to blank
out in the nebular images those areas with bright stars and when measuring the integrated
nebular fluxes we have interpolated in those regions using the neighboring pixels. Another
issue related to the obtention of pure nebular images, the mutual contamination between
Hα and [N II] λ6548+6584 in the F656N and F658N filters is discussed later.
Another topic that needs to be addressed here is that of the absolute photometric
calibration of the WFPC2 nebular filters. It is stated in the HST Data Handbook (STScI
2002) that their accuracy is estimated to be . 5%. A recent study by Rubin et al. (2002)
found this to be true for three of the filters used here, F487N, F502N, and F656N (see
Table 2). We performed an independent calibration of four of the nebular filters (F487N,
F656N, F658N, and F673N) using the NGC 604 Hα and Hβ images published by Bosch
et al. (2002), which were kindly made available to us by Guillermo Bosch, and our WHT
long-slit spectra. For each filter we used synphot (STScI 1999) to obtain the conversion
factors between detector and physical units both for a continuum source and for an emission
line with the appropriate blueshift corresponding to NGC 604.
1. For F487N, we compared the continuum-subtracted integrated flux over a 50′′ × 50′′
region (after smoothing the image with a Moffat kernel to degrade its resolution) with
the same value derived from the Bosch et al. (2002) data.
2. For F656N, we used a similar but more complex procedure. The photons detected in
that filter can originate in the continuum, Hα itself, or one of the neighbor [N ii] lines.
Given that the expected ratios are rather low ([N II] λ6584/Hα∼ 0.1, [N II] λ6548/Hα∼
0.03) and that the throughput peaks near the wavelength of Hα, the contamination
by [N ii] is expected to be not too important (< 10%) but still significant. On the
other hand, F658N is also contaminated by Hα emission to a similar degree (Hα falls
at the tail of the filter throughput at the blueshift of NGC 604). Therefore, we de-
cided to eliminate the mutual contamination by establishing the 2 linear equations with
the coefficients determined from synphot and solving for the two unknowns, Hα and
[N II] λ6584, using [N II] λ6584/[N II] λ6548 = 3. We then compared the obtained Hα
flux with the Bosch et al. (2002) data in a manner analogous to what we did previously
for Hβ.
3. To verify the calibration for F658N and F673N we first obtained the continuum-
subtracted images (and for the case of F658N we also eliminated the Hα contribution
5There are, however, at least two more effects, detector saturation and emission associated with the star
itself, which are also important in a number of cases.
– 9 –
as detailed in the previous point) and degraded their resolution with a Moffat kernel.
We then calibrated the Hα fluxes from the long-slit spectra with the Hα image from
Bosch et al. (2002) after shifting the relative positions of the long-slits to match the
structures seen in the corresponding cuts of the images. We used the absolute cali-
bration thus provided to obtain the corresponding [N II] λ6584 and [S II] λ6717+6731
fluxes for the long-slit spectra and then compared the integrated fluxes with those
obtained from the WFPC2 F658N and F673N data.
Given the mutual influence in the calibration of F656N and F658N, the above procedure
was iterated several times until convergence was reached. The results are shown in Table 2 in
the form of the ratio between the values of the EMFLX parameter (the factor used to convert
from ADU/s to monochromatic flux) as provided by synphot and as measured here. For
all four cases we detect that synphot slightly overestimates the throughput of the nebular
filters, which is the same effect that was detected by Rubin et al. (2002) for two of them,
F487N and F656N (they did not analyze F658N and F673N). Our values are similar but not
identical, which is not unexpected given that the different radial velocities of the objects
studied is sufficient to shift the wavelengths of the emission lines by several Angstroms. In
the rest of this paper we will use the calibration discussed here for F487N, F656N, F658N,
and F673N and the one obtained by Rubin et al. (2002) for F502N.
It should be noted that Charge Transfer Efficiency (CTE) effects are not included in
the results of Table 2, either for the new values or for those extracted from Rubin et al.
(2002), and that could be a major factor in the existing differences6 Since our data have
been directly calibrated using ground-based fluxes (except for F502N) this should not be a
problem for the results in this paper. However, one should be careful when using the results
in Table 2 with other data and, if possible, should attempt a similar procedure as the one
outlined in the previous paragraphs7.
6A sign pointing in that direction is that the largest correction is found for F658N; those exposures have
lower backgrounds and were obtained later than F656N or F673N, both of which have smaller corrections.
This is what would be expected if CTE effects were playing a significant role.
7In a recent paper, Calzetti et al. (2004) did a similar analysis for four different targets. They found
object-to-object variations in the calibration, likely caused by differences in redshift, among other effects.
See also the work by O'Dell & Doi (1999) for a calibration of the nebular filters applicable to resolved
Galactic H II regions.
– 10 –
Table 1. NGC 604 WFPC2 data available in the HST archive.
Prog.
P.I.
Filter
Band
Data sets
Exp. times (s)
5384 Waller
5773
Hester
WFPC2 U
WFPC2 V
WFPC2 I
5237 Westphal F336W
F555W
F814W
F656N
F170W
F502N
F656N
F673N [S II] λ6717+6731
F547M Wide Stromgren y
F375N [O II] λ3726+3729
9134
Garnett
Hα
Far UV
[O III] λ5007
Hα
u2ab0207t + 208t
u2ab0201t + 202t
u2ab0203t + 204t
u2ab0205t + 206t
u2c60b01t + 202t
u2lx0305t + 306t
u2lx0301t + 302t
u2lx0303t + 304t
u2lx0307t + 308t
u6cj0201m + 202r +
203m + 301m
600 + 600
200 + 200
200 + 200
1 000 + 1 000
350 + 350
1 100 + 1 100
1 100 + 1 100
1 100 + 1 100
500 + 500
2 700 + 2 700 +
2 700 + 2 700
F487N
F658N
F953N
Hβ
u6cj0101m + 102m + 103m
2 700 + 2 700 + 2 700
[N II] λ6584
[S III] λ9531
u6cj0104m + 105m
1 300 + 1 300
u6cj0401m + 402m + 403m +
404m + 405m + 406m
1 300 + 1 300 + 1 300 +
1 300 + 1 300 + 1 300
Table 2. Absolute flux calibration of the nebular WFPC2 filters. The quantity plotted in
each case is the EMFLX parameter (conversion factor between count rate and
monochromatic line flux) provided by synphot (STScI 1999) divided by the value of
EMFLX measured by alternative calibrations. A value greater than one indicates that
synphot overstimates the throughput and, therefore, that the sense of the correction is to
increase the line flux (since in that case a larger real count rate is necessary to produce a
given monochromatic flux).
Data
F487N F502N F656N F658N F673N
1.000
synphot calibration
1.032
Table 1 of Rubin et al. (2002)
Bosch et al. (2002) data
1.012
Bosch et al. (2002) + long-slits -–
1.000
0.959
-–
-–
1.000
1.037
1.101
-–
1.000
-–
-–
1.210
1.000
-–
-–
1.114
– 11 –
2.3. Other data
In order to provide a more complete picture of the ISM in NGC 604 we obtained some
additional data from the literature and other archives. Ed Churchwell generously provided
us with the data shown in Fig. 2 of Churchwell & Goss (1999). Those authors obtained a 3.6
cm radio continuum map of NGC 604 with the VLA and combined the data with a ground-
based Hα image of the region to produce a map of τrad, the (true) optical depth experienced
by Hα emission (see Appendix). Here we will use the τrad map to compare it with a τBal
map, that is, a map of the optical depth at Hα as measured from the ratio of the fluxes of
the Hα and Hβ emission lines. A vital step to do an study of the attenuation by dust is
to obtain a correct registration between the images obtained at different wavelengths. In
order to do this we first degraded the spatial resolution of our WFPC2 Hα and Hβ images
in order to match that of the radio data. As it is discussed later in the paper, some of
the bright H II knots are located on extinction gradients and this can cause a significant
displacement between the positions of the intensity peaks in Hα and radio, rendering those
knots useless for registration purposes. Therefore, we decided to use instead for registration
(a) the SNR present in NGC 604 (knot E in Churchwell & Goss 1999), which is located in a
relatively dust-free area and should have the same coordinates in both the optical and radio
data (D'Odorico et al. 1980; Gordon et al. 1998); and (b) the intensity minima determined
by the cavities in the ionized gas, since those structures are also located on areas with low
attenuation. This process yielded a registration correct to better than 0.′′5 without requiring
the use of a rotation between the optical and radio data, as estimated from the residuals
in the fit. Given that the radio beam had a HPBW of 4.′′18 × 3.′′63, the precision in the
registering is good enough for our purposes.
We also extracted from the M33 CO (1 → 0) survey of Engargiola et al. (2003) a map of
the NGC 604 region. That survey was obtained with BIMA and has a spatial resolution of
13′′. Finally, we also retrieved from the Chandra archive a 90 ks image of NGC 604 obtained
with ACIS (P.I.: Damiani). The X-ray data were registered to the optical data using the
SNR in NGC 6048. Here we will use the CO and X-ray data to compare the morphologies
of the different phases of the ISM in NGC 604.
8The registering of the CO and X-ray data is not as vital as that of the radio continuum, since in those
cases we do not calculate intensity ratios using data from the different wavelength ranges.
– 12 –
3. NEBULAR MORPHOLOGY AND ATTENUATION BY DUST
3.1. Nebular morphology
The nebular structure of NGC 604 consists of a central bright, high-excitation region sur-
rounded by a dimmer low-excitation halo. The central region is detected in all bright nebular
lines while the halo can be easily seen only in low-excitation lines such as [S II] λ6717+6731
or [N II] λ6584 (see Fig. 2). The high excitation region is composed of a 16′′ × 14′′ shell sur-
rounding a central cavity (which we will call A following the nomenclature of Ma´ız-Apell´aniz
2000) centered on coordinates (90′′,9′′)9; two filaments that extend N/S along x = 100′′ and
x = 104′′, respectively; and a filled H II region centered at (76′′,−5′′).
The SOBA has two loosely-defined components, a main one centered at cavity A and
elongated in the N-S direction (from approximately (92′′,2′′) to (85′′,18′′), see Fig. 3) and a
secondary one, more dispersed and approximately cospatial with the eastern part of the shell
around cavity A and the two high-excitation filaments located towards the E (in the range
x = 92′′ − 103′′ and y = 2′′ − 25′′, see Fig. 3). Another more distant group of bright stars can
be seen at the location of the filled H II region towards the NW (this group is called cluster
B in Hunter et al. 1996).
The shell around cavity A is far from uniform. Towards the SW it is very bright and
very thin (likely unresolved) in [O III] λ5007. Towards the E it is of intermediate brightness,
much thicker and less well defined, with a bright intrusion extending towards the center of
the cavity at (90′′, 12′′) (Tenorio-Tagle et al. 2000). Towards the N it is thin and well defined
again but dimmer. The two eastern filaments are also not uniform, with the southern one
brighter than the northern one. We want to stress that the optical appearance of all the
high excitation regions is consistent with them being thin structures (thicknesses of 1 − 2
pc), with an apparent variable thickness created by different inclinations with respect to the
observer. As described in the introduction, this is the observed geometry for well-studied
galactic and extragalactic H II regions. This effect is seen, for example, in that the areas with
the highest surface brightness appear to be one-dimensional, as expected from orientation
effects when the geometry is such that we observe the surface edge on. Later on, we describe
other observational evidence that support this model for the structure of the H II gas in
NGC 604.
Tenorio-Tagle et al. (2000) explored the kinematics of the H II gas in NGC 604 and
found that the shell around cavity A can be characterized by a single gaussian profile with
9See Fig. 9 for an explanation of the coordinate system.
– 13 –
an approximately constant velocity of −255 km s−1. The H II gas on cavity A itself needs
to be described by at least two kinematic components, one shifted towards the red and one
towards the blue. The red component is present everywhere and shows the spatial profile
of an expanding bubble while the blue one has a patchy coverage. The bright intrusion at
(90′′, 12′′) is part of the blue-shifted component. Tenorio-Tagle et al. (2000) interpreted this
kinematic profile as a sign that the superbubble originally associated with cavity A had burst
in the direction towards us.
Two other cavities are visible in the nebular-line images towards the S and N; we will call
them B and D, respectively, following the previously established nomenclature (see Fig. 2).
They contain only a few massive stars each. Their kinematics were partially mapped by
Tenorio-Tagle et al. (2000) and they were found to be similar to that of cavity A: the walls
have single gaussian components with approximately constant velocity while the cavity itself
shows line-splitting indicative of expansion. Given the low number of massive stars these
two structures contain and that they are apparently connected to the main cavity (there are
regions at the boundaries where only weak low-excitation nebular emission can be seen), a
likely origin for these two structures is that they were also produced by the superbubble in
cavity A bursting out into the surrounding medium (in this case along the plane of the sky
rather than along the line of sight).
The region towards the E of the x = 100′′ − 104′′ filaments is marked by a sharp drop
in the excitation ratios (see Fig. 12). A fourth cavity (C) is centered on (119′′,5′′). It differs
from the previous three in that the intensity of the nebular lines in its central region is
weaker and does not show line splitting. Only a few stars are present inside and, judging
from their magnitudes, they are likely to be of late-O or early-B type at most. All of this
suggests that the fourth cavity is an older structure which has been recently re-ionized by
the current burst.
The distribution of the CO (1 → 0) emission in NGC 604 with respect to the H II region
is shown in Fig. 4. Two clouds can be seen close to the SOBA. The largest one is centered
on (102′′,27′′) while the second one is centered on (80′′,17′′). Farther away, a third cloud is
centered on (113′′,51′′) and a fourth, much less intense, on (82′′,-23′′). Looking at Fig. 4 and
allowing for the difference in spatial resolution between the two data sets, it can be seen that
all the bright areas of the H II region are (a) located at the edges of the first two molecular
clouds and (b) oriented towards the ionizing sources10. As we have already mentioned, this
is the same configuration observed in Galactic H II regions and in more massive objects
such as 30 Doradus and N11 and provides further evidence towards the near-2D character
10That is, they are on the edges of the molecular clouds that point towards the massive stars in the SOBA.
– 14 –
of these areas. Therefore, it appears to be a general configuration that the bright areas of
H II regions (excepting maybe the oldest and the youngest ones) originate in the surface of
molecular clouds directly exposed to massive young stars.
Diffuse X-rays are observed filling the four cavities in NGC 604 (see Fig. 5), with the
highest intensity originating in the main one. Two point sources are also detected: one is the
SNR and the other one, located at (137′′,0′′), is not associated with any of the other sources
described in this paper. The integrated X-ray emission from the diffuse gas was modeled
by Yang et al. (1996). The X-ray emission is very soft, with a best-fitting Raymond-Smith
model with a plasma temperature of 1.3 × 106 K (kT = 0.12 keV), giving a luminosity in
the 0.5–2 keV range of around 8 × 1038 erg s−1. These authors concluded that the X-ray
emission is most likely dominated by the contribution from a hot, thin plasma energized by
stellar winds. Cervino et al. (2002) predict a soft X-ray emission of around 1034 erg s−1 M−1
for a MYC of 3 Myr, assuming average conditions. For a total stellar mass of around 105 M⊙
in the range 2–120 M⊙ (Yang et al. 1996), we conclude that the observed X-ray luminosity
is indeed consistent with a hot plasma energized by a young massive cluster. As discussed
by Cervino et al. (2002), the contribution by individual stars of any kind at this age should
be only around 1030 erg s−1 M−1
⊙ , which is negligible in comparison with the contribution of
the coronal gas.
⊙
3.2. Attenuation data
We have smoothed our continuum-subtracted Hα and Hβ images with a 5 × 5 pixel
box, calculated their ratio, and applied Eq. 6 to obtain a map of τBal, the optical depth at
Hα measured from the ratio of the two Balmer lines. The result is shown in Fig. 6, where
we have masked out bright stars (as previously described) and areas with low values of the
S/N ratio. We stress again the need for an accurate registering and flux calibration for such
a map to be a realistic representation of τBal, topics that we have covered in the previous
section. Other aspects that require some words of caution regarding Fig. 6 are also detailed
here:
• An incorrect temperature can induce systematic differences in the value of τBal (see
Eq. 5 in the Appendix). For NGC 604 we use a value of 8 500 K based on the mea-
surements of 8 150 ± 150 K for Te([O III]) and 8 600 ± 450 K for Te([N II]) of Esteban
et al. (2002) (see also D´ıaz et al. 1987; Gonz´alez Delgado & P´erez 2000). However,
the temperature dependence is rather weak: a change from 8 500 K to 10 000 K (which
is too large to be realistic for NGC 604, see D´ıaz et al. 1987; Esteban et al. 2002)
changes τBal only from 0.245 to 0.273 for Fα/Fβ = 3.2. The dependence is stronger
– 15 –
for the case of τrad (see Eq. 2) but the uncertainties introduced by a possible error in
the temperature are not large. A change from 8 500 K to 10 000 K would change the
overall value of τBal for NGC 604 only from 0.54 to 0.63.
• The existence of scattered Balmer radiation from a dust cloud can yield localized values
of Fα/Fβ lower than the canonical value corresponding to τBal = 0. Such a phenomenon
has been observed in e.g. NGC 4214 (Ma´ız-Apell´aniz et al. 1998) and it is also present
here in a few pixels in the central cavity. In those cases we have simply used τBal = 0.
Our τBal map provides us with a 0.′′5 (2 pc) resolution map of the attenuation experienced
by the H II gas in NGC 604. The problem of recovering the flux lost by geometrical effects
in the dust distribution (as opposed to the flux lost by the total amount of dust present
between the source and the observer) can be divided into two parts: (a) the patchiness of
the dust geometry in the plane of the sky and (b) the possibility that some of the dust may be
interspersed among sources along the same line of sight. (a) has the effect of underestimating
the real Hα flux. If we have observations with a spatial resolution comparable to the scales
in which the dust distribution varies, then we can correct for the effects of dust by applying
an extinction correction pixel by pixel. This is the main motivation for producing a τBal map
with the highest possible spatial resolution.
The solution to (b) is more complicated. Suppose we have a thick dust cloud located
between sources A and B, both in the same line of sight with A closer to the observer. In that
case, most or all of the Balmer photons emitted from B will be absorbed in the cloud and our
value of τBal will be only a measurement of the possible foreground extinction experienced
by A. From the point of view of Hα and Hβ the existence of B will be unknown to us and if
we were to estimate the ionizing flux from that information, we would miss the contribution
from B completely. In a general case, it can be shown that the effect of (b) is the same as
that of (a): to introduce an underestimation of the total ionizing flux. What can we do to
recover the lost flux? One solution is to use the thermal radio continuum, which is unaffected
by dust and can give us a better estimate of the ionizing flux. Nevertheless, there are two
considerations that affect radio observations:
• At the present time, radio observations of low-to-medium emission-measure thermal
sources such as extragalactic H II regions have lower spatial resolution than that avail-
able from optical observations. The VLA can reach HST-like resolution but only for
high emission-measure sources.
• It is possible to have contamination from non-thermal continuum radio sources such
as SNRs. This can be mitigated by observing at multiple wavelengths but such a
procedure can introduce large uncertainties if the non-thermal contribution is large.
– 16 –
In addition, a fraction of the ionizing photons might simply be destroyed by dust and/or
might escape the region without contributing to the ionization process. In our next paper
we will characterize the massive stellar population and will compare the predicted with the
measured fluxes, aiming to constrain the fraction of escaping photons.
In order to explore the relative importance of the non-uniform dust distribution in the
plane of the sky and along the line of sight we have produced an ≈ 4′′ resolution map of τrad,
the real optical depth at Hα as measured from the ratio of the thermal radio continuum to
Hα (see Eq. 3 in the Appendix). The map is shown in Fig. 7 along with the original radio
image, the resolution-degraded WFPC2 Hα image used to calculate τrad, and a τBal map at
the same resolution and scale as the τrad one. We have also integrated the fluxes along the
polygonal regions shown in that same figure, which correspond to the main H II knots11,
and have computed the corresponding values of τrad and τBal (see Table 3). In that same
table we include three other quantities. The first one, τ ′
Bal, is the value of τBal calculated by
correcting for extinction pixel by pixel in the region using the full-resolution WFPC2 images,
i.e. the weighted value we recover by using the spatial information in the plane of the sky up
to the resolution provided by WFPC2. One expects that τBal < τ ′
Bal does
not correct for structures present in the dust at smaller spatial scales and along the line of
sight, and that is indeed the case for all entries in Table 3. The other two quantities are the
values of τcov and γ corresponding to τrad and τBal applying the patchy foreground model of
the Appendix and using the one-to-one mapping that can be derived from Fig. 8.
Bal < τrad, since τ ′
In order to make better use of the high-resolution optical data, we have subdivided the
A to H regions into sub-regions and calculated the corresponding values of τBal and τ ′
Bal. The
results are shown in Table 4, where the entry "diffuse" refers to the totality of the integrated
area except the previous line, which is the sum of all sub-regions. Note that the value of
τ ′
Bal for the "diffuse" region is likely to be an overestimation of the real value caused by the
low S/N ratio of the Balmer ratio for the outer regions of NGC 604 because the logarithm
in Eq. 6 introduces a bias by giving more weight to the pixels where Fβ is close to zero due
to noise fluctuations.
We end this subsection by pointing out that the values listed in Tables 3 and 4 for the
entry "NGC 604" refer to the area (a) where nebulosity can be easily detected in Hα (b)
that is covered by all the WFPC2 exposures in programs 5237, 5773, and 9134. In order to
check that we were not leaving out any outlying regions with a significant flux contribution,
11The notation follows Churchwell & Goss (1999) joining C and D due to their proximity and including
G and H as two new knots. We also include the sum: (a) of the 7 regions, (b) of all the integrated area, and
(c) of all the integrated area without knot E (the SNR).
– 17 –
Table 3.
Integrated values for A to H regions.
Apert.
Areaa
Frad
b
c
Fα
Frad/Fα
d
e
τrad
c
Fβ
Fα/Fβ
e
τBal
f
τ ′
Bal
τcov
g
γg
A
B
CD
E
F
G
H
A-H
NGC 604
NGC 604 (no E)
48.80
116.39
347.47
93.96
349.04
208.20
161.82
1325.68
6120.94
6026.98
3.62
6.55
17.92
0.67
4.82
6.90
3.29
43.78
56.77
56.10
13.73
20.96
87.71
3.48
29.35
47.22
11.09
213.55
314.60
311.12
0.264
0.313
0.204
0.192
0.164
0.146
0.297
0.205
0.180
0.180
0.92
1.09
0.67
0.61
0.45
0.33
1.04
0.67
0.54
0.54
4.12
6.24
28.13
1.07
9.05
15.71
2.92
67.25
99.17
98.10
3.330
3.357
3.118
3.258
3.244
3.006
3.798
3.176
3.172
3.172
0.34
0.36
0.18
0.29
0.28
0.09
0.66
0.23
0.22
0.22
0.47
0.45
0.25
0.51
0.40
0.21
0.80
0.34
0.42
0.41
2.02
2.27
2.17
1.48
1.03
1.89
1.52
1.92
1.59
1.59
0.695
0.741
0.549
0.588
0.561
0.333
0.827
0.572
0.525
0.525
aIn square arcseconds.
bIn mJy.
cIn 10−13 erg s−1 cm−2.
dIn 1013 mJy erg−1 s cm2.
eCalculated from previous column (low resolution data).
f Calculated by correcting for extinction pixel by pixel in the high resolution data.
gCalculated from τrad and τBal, see Appendix and Fig. 8.
Note. - The two largest uncertainty sources for the optical depths are the absolute flux calibration and the assumed
temperature. See the discussion in subsections 2.2 (see Table 2) and 3.2, respectively.
– 18 –
Table 4.
Integrated values for A1 to H2 subregions.
Apert.
Areaa
b
Fα
Fα/Fβ
c
τBal
d
τ ′
Bal
ratio 1e
ratio 2f
ratio 3g
3.203
3.757
3.253
3.712
2.996
3.103
3.145
3.307
3.108
3.414
3.209
3.500
2.877
3.002
2.945
3.598
3.510
3.925
3.175
3.168
3.173
3.172
0.25
0.63
0.28
0.61
0.09
0.17
0.20
0.32
0.17
0.40
0.25
0.46
0.00
0.09
0.04
0.53
0.47
0.74
0.23
0.22
0.22
0.22
0.33
0.73
0.36
0.69
0.25
0.23
0.22
0.38
0.24
0.57
0.31
0.81
0.21
0.16
0.21
0.68
0.56
0.87
0.34
0.56
0.42
0.42
2.63
2.42
2.07
1.74
1.75
2.49
2.10
1.78
1.77
1.48
1.82
1.50
1.93
1.70
1.15
1.27
1.16
1.48
1.90
1.32
1.71
1.71
0.094
0.159
0.121
0.196
0.136
0.088
0.105
0.149
0.155
0.553
0.156
0.240
0.118
0.122
0.205
0.229
0.245
0.234
0.141
0.266
0.182
0.181
0.090
0.135
0.104
0.151
0.117
0.081
0.099
0.132
0.130
0.218
0.121
0.170
0.102
0.106
0.155
0.172
0.166
0.163
0.116
0.194
0.142
0.142
A1
A2
B1
B2
CD1
CD2
CD3
CD4
CD5
E1
F1
F2
G1
G2
G3
G4
H1
H2
A1-H2
diffuse
NGC 604
NGC 604 (no E1)
22.51
26.26
73.13
43.26
69.39
44.07
55.82
43.94
129.45
20.91
230.97
116.01
29.65
90.78
63.25
25.17
36.67
124.68
1245.92
4875.02
6120.94
6100.03
9.88
4.25
16.16
5.10
14.03
19.80
28.77
9.83
14.38
0.87
25.32
3.84
7.21
29.32
9.48
2.15
2.77
8.54
211.70
102.90
314.60
313.73
aIn square arcseconds.
bIn 10−13 erg s−1 cm−2.
cCalculated from previous column.
dCalculated by correcting for extinction pixel by pixel.
e[O iii] λ4959/Hβ.
f [S ii] λ6717+6731/Hα.
g[N ii] λ6584/Hα.
Note. - The two largest uncertainty sources for the optical depths are the absolute flux calibration
and the assumed temperature. See the discussion in subsections 2.2 (see Table 2) and 3.2, respectively
– 19 –
we compared our total value for the Hα flux with the one measured by Bosch et al. (2002)
using a larger area. Their value, (3.1 ± 0.4) · 10−11 erg cm−2 s−1, is in excellent agreement
with ours.
3.3. Results
Four attenuation maxima are present in the τrad map of Fig. 7, three in the upper half
of the map centered around (104′′,24′′), (80′′,15′′), and (93′′,35′′), respectively, and another
one more extended near the bottom of the map. The third maximum is not really a high
attenuation region:
it is the previously mentioned SNR, for which Eq. 3 does not provide
an accurate measurement of τrad given the non-thermal character of the radio flux, and will
not be discussed any further here. The other three, however, coincide with the positions of
three of the CO (1 → 0) maxima in Fig. 4. In general, once we account for the SNR and
the difference in resolution, there is a very good correlation between τrad and CO (1 → 0)
intensity.
If we try the same comparison using the τBal map instead of τrad, we find that the
correlation is also present but that it is not as strong. The main culprit is the attenuation
maximum at (80′′,15′′), which is much weaker in τBal than in τrad. It is also clear from Fig. 7
that τrad ≥ τBal everywhere (except for a few weak areas where resolution and scattering
effects may be important), as expected from the discussion in the previous subsection.
The first attenuation maximum corresponds to the main molecular cloud of NGC 604
(cloud 1 of Viallefond et al. 1992, cloud 2 of Wilson & Scoville 1992) and extends over region
H and part of regions A, B, and G (more specifically, over sub-regions A2, B2, G4, H1,
and H2). The right panel of Fig. 10 shows the correlation between τrad and CO(1 → 0)
intensity in this region. The brightest regions in Hα are located in A and B, which span
the SE bright filament. A strong E-W attenuation gradient is visible on those regions, as
it can be seen comparing A1 with A2 and B1 with B2 (also G2 with G4) in Fig. 6 and in
the cuts in Fig. 11 (note how in the first plot the attenuation gradient coincides with the
Hα intensity maximum at the center of region A). The existence of this gradient could in
principle justify the fact that τBal ≈ τrad/3 for both A and B by indicating that a patchy
foreground screen is responsible for the difference. However, τ ′
Bal/τrad is 0.51 for A and 0.41
for B, so either the patchiness extends to scales below 2 pc or some Balmer emission is
completely hidden from view. In this context, it is interesting to note that the Hα and radio
peaks are nearly coincident for A but that the radio peak is displaced towards the SE by ≈ 1′′
with respect to the Hα peak for B, the region of the two with the lowest value of τ ′
Bal/τrad.
This suggests that the H II gas is located along the surface of the main molecular cloud and
– 20 –
that this cloud creates a high-obscuration "flap" (seen in Fig. 3 running from (101′′,21′′) to
(100′′,27′′)) that absorbs most or all of the Balmer photons located behind, letting only the
radio continuum photons traverse it. The interpretation of that intensity drop as such a flap
would be consistent with the observed values of the optical depth and, as we will see later,
may not be the only such structure in NGC 604.
Region A, on the other hand, appears to be dominated by a single, compact, barely-
resolved H II region. Our measured value of τ ′
Bal/τrad is more likely to be explained by our
lack of better spatial resolution. Farther towards the E into the main molecular cloud, region
H shows high attenuation over most of its extension but does not include very bright areas.
Its values for τrad and τ ′
Bal are quite similar, an indication that a simple patchy foreground
screen provides an accurate description of the H II region without the need to include much
highly-obscured H II gas or unresolved dust clouds.
The second attenuation maximum corresponds to a molecular cloud that was outside
the region studied in CO by Viallefond et al. (1992) and Wilson & Scoville (1992) but that
is well defined in the Engargiola et al. (2003) map. This region also corresponds to the SW
quadrant of the main cavity, where the high-excitation shell is brightest, and is covered by
our low-resolution region CD. We note the following:
• τBal, τ ′
Bal, and τrad here are lower than in regions A, B, or H, indicating an overall lower
importance of attenuation.
• τ ′
Bal/τrad = 0.39, a value similar to that of region B, suggesting the existence of hidden
H II gas.
• The values of τBal (≈ 0.2) for the sub-regions CD2 and CD3 that correspond to the
brightest regions (the high excitation shell) are almost identical to the overall value of
τBal for the whole CD region.
• A sharp intensity drop in all emission lines is seen between sub-regions CD2 and CD3
and subregion CD4 (sub-regions were specifically chosen to follow this boundary), as
it can be seen in Figs. 3 and 11.
All of the above point towards the existence of a more or less uniform foreground ex-
tinction with τBal ≈ 0.2 combined with another "flap" that is covering the central part of
the near-edge-on high-excitation region, almost dividing it into two to the point of making it
appear as two separate knots (C and D) in the low-resolution Hα data (Churchwell & Goss
1999). Can we test this hypothesis any further? Yes, there are three more pieces of evidence
that point in the same direction. First, the radio peaks are displaced by ≈ 2′′ towards the
– 21 –
SW with respect to the low-resolution Hα peaks, which is exactly what would be expected
if the flap detectable in the high-resolution Hα images was occulting a flux similar to that in
sub-regions CD2 and CD3. Second, the value of τBal for subregion CD4 is indeed higher than
that for CD2 and CD3 but nowhere near the value required to raise the total value of τBal
for all of CD to 0.67. This is also seen in Fig. 11, where the sharp drop in Hα intensity as
one moves towards the left (W) is accompanied only by a slow increase in τBal for y = 16.′′6
(left panel) and by a more abrupt (but still insufficient to raise the total τBal to 0.67) one
at y = 13.′′4 (right panel). What is likely happening here is that we are seeing in the optical
is ionized gas near the front side of the attenuation flap, likely affected by it, but not the
material behind it, which is obscured to the point that we can only see it in radio continuum.
Third, if we apply the model described in the Appendix, we obtain τcov = 2.17 and γ = 0.549.
The value of γ appears to be consistent with the observed morphology but τcov is too small
(if it were really that low, we would detect a much higher value of τBal at the location of
the flap). However, one should bear in mind that the model assumes no attenuation for the
uncovered region and, as we have seen, the low-attenuation part of CD has τBal ≈ 0.2. If
we add an additional uniform screen that affects all the ionized gas with τBal = 0.18, all the
curves in Fig. 8 are displaced towards the top and we end up with a γ of approximately 0.5
and an arbitrarily large τcov. In this sense, the values of τcov provided in Table 4 should be
understood only as a lower limit (on the other hand, it is easy to see from Fig. 8 that the
values of γ are not so strongly affected by the vertical displacement induced by a uniform
foreground screen if its optical depth is not too large). We can conclude that all available
evidence suggests the existence of a high-attenuation flap that covers a significant fraction
of the SW high-excitation shell.
The last attenuation maximum, located towards the N of subregion F1, corresponds to
a weak CO (1 → 0) maximum. In this case, the effect on the H II gas in region F is not as
important (τrad = 0.45) and τ ′
Bal is very close to τrad, so a patchy foreground screen with a
low value of τcov provides a good fit to the observed properties (this can be seen also in the
proximity to the diagonal (τBal=τrad) of F in Fig. 8). The last of the regions, G, is located
along the eastern edge of the main cavity and also includes the NE high-excitation filament.
Most of it experiences low attenuation, with the only exception of G4, which covers the
northern tip of the attenuation feature of the main molecular cloud. All of this results in
the lowest values of τrad, τBal, and τ ′
Bal among all regions. Note also how the simple patchy
foreground model of the Appendix correctly characterizes what is seen in the high resolution
data: the value of τcov is similar to that of its neighbors, A and H, but γ is much lower since
attenuation affects significantly only G4.
On the issue of the relative location of the CO clouds with respect to the extinction
experienced by different sources it is interesting to point out the recent result by Bluhm
– 22 –
et al. (2003). Those authors used FUSE to try to detect H2 absorption towards NGC 604
but they failed to do so. The morphology described in this paper easily explains the reasons
for that non-detection: it is not that there is no molecular gas in NGC 604 but rather that
there is very little in front of most of the UV-bright stars. Note that even if there were many
massive stars embedded in the CO cloud it would be hard to detect the H2 signature in the
integrated FUV spectrum, since those stars would be too extinguished in that wavelength
range and most of the detected FUV photons would come from the unextinguished stars. In
order to detect a strong FUV H2 signature in a integrated spectrum, it is required that the
molecular gas covers most or all of the massive stars.
4. DENSITY AND EXCITATION
4.1. Density maps
We show in Fig. 9 the maps of the electron density ratio [S II] λ6717/[S II] λ6731 derived
from the long-slit data, along with a synthetic Hα map obtained in the same way. Given
that low S/N data have large uncertainties that render the values of the density ratios
meaningless, only those points where the uncertainty in [S II] λ6717/[S II] λ6731 (as derived
from the fitting to the spectra) is less than 0.08 are shown. In order to extract as much
information as possible, two maps and one table are shown at different spatial resolutions.
Lower spatial resolutions allow us to extend the maps to larger areas, increasing the S/N by
smoothing over more pixels.
Most of NGC 604 has values of [S II] λ6717/[S II] λ6731 > 1.30, which corresponds to
the low-density regime ne . 130 cm−3 (Castaneda et al. 1992). The low density areas include
the central cavity, most of the surrounding shell, and the blue-shifted Hα knot. Only a few
small regions of the main shell show values of [S II] λ6717/[S II] λ6731 ≈ 1.30 in the high-
spatial resolution map. The fact that most of the bright regions of the shell are not especially
dense is another point in favor of the description of these regions as 2-D surfaces since it
implies that their high surface brightness is caused by large column densities (explained by
near-edge-on orientations of the surfaces) and not by high densities12.
Three regions are characterized by having a distinct high density in Fig. 9. The first one
is the compact Hα knot in region A centered at (100′′,17.′′3). Its high density is shown in the
two maps in Fig. 9 and also in the left panel of Fig. 10. There we can see that the density is
12Note that we are unable to calculate detailed filling factors for the gas in NGC 604 using a combination
of the emission measure and the density since we can only provide an upper bound to the density.
– 23 –
well correlated with the Hα intensity in the 5′′ around the knot, reaching a maximum value
around 250 cm−3. Compare this with the Hα-bright southern part of the main shell (around
x = 87.5′′ in the same plot), which shows lower densities and a much poorer correlation
between intensity and density.
The second high-density region is located at the southernmost (top) edge of the maps
around (105′′,20.′′9). The region is dense enough to appear in yellow in the bottom table of
Fig. 9. We also show in the right panel of Fig. 10 the density as a function of the y coordinate,
with a maximum value around 360 cm−3. The density gradient in that plot shows a good
correlation with τrad which itself increases as one goes into the largest molecular cloud in NGC
604. The higher-resolution τBal map shows that this specific region is highly extinguished,
which leads to the possibility that there may be a barely-visible compact H II region. This
area is adjacent to the location of a stellar group centered at (103′′,19.′′5) that contains a
WR/Of star (WR11, Drissen et al. 1993) and a RSG (Terlevich et al. 1996). This leads to
the interesting possibility of having a compact H II region, a WR star, and a RSG within 4′′
(16 pc) of each other, which is quite surprising, given that each of these objects represents
a different stage in the evolution of massive stars separated from the rest by several Myr. It
should be pointed out, however, that the current data does not exclude the possibility of a
chance alignment between the RSG and the other two objects. Furthermore, if the star with
He II λ4686 excess (which is really what Drissen et al. 1993 detected) turned out to be an Of
star instead of a WR, there would not necessarily be an age discrepancy with the compact
H II region. The situation would be similar to that of knot 2 in 30 Doradus, where an early-O
star is partially embedded in a molecular cloud adjacent to the main stellar cluster (Walborn
et al. 2002).
The third high-density region is located around (105′′,17.′′3), 4′′ to the N of the second
one. It can be seen in the third panel of Fig. 9 and as a secondary maximum in the right
panel of Fig. 10. This region corresponds to the location of another WR/Of star, WR7
(Drissen et al. 1993) and is further discussed in the next subsection.
4.2. Excitation structure
We show in Fig. 12 maps of the three excitation ratios derived from the WFPC2 data,
[O III] λ5007/Hβ, [S II] λ6717+6731/Hα, and [N II] λ6584/Hα. As we did for the Hα/Hβ
map, we first smoothed with a 5×5 pixel box the continuum-subtracted emission-line images,
calculated their ratios, and plotted the result with the low S/N ratio areas and bright stars
masked out. Inspection of emission line diagnostic diagrams (Fig. 13) reveals two types of
ionization structures: (a) a main central ionization structure (CIS) produced by the central
– 24 –
SOBA, and (b) a number of secondary ionized structures (SISs) localized outside the main
shell and energized by nearby small stellar groups.
The CIS is dominated by the main shell (shell A) centered at (90′′,9′′). The upper
panel in Fig. 14 shows the radial distribution of the CIS Hα flux (in log scale), where it can
be seen that the general appearance is that of an empty shell integrated along the line of
sight, with an inner radius of 8′′, plus an extended low-surface-brightness halo. The overall
appearance of the CIS is that of an inner ≈ 20′′ radius high excitation zone surrounded by
an outer larger low excitation halo. The high excitation zone is relatively symmetrical about
the main shell. The low excitation halo is asymmetrical; it extends out to ≈ 45′′ towards the
North, East and South, but it is significantly less important towards the West. Figure 13
shows the diagnostic diagram for four quadrants and a difference is readily apparent there
in one of the four cardinal directions: the W±45◦ diagram has all its pixels in the high to
intermediate excitation regime, log([O III] λ5007/Hβ) > −0.2, indicating that the nebula is
density bounded towards the West.
The lower panel in Fig. 14 shows the log([O III] λ5007/Hβ) vs. log([S II] λ6717+6731/Hα)
diagnostic diagram of the CIS. This diagram has been constructed from the pixel-by-pixel
diagnostic diagram of the points belonging to the CIS, and where each square represents
the density of individual points located in that part of the diagram. The contours give the
density of points in log scale, while color is used to code the average radial distance of those
points to the CIS center. We see that those points closer to the center (blue colors) have
high excitation, while those further away (red) tend to be of lower excitation. Thus, this
diagnostic diagram traces the overall ionization structure of the CIS, as expected for a simple
shell+halo structure. We have run a series of models similar to those described in the simple
approach followed by Gonz´alez Delgado & P´erez (2000). The density distribution used is
taken as the azimuthal average rms density distribution as obtained from the Hα flux, and
converted to actual density via the filling factor φ (taken as a parameter). For a range of ages
between 1 and 5 Myr, and φ = 0.1, 0.01, 0.001, we have used Cloudy (Ferland 1997) with the
same input SEDs as in Gonz´alez Delgado & P´erez (2000). The output radial distribution
of emission line fluxes are integrated along the line of sight, assuming spherical symmetry.
Only those models in the age range 2.75−3.0 Myr and φ = 0.1 fall close enough to the CIS
ionization structure (see Fig. 14). We notice that: (i) as concluded by Gonz´alez Delgado &
P´erez (2000), the age of the cluster ionizing the CIS is 2.5−3 Myr; and (ii) simple spherically
symmetric models produce a well defined line in the diagnostic diagram, while the CIS points
in NGC 604 are distributed along a thick region in the ionization structure. This spread in
NGC 604 is produced by inhomogeneities in the detailed structure, with each radial direction
from the center having a different actual density distribution. An additional cause for the
spread is the fact that the ionizing stars are not all located at the center of the CIS, i.e.,
– 25 –
NGC 604 is ionized by a SOBA and not by a SSC.
Outside the main shell a number of SISs can be identified by their clear footprints on
the diagnostic diagrams; they are ionized by a few or even just a single massive star. They
are as follows.
• The filled H II region in region F, centered at (76′′,5′′) with a radius of 6′′. The ionizing
source consists of a small group that produces a maximum log([O III] λ5007/Hβ) = 0.4.
The extinction is low, with τrad = 0.45.
• The region around WR7, located at (104′′,18′′) with a radius of 1.′′5. It is ionized by
a single WR/Of star, identified as WR7 by Drissen et al. (1993), and reaches a high
excitation value of log([O III] λ5007/Hβ) = 0.6. This star is located in a ridge where
τBal grows rapidly from 0.7 to 1.2. As mentioned in the previous subsection, it also
corresponds to a region of high density as measured from the [S II] λ6717/[S II] λ6731
ratio, possibly a compact H II region.
If the star and the compact H II region are
physically associated, the star should be an Of rather than a Wolf-Rayet.
• The filled H II region in region A, centered at (100′′,17′′) and with a radius of 2.′′5, is
the brightest and more compact within NGC 604. Ionized by a handful of UV-bright
stars, it presents the highest ionization level, with up to log([O III] λ5007/Hβ) = 0.75.
Located at the edge of the main molecular cloud, it has a value of τrad of 0.92.
• The high-intensity H II gas in region B is elongated along the SN direction and just
south of the filled H II region in region A. It has an approximate size of 6′′ × 9′′ and
its ionization level is not very high, log([O III] λ5007/Hβ) = 0.35 − 0.40, except in an
unresolved high excitation knot at (100′′,26′′) where it reaches 0.6.
• Region E corresponds to the SNR described by D'Odorico et al. (1980) and it is centered
at (94′′,35′′) with a radius of 1.′′5. Its ionization footprint is conspicuous only in the
ratio [S II] λ6717+6731/Hα, with logarithmic values between −0.2 and 0.0, with all
other points in NGC 604 having values smaller that −0.5 for the log of this ratio. The
excitation ratio [O III] λ5007/Hβ has rather low values (logarithm between −0.3 and
0.2) except for the SE part of the SNR, where an [O III] λ5007-bright knot raises the
value to log([O III] λ5007/Hβ) = 0.4 − 0.5. Contamination from the SNR hampers
the measurement of the extinction, but an analysis of the surroundings indicate a low
value. It is interesting to point out that the effect of the SNR on the global emission-
line ratios for NGC 604 is very small, as evidenced by comparing the last two lines
in either Table 3 or Table 4.
It would be impossible to detect its presence using
spatially-unresolved optical data.
– 26 –
• Another high-excitation region is located at (77′′,9′′) and has a radius of 1.′′5. Given
the symmetry of the local ionization structure, this node is probably ionized by a
single star with a hard ionizing spectrum. The star is 0.′′5 towards the W of the
star identified as 139 by Drissen et al. (1993). The ionization level is high, with
log([O III] λ5007/Hβ) = 0.7. Extinction is intermediate.
The detailed ionization structure of an H II region is related to the issue of whether all the
ionizing radiation produced by the massive stars is processed within the nebula or whether
some of it escapes into the more general interstellar medium of the host. Gonz´alez Delgado &
P´erez (2000) modeled the integrated spectral properties of NGC 604, including an analysis of
both the stellar spectral energy distribution (SED) and the photoionization of the integrated
nebular spectrum. They concluded that, within this integrated modeling, the SED of the
stars was adequate to account for the gas ionization and extension, and that there was
no leak of ionizing radiation. The data presented here shows two pieces of evidence that
argue against that statement. First, the western quadrant of the nebula appears to be
matter bounded, as seen in the ionization structure and diagnostic diagrams shown above.
Notice that the second largest molecular cloud is located in this direction, which in principle
might seem incompatible with the matter bound scheme. This apparent discrepancy can be
resolved if we locate the molecular cloud "behind" (along the line of sight) and the ionized
gas is "in front", as indicated by the extinction analysis above, and assume that the ionized
cloud is "broken" so that the radiation escapes after producing the O++ zone and there is no
ionization front trapped in this direction. Second, we have measured the velocity field along
the 14′′-wide WHT scan spectrum at PA=60◦. Figure 15 shows the velocity field of Hβ (filled
points) and of [O III] λ5007 (open circles), together with the Hβ flux distribution (dotted
line). The horizontal line marks the systemic velocity of −255 km s−1 (Tenorio-Tagle et al.
2000) and the horizontal scale (in ′′) is centered in the main shell (seen as the two rather
asymmetrical peaks in the flux distribution), increasing towards the SW. At the SW edge of
the shell the gas velocity suddenly jump to −270 km s−1. This blue-shifted high-excitation
ionized gas can be interpreted as further evidence that the shell has been broken and that
the shreds have been blown out onto the line of sight, as suggested by Tenorio-Tagle et al.
(2000). In any case, the final word on the nature of NGC 604 as an ionization- or matter-
bounded nebula requires a complete characterization of the young stellar population, which
we will address in a subsequent paper.
– 27 –
5. DISCUSSION AND CONCLUSIONS
A consistent picture emerges from our analysis of the gas distribution in NGC 604: the
≈ 3.0 Myr-old MYC (Gonz´alez Delgado & P´erez 2000; Ma´ız-Apell´aniz 2000) has carved a
hole in its surrounding ISM and HAS punctured it in several places, leading to the formation
of a series of cavities and tunnels. This low-density medium is filled with hot coronal gas
that emits in X-rays and is transparent to the ionizing UV radiation. The leftover molecular
gas from the parent cloud is still visible along several directions, but for others, including
the direct line of sight to the central part of the SOBA, it appears to have been almost
completely cleared out. What we see as a giant H II region is a composite of (a) localized
high-intensity, high-excitation gas on the surfaces of the molecular clouds directly exposed to
the ionizing radiation, and (b) a diffuse low-intensity, low-excitation component that extends
for several hundreds of pc. A similar morphology is also observed in 30 Doradus. Both objects
have a similar partition of the fluxes for different emission lines, with most of the photons
from medium and high excitation species (e.g. Hα and [O III] λ5007) being produced in
the surfaces adjacent to the molecular clouds and a more even distribution for the photons
from low-excitation species (e.g.
[S II] λ6717+6731 and [N II] λ6584). The high-excitation
regions are indeed near-bidimensional in character, given that their thicknesses (∼ 1 pc) are
much smaller than their extensions (several tens of pc), so that a more appropriate name
for them may be H II surfaces. The morphology of the H II gas in the halo is less clear:
does it fill most of the volume around the giant H II region or is it concentrated in a series
of thin shells? The complex kinematics of the halo of both 30 Doradus (Chu & Kennicutt
1994) and NGC 604 (Yang et al. 1996; Ma´ız-Apell´aniz 2000), where two or more individual
kinematic components are detected in most positions, favors the second option. This H II
surface + extended halo morphology observed in NGC 604 and 30 Doradus (Walborn et al.
2002) is also consistent with what is observe in objects at larger distances and lower spatial
resolutions, such as NGC 4214-II (MacKenty et al. 2000), where the low-excitation halo is
easily resolved but the H II surface is reduced to a quasi-point-like high excitation core.
Assuming the validity of the comparison between the structure of the giant H II region
in NGC 604 and in other objects such as 30 Doradus, we can make two predictions for future
observations. One is that wherever H II surfaces are present one should also detect the PDR
in e.g. the NIR H2 emission lines (Rubio et al. 1998). The second one is that if we obtain
higher-resolution images of the H II surfaces, we should detect dust pillars similar to the
ones seen from the Eagle Nebula to 30 Doradus (Scowen et al. 1998; Walborn 2002). Those
pillars should be easier to detect where the H II surface is seen near-edge-on and should be
locations where induced star formation may be taking place. Another such place where we
may be witnessing induced star formation is the bright compact H II region in region A,
which is similar to knot 1 in 30 Doradus in terms of apparent stellar content, orientation
– 28 –
with respect to the molecular cloud, and high surface brightness (Walborn et al. 2002).
For distant giant H II regions we cannot resolve the individual components and we can
only analyze their spatially-integrated properties. Our analysis suggests that characteriza-
tion of the extinction from spatially-integrated studies might be quite uncertain, especially
when the amount of dust is relatively large. Taking NGC 604 as an example, if we would
perform an integrated study of its properties, the underestimation of the number of ionizing
photons as derived from the Fα/Fβ ratio would be around 27%, compared to an underesti-
mation of around 11% derived from our high-spatial-resolution data. Just for comparison,
if no correction at all is applied, the underestimation increases to around 42%. In principle,
the analysis of the radio continuum could provide a good estimate of the ionizing flux being
emitted, indeed more accurate than the value derived from the Balmer lines ratio. Neverthe-
less, we want to stress that the non-thermal contribution to the radio continuum could lead
to an overestimation of the ionizing flux, unless high-resolution radio observations are used
to separate the contributions from non-thermal emitting regions. In the case of relatively
unevolved regions, like NGC 604, the non-thermal component is negligible (≈ 1% at 3.6 cm),
but it might be significant for other objects (see MacKenty et al. 2000 for NGC 4214). We
want to stress that the uncertainties in the ionizing flux values derived from emission lines
or thermal radio continuum make it very uncertain to derive the number of ionizing photons
escaping unabsorbed from the region, unless a careful analysis of both the stellar population
and the structure of the absorbing agents is performed.
The reason why it is not possible to produce a simple correction for the attenuation
experienced by the gas is the intrinsically complex geometry of a giant H II region. The
sources (H II surfaces and halos) and the extinction agents (the dust particles located mainly
in the adjacent molecular clouds) are extended and, to a certain degree, intermixed. As a
consequence, sources located in different points in the plane of the sky may have quite
different amounts of dust in front of them. Furthermore, two sources along the same line
of sight may have large amounts of dust between them, as we have seen for the case of the
"flaps". Our WFPC2 images show how extinction can rapidly vary in scales of a pc or less
and that even relatively small structures of the order of a few tens of pc can hide a significant
fraction of the optically-unobservable H II gas. We estimate that most of the . 11% of the
Hα flux that is not recoverable except using radio (or maybe NIR) observations could be
accounted for simply by extending the optically bright H II surfaces in the B and CD regions
into an area equivalent to the one covered by each flap, as derived from the optical geometry
in Fig. 3 (10-20 square arcseconds in each case).
This complex geometry of GHRs explains the well known effect that in many galaxies
hosting strong starbursts the average extinction experienced by the stars is significantly
– 29 –
lower than the integrated extinction derived from the Balmer emission lines, as discussed
e.g. by Calzetti et al. (2000). In NGC 604 the SOBA is located in a low extinction region
because it has cleared out a cavity around it. A similar result was obtained for NGC 4214
(Ma´ız-Apell´aniz et al. 1998). In the next paper we will analyze the extinction affecting the
stellar continuum of the individual stars in NGC 604 one by one, in more detail.
We can conclude that in NGC 604 a large fraction of the extinction is caused by dust
associated with the GHR itself, although not evenly distributed. A similar conclusion was
reached by Caplan & Deharveng (1986) for H II regions in the LMC using unresolved data.
It should be pointed out that those authors indicated already almost twenty years ago that
one of the ways to continue work in this field was by using "point-by-point emission-line
photometry". Our analysis shows that this technique allows indeed a better characterization
of the extinction properties and, moreover, confirms the results obtained by these authors
on the LMC GHRs.
The main cavity in NGC 604 has a diameter of ≈ 60 pc, which is several times smaller
than the value predicted by applying the single-star wind-blown model of Weaver et al.
(1977). According to that model, bubble sizes should be very weakly dependent on wind
luminosity or on the density of the surrounding medium, so varying those parameters does
not provide a solution to the discrepancy. As a matter of fact, any of the four cavities in
NGC 604 could have been created by the kinetic energy released into their surroundings by
only one or a few massive stars (as opposed to the ≈ 200 here) in a period of 3 Myr if Weaver
et al. (1977) models were applicable here. This discrepancy is similar to the ones found for
NGC 4214 and other objects by MacKenty et al. (2000) and Ma´ız-Apell´aniz (2001). We
believe that the explanation lies in that the ISM is far from being in pressure equilibrium.
Recent numerical simulations by Mac Low (2004) show that the ISM is extremely dynamic,
with molecular clouds being transient objects formed along the edges of superbubbles by the
collision of two or more of them. In those simulations, the gas compressed in such a way can
collapse to form stellar clusters but may have only a short time to do so, since superbubbles
come and go in scales of only ∼ 10 Myr. In other environments, such as dwarf galaxies,
MYCs may form in a different way but the requirement for a source of external pressure
should still be present. NGC 604 is known to be located at the edge of an H I-detected
superbubble centered ∼ 1 kpc towards the SE (Thilker 2000). According to the previously
mentioned simulations, that superbubble may have triggered the formation of NGC 604 and
should still be pushing the two largest molecular clouds associated with NGC 604 in the
opposite direction to the one in which the winds and SNe from the NGC 604 SOBA are
pushing them (see Fig. 3 in Thilker 2000). That source of external pressure is likely to be
one of the reasons for the smaller-than-expected sizes of the bubbles in NGC 604.
– 30 –
Two additional mechanisms that act to make superbubbles smaller than their expected
sizes at a given age may also be in effect. On the one hand, when a giant molecular cloud
is formed and once the gas becomes dense enough, its pressure should increase locally due
to self-gravity. Under those conditions, wind bubbles from isolated stars have to fight an
additional pressure and have difficulty expanding to their expected sizes, especially in the first
stages of their development (Garc´ıa-Segura & Franco 1996). In a stationary isolated case,
the growth of a wind-blown bubble can stall completely. In a more realistic environment, the
motion of the star should eventually take it out of its dense birth region and its bubble may
expand to a size large enough to find an adjacent bubble. The resulting wind-wind collision
should lead to a second thermalization (the first one being produced by the initial reverse
shock wave created by the interaction of the stellar wind with its surrounding dense medium)
and, after adding more recently-formed massive stars to the bubble, become the seed for a
superbubble. However, the whole process may be delayed by several hundreds of thousands
of years by the initial confinement of the individual bubbles, thus contributing to the excess
in the number of giant H II regions without superbubbles detected by Ma´ız-Apell´aniz (2001).
The second mechanism that should contribute to the small observed sizes of superbubbles
lies in the inhomogeneity of the large-scale ISM. Once a superbubble expands to a size of
∼ 100 pc, it is likely to encounter a density gradient in one or several directions. The
superbubble is then punctured and the hot coronal gas flows out of the cavity through the
created hole(s). The subsequent reduction in the internal pressure of the cavity should slow
down its expansion and, eventually, halt it and even reverse it. As discussed by Tenorio-
Tagle et al. (2000), the observed morphology and kinematics of NGC 604 is consistent with
such a puncture.
In this paper we have presented our results on the spatial distribution of the different gas
phases and the dust in NGC 604. In the next paper we will study the massive stellar popu-
lation by means of HST UV-optical photometry and HST UV objective-prism spectroscopy.
The analysis of the individual stars will provide a complete picture of the relationship be-
tween the gas and the stars in NGC 604, and will allow to measure the fraction of ionizing
photons that potentially escape from the nebula into the general ISM of M33.
We would like to thank Greg Engargiola, Ed Churchwell, and Guille Bosch for giving
us access to their data. We would also like to thank Rodolfo Xeneize Barb´a for his help
with the CO data processing and Henri Plana for his help with the calibration of the long-
slit data. This work started as part of the GEFE collaboration. We have enjoyed the
benefits from many a discussion at different workshops and brainstorming sessions over the
years with our colleagues inside and outside of GEFE. Support for this work was provided
by NASA through grant GO-09096.01-A from the Space Telescope Science Institute, Inc.,
– 31 –
under NASA contract NAS5-26555; by the Spanish Government grants CICYT-ESP95-0389-
C02-02, AYA-2001-3939-C03, and AYA-2001-2089; and by the Mexican government grant
CONACYT 36132-E.
REFERENCES
Anderson, J., & King, I. R. 2003, PASP, 115, 113
Barb´a, R. H., Rubio, M., Roth, M. R., & Garc´ıa, J. 2003, AJ, 125, 1940
Bluhm, H., de Boer, K. S., Marggraf, O., Richter, P., & Wakker, B. P. 2003, A&A, 398, 983
Bosch, G., Terlevich, E., & Terlevich, R. 2002, MNRAS, 329, 481
Calzetti, D., Armus, L., Bohlin, R. C., Kinney, A. L., Koornneef, J., & Storchi-Bergmann,
T. 2000, ApJ, 533, 682
Calzetti, D., Harris, J., Gallagher, J. S., III, Smith, D. A., Conselice, C. J., Homeier, N., &
Kewley, L. 2004, AJ, 127, 105
Caplan, J., & Deharveng, L. 1986, A&A, 155, 297
Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245
Casertano, S., & Wiggs, M. S. 2001, WFPC2 Instrument Science Report 2001-10 (STScI)
Castaneda, H. O., V´ılchez, J. M., & Copetti, M. V. F. 1992, A&A, 260, 370
Cervino, M., Mas-Hesse, J. M., & Kunth, D. 2002, A&A, 392, 19
Chu, Y.-H., & Kennicutt, R. C., Jr. 1994, ApJ, 425, 720
Churchwell, E., & Goss, W. M. 1999, ApJ, 514, 188
D´ıaz, A. I., Terlevich, E., Pagel, B. E. J., V´ılchez, J. M., & Edmunds, M. G. 1987, MNRAS,
226, 19
D'Odorico, S., Dopita, M. A., & Benvenuti, P. 1980, A&AS, 40, 67
Dolphin, A. E. 2000, PASP, 112, 1383
Drissen, L., Moffat, A. F. J., & Shara, M. M. 1993, AJ, 105, 1400
Engargiola, G., Plambeck, R. L., Rosolowsky, E., & Blitz, L. 2003, ApJS, 149, 343
– 32 –
Esteban, C., Peimbert, M., Torres-Peimbert, S., & Rodr´ıguez, M. 2002, ApJ, 581, 241
Fall, S. M., & Rees, M. J. 1977, MNRAS, 181, 37P
Ferland, G. J. 1997, Hazy, a brief introduction to Cloudy (University of Kentucky, Depart-
ment of Physics and Astronomy Internal Report)
Ferland, G. J. 2001, PASP, 113, 41
Freedman, W. L., et al. 2001, ApJ, 553, 47
Garc´ıa-Segura, G., & Franco, J. 1996, ApJ, 469, 171
Gonz´alez Delgado, R. M., & P´erez, E. 2000, MNRAS, 317, 64
Gordon, S. M., Kirshner, R. P., Long, K. S., Blair, W. P., Duric, N., & Smith, R. C. 1998,
ApJS, 117, 89
Hollenbach, D. J., & Tielens, A. G. G. M. 1997, ARA&A, 35, 179
Hunter, D. A., Baum, W. A., O'Neil, E. J., Jr., & Lynds, R. 1996, ApJ, 456, 174
Kozhurina-Platais, V., Anderson, J., & Koekemoer, A. M. 2003, WFPC2 Instrument Science
Report 2003-02 (STScI)
Mac Low, M.-M. 2004, in How does the Galaxy work? A Galactic tertulia with Don Cox
and Ron Reynolds, E. Alfaro, E. P´erez, and J. Franco (eds.), (Dordrecht: Kluwer
Academic Publishers), in press
MacKenty, J. W., Ma´ız-Apell´aniz, J., Pickens, C. E., Norman, C. A., & Walborn, N. R.
2000, AJ, 120, 3007
Ma´ız-Apell´aniz, J. 2001, in Highlights of Spanish astrophysics II, J. Zamorano, J. Gorgas,
and J. Gallego (eds.), Fourth Scientific Meeting of the Spanish Astronomical Society
(Dordrecht: Kluwer), 113
Ma´ız-Apell´aniz, J. 2000, PASP, 112, 1138
Ma´ız-Apell´aniz, J. 2001, ApJ, 563, 151
Ma´ız-Apell´aniz, J., Mas-Hesse, J. M., Munoz-Tun´on, C., V´ılchez, J. M., & Castaneda, H. O.
1998, A&A, 329, 409
Ochsenbein, F., Bauer, P., & Marcout, J. 2000, A&AS, 143, 221
– 33 –
O'Dell, C. R. 2001, ARA&A, 39, 99
O'Dell, C. R., & Doi, T. 1999, PASP, 111, 1316
Parker, J. W., Garmany, C. D., Massey, P., & Walborn, N. R. 1992, AJ, 103, 1205
Rubin, R. H., et al. 2002, MNRAS, 334, 777
Rubio, M., Garay, G., & Probst, R. 1998, The Messenger, 93, 38
Russell, J. L., Lasker, B. M., McLean, B. J., Sturch, C. R., & Jenkner, H. 1990, AJ, 99, 2059
Sabalisck, N. S. P., Tenorio-Tagle, G., Castaneda, H. O., & Munoz-Tun´on, C. 1995, ApJ,
444, 200
Scowen, P. A., et al. 1998, AJ, 116, 163
STScI. 1999, Synphot User's Guide, Howard Bushouse and Bernie Simon (eds.)
STScI. 2002, WFPC2 Data Handbook, Sylvia Bagget and Matthew Mc Master (eds.)
Tenorio-Tagle, G., Munoz-Tun´on, C., & Cid-Fernandes, R. 1996, ApJ, 456, 264
Tenorio-Tagle, G., Munoz-Tun´on, C., P´erez, E., Ma´ız-Apell´aniz, J., & Medina-Tanco, G.
2000, ApJ, 541, 720
Terlevich, E., D´ıaz, A. I., Terlevich, R., Gonz´alez Delgado, R. M., P´erez, E., & Garc´ıa-
Vargas, M. L. 1996, MNRAS, 279, 1219
Thilker, D. A. 2000, in The Interstellar Medium in M31 and M33, 3
Viallefond, F., Boulanger, F., Cox, P., Lequeux, J., P´erault, M., & Vogel, S. N. 1992, A&A,
265, 437
Walborn, N. R. 2002, in ASP Conf. Ser. 267: Hot Star Workshop III: The Earliest Phases
of Massive Star Birth, 111
Walborn, N. R., Barb´a, R. H., Brandner, W., Rubio, M., Grebel, E. K., & Probst, R. G.
1999, AJ, 117, 225
Walborn, N. R., Ma´ız-Apell´aniz, J., & Barb´a, R. H. 2002, AJ, 124, 1601
Walborn, N. R., & Parker, J. W. 1992, ApJ, 399, L87
Weaver, R., McCray, R., Castor, J., Shapiro, P., & Moore, R. 1977, ApJ, 218, 377
– 34 –
Wilson, C. D., & Scoville, N. 1992, ApJ, 385, 512
Yang, H., Chu, Y.-H., Skillman, E. D., & Terlevich, R. 1996, AJ, 112, 146
APPENDIX: DUST MODELS
The ratio between the Hα and radio emissivities of a photoionized region with n(He+)/n(H+) =
0.09 for case B is (Caplan & Deharveng 1986; Churchwell & Goss 1999):
q
jα (erg s−1 cm−2)
jν (Jy)
= 7.90 · 10−10(cid:18) T
104 K(cid:19)−0.59
ν
109 Hz(cid:17)0.1
(cid:16)
,
(1)
where the possible errors due to the power-law approximations used and the possible uncer-
tainty in the He abundance are of the order of a few percent. From Eq. 1 we obtain that the
(true) optical depth at Hα measured from the radio and Hα fluxes, Fν and Fα:
τrad = ln"1.27 · 109(cid:18) Te
104 K(cid:19)0.59
ν
109 Hz(cid:17)−0.1
(cid:16)
Fν(Jy)
Fα(erg s−1 cm−2)# ,
(2)
For NGC 604, we use T = 8 500 K (D´ıaz et al. 1987; Gonz´alez Delgado & P´erez 2000;
Esteban et al. 2002) and for the frequency of ν = 8.44 GHz of Churchwell & Goss (1999)
Eq. 2 becomes:
τrad = ln(cid:20)9.32 · 108
Fν(Jy)
Fα(erg s−1 cm−2)(cid:21) ,
The corresponding ratio between the Hα and Hβ emissivities is, within ∼ 1%:
jα
jβ
= 2.859(cid:18) T
104 K(cid:19)−0.07
,
(3)
(4)
where the two emissivities are measured in the same units. From Eq. 4 and the Cardelli
et al. (1989) law for RV = 3.2 we obtain the optical depth at Hα measured from the ratio of
the two Balmer fluxes, Fα and Fβ:
This preprint was prepared with the AAS LATEX macros v5.2.
– 35 –
τBal = 2.42 ln"Fα/Fβ
2.859 (cid:18) Te
104 K(cid:19)0.07# ,
which for T = 8 500 K becomes:
τBal = 2.42 ln(cid:20) Fα/Fβ
2.892 (cid:21) .
(5)
(6)
One of the largest uncertainties in measuring τBal comes from the assumed value of RV .
For example, values of 2.6 or 4.4 change the constant that multiplies the logarithm in Eq. 5
to 2.04 and 3.07, respectively. Here we will ignore this question but we plan to study it in
future works.
The standard attenuation model is that of a uniform foreground screen model. There
we have that τrad=τBal, since all areas observed are affected by dust in the same degree and
no scattering is injecting photons back into the line of sight.
A uniform foreground screen is not very realistic for H II regions, since images reveal
the existence of localized dust clouds. Here we explore a model in which we assume that
the dust in the aperture is distributed in a patchy foreground screen, in such a way that the
area covered is γ times the area free of dust and the screen yields an optical depth at Hα
of τcov for the areas behind it. In this case, the values of τrad and τBal for the aperture will
be weighted means of τcov (the optical depth experienced by the areas covered by dust) and
zero (the optical depth experienced by the areas free of dust).
Our results are shown in Fig. 8, where we represent τBal as a function of τrad for different
values of γ. As expected, τBal is always smaller than τrad when the dust is not in a uniform
foreground screen and scattering is not relevant (see e.g. Caplan & Deharveng 1986. For
our patchy foreground screen model, τBal first increases while the obscured areas contribute
significantly to the detected radiation. When the amount of dust present in the screen
becomes large enough, the curve acquires a negative slope and ends up going back to zero
since the dominant effect is to let only the unobscured regions contribute to the detected
Balmer photons. Note that the region defined by τBal ≥ 0, τrad ≥ τBal is completely covered
by the family of models for τcov ≥ 0, 0 ≤ γ ≤ 1, with a one-to-one mapping which is only
degenerate for τBal = τrad = 0.
– 36 –
Fig. 1.- The ten long-slits superimposed on one of the F656N WFPC2 images. Note that
north is towards the bottom and that this orientation is used throughout the article.
– 37 –
Fig. 2.- WFPC2 continuum-subtracted [S II] λ6717+6731 (left) and [O III] λ5007 (right)
images. Both images are displayed with a hyperbolic intensity scale in order to show both
low and high intensity structures. White pixels are used to block regions with bright stars
for the sake of clarity (continuum subtraction is not perfect there due to the under-sampled
nature of the WFPC2 PSF). The larger extension of the emission in the [S II] λ6717+6731
image as compared to that of the [O III] λ5007 is not an artifact of the choice of displayed
levels. On the contrary, the dynamic range (defined as the ratio between maximum and
minimum displayed levels) for the image on the left is 100 while that for the image on the
right is 1 000. Therefore, if both lines were emitted from regions of similar size, one would
expect the [O III] λ5007 region to appear larger instead of smaller. The letters A to D
indicate the location of the four cavities described in the text. The field is 80′′ on a side and
the orientation is the same as in the rest of the figures. See Fig. 9 for an explanation of the
coordinate system.
– 38 –
Fig. 3.- (left) Color mosaic of the central region of NGC 604 composed by assigning F673N
to the red channel, F555W and F656N to the green one, and F336W to the blue one (the
filters were not processed for continuum or line subtraction). The field is 39′′ on a side and
the orientation is the same as in the rest of the figures. See Fig. 9 for an explanation of the
coordinate system. (right) Explanatory diagram for some of the structures discussed in the
text. Violet is used to display cavity contours, with a dashed line style when the boundary
between cavities is uncertain (possibly due to superbubble bursting), and shading is used to
mark the high excitation parts of the nebula (H II surfaces), with yellow used for the regions
with the highest intensity and green used for the rest.
– 39 –
Fig. 4.- Contour diagram of the CO (1 → 0) emission in NGC 604 (adapted from the data
of Engargiola et al. 2003) superimposed on an F656N WFPC2 image (without continuum
subtraction). The spatial resolution of the contour diagram is 13′′. The field is 110′′ on a side
and the orientation is the same as in the rest of the figures. See Fig. 9 for an explanation of
the coordinate system.
– 40 –
Fig. 5.- Contour diagram of the X-ray emission in NGC 604 superimposed on an F656N
WFPC2 image (without continuum subtraction). The field and orientation is the same as
in Fig. 4.
– 41 –
Fig. 6.- (Left) Continuum-subtracted Hα WFPC2 image of NGC 604. (Right) τBal map of
the same region as derived from WFPC2 data smoothed with a 5 × 5 WF pixels box. Areas
in white have been masked due to strong stellar contamination or low signal-to-noise. The
sub-regions used in this paper are shown. The field is the same as the one in Fig. 2. See
Fig. 9 for an explanation of the coordinate system.
– 42 –
Fig. 7.- (Upper left) Low (≈ 4′′) resolution continuum-subtracted Hα image of NGC 604.
(Upper right) 8.4 GHz radio continuum image of NGC 604 at the same resolution. (Lower
left) τBal map of the same region at that resolution. (Lower right) τrad map of the same region
at that resolution. The regions used in this paper are shown in the upper two panels. Areas
in white in the two lower panels have been masked due to low signal-to-noise. Contours from
the CO data shown in Fig. 4 are plotted in the lower two panels. The field is the same as
the one in Fig. 2. See Fig. 9 for an explanation of the coordinate system.
– 43 –
τrad
0.00
0.25
0.50
0.75
1.00
1.00
0.75
0.50
0.25
l
a
B
τ
γ = 0.3
γ = 0.4
γ = 0.5
γ = 0.6
γ = 0.7
γ = 0.8
γ = 0.9
γ = 1.0
τcov = 1
τcov = 2
τcov = 3
τcov = 4
regions
F
NGC 604
A−H
CD
G
H
A
B
0.00
0.00
0.25
0.50
0.75
1.00
τrad
τ
B
a
l
1.00
0.75
0.50
0.25
0.00
Fig. 8.- τBal vs. τrad for a patchy foreground model with different values of γ, the area
covered by the screen. Dashed lines show the location of different values of the optical depth
of the screen. The values obtained for our regions are also plotted.
– 44 –
60
70
80
90
100
110
120
130
140
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
60
Hα map
sm=3
<
1
1
−
1
1
−
1
1
−
1
1
−
1
1
−
1
1
<
1
2
2
3
3
4
5
0
5
0
5
0
.
.
.
.
.
.
sm=9
<
1
1
−
1
1
−
1
1
−
1
1
−
1
1
−
1
1
<
1
2
2
3
3
4
5
0
5
0
5
0
.
.
.
.
.
.
sm=30
<
1
1
−
1
1
−
1
1
−
1
1
−
1
1
−
1
1
<
1
2
2
3
3
4
.
.
.
.
.
.
5
0
5
0
5
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
2
2
3
3
4
5
0
5
0
5
0
1
2
2
3
3
4
5
0
5
0
5
0
1
2
2
3
3
4
5
0
5
0
5
0
1.39 1.33 1.37 1.31 1.37 1.26 1.28 1.41 1.46
1.55 1.40 1.34 1.36 1.37 1.34 1.38 1.45 1.54 1.44
1.48 1.44 1.37 1.39 1.43 1.40 1.41 1.42
1.48
1.40 1.40 1.45 1.44 1.43 1.43 1.41
1.42 1.41 1.41 1.34 1.42 1.36 1.39 1.39
70
80
90
100
110
120
130
140
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
Fig. 9.- Synthetic Hα map (top panel), density ratio maps (two middle panels), and density
ratio table (bottom panel) produced from the long-slit data. The first density ratio map has
been smoothed over 3 pixels (1′′) along each slit while the second density ratio map has been
smoothed over 9 pixels (3′′). The color-coded density ratio table shows data averaged over
2 adjacent long slits and 15 pixels along the slit direction. In each of the lower three panels,
data is shown where the uncertainty of the ratio is less than 0.08 in order to ensure that the
values shown are relevant. The coordinates are expressed in arcseconds with north towards
the bottom and east towards the right (see Fig. 1) with the first slit centered at y = 1′′
and its first pixel at x = 1/3′′. In these coordinates (x, y) = (93.′′4, 14.′′25) corresponds to
(1h 34m 33s, 30◦ 47′) in J2000.
– 45 –
80.0
82.5
85.0
87.5
90.0
92.5
95.0
97.5
100.0
10.0
12.5
15.0
x (arcsec)
y (arcsec)
17.5
20.0
22.5
25.0
500
400
300
200
100
0
Hα
ne
τrad
CO (1−0)
ne
500
1.50
400
1.25
1.00
0.75
0.50
0.25
0.00
300
200
100
0
1.50
1.25
1.00
0.75
0.50
0.25
0.00
80.0
82.5
85.0
87.5
90.0
92.5
95.0
97.5
100.0
10.0
12.5
15.0
x (arcsec)
17.5
y (arcsec)
20.0
22.5
25.0
Fig. 10.- (left) Hα flux and electron density at y = 17.′′3 derived from slit data as a function
of the horizontal coordinate established in Fig. 9. The density has been smoothed with a
3-pixel (1′′) box. Hα is expressed in units of 20 · 10−17 erg s−1 cm−2 arcsec−2 and ne is
expressed in cm−3. (right) True optical depth at Hα, CO (1 → 0) intensity, and electron
density at x = 105.′′3 as a function of the vertical coordinate established in Fig. 9. τrad is
obtained from Churchwell & Goss (1999), the CO data is from Engargiola et al. (2003),
and ne is derived from slit data. The density has been smoothed with a 9-pixel (3′′) box.
The value of τrad can be read directly from the labels in the vertical axis, the CO scale is
arbitrary, and ne is expressed in units of 300 cm−3.
– 46 –
x (arcsec)
x (arcsec)
70
6
75
80
85
90
95
100
105
Hα
τBal
[O III] λ5007/Hβ
75
80
85
90
95
100
105
110
70
75
80
85
90
95
100
105
110
x (arcsec)
x (arcsec)
Fig. 11.- Hα flux, τBal, and [O III] λ5007/Hβ measured from WFPC2 data as a function
of the horizontal coordinate established in Fig. 9. The left figure corresponds to y = 16.′′6
and the right one to y = 13.′′4. Hα is shown in arbitrary units while the scale for τBal and
[O III] λ5007/Hβ can be read from the left and right sides, respectively.
5
4
3
2
1
/
β
H
7
0
0
5
λ
]
I
I
I
O
[
0
70
110
1.50
1.25
1.00
0.75
τ
B
a
l
β
H
/
7
0
0
5
λ
]
I
I
I
O
[
0.50
0.25
0.00
70
75
80
85
90
95
100
105
110
3.0
2.5
2.0
1.5
1.0
0.5
0.0
Hα
τBal
[O III] λ5007/Hβ
1.50
1.25
1.00
τ
B
a
l
0.75
0.50
0.25
0.00
– 47 –
Fig. 12.- A continuum-subtracted Hα WFPC2 image of NGC 604 is shown in the upper left
panel. The other three panels show excitation maps of the same region: [O III] λ5007/Hβ
(upper right) [S II] λ6717+6731/Hα (lower left), and [N II] λ6584/Hα (lower right). Areas
in white have been masked due to strong stellar contamination or low signal-to-noise. The
field is the same as the one in Fig. 2. See Fig. 9 for an explanation of the coordinate system.
– 48 –
diagnostic diagram,
log([O III] λ5007/Hβ)
13.- CIS pixel-by-pixel
Fig.
vs.
for the four quadrants (±45◦ centered on the four cardinal
log([S II] λ6717+6731Hα),
points, from left to right and from top to bottom, N, E, S, and W). The inset in each
panel shows the [S II] λ6717+6731/Hα image with superposed circles at radii of 10′′, 20′′,
and 30′′, and two lines indicating the quadrant plotted. The overall loci of the points
in the diagram mark the general features of the CIS, and it is apparent that while the
three diagrams corresponding to N±45◦, E±45◦, and S±45◦ reach to very low values of
log([O III] λ5007/Hβ) < −0.5 at large distances, the W±45◦ diagram has all its pixels in
the high to intermediate excitation regime, log([O III] λ5007/Hβ) > −0.2. This is strong
evidence that the nebula is matter bounded towards the West.
– 49 –
Fig. 14.- The top panel shows the radial distribution of the CIS Hα flux (in log scale). The
general appearance is that of an empty shell integrated along the line of sight, with an inner
radius of 8′′, given by the peak flux, plus an extended low surface brightness halo. The lower
panel shows the log([O III] λ5007/Hβ) vs.
log([S II] λ6717+6731/Hα) diagnostic diagram
of the CIS. This diagram has been constructed from the pixel-by-pixel diagnostic diagram
of the points belonging to the CIS, excluding those regions belonging to SISs. Each square
represents the density of individual points located in that part of the diagram. The contours
give the density of points in log scale (i.e., the contour labeled 3 indicates that that part of the
diagram is populated by 1000 individual image pixels). The color codes for the average radial
distance of those points to the CIS center. Two photoionization models (calculated using
Cloudy) similar to those described in the simple approach followed by Gonz´alez Delgado &
P´erez (2000) are plotted, the black points for an age of 2.75 Myr, and the red ones for an
age of 3 Myr. The density distribution used is taken as the azimuthal average rms density
distribution as obtained from the Hα flux, and converted to actual density via the filling
factor φ (taken as a parameter, and equal to 0.1 in the plot).
– 50 –
Fig. 15.- Velocity field of Hβ (filled points) and of [O III] λ5007 (open circles), together
with the Hβ flux distribution (dotted line). The horizontal line marks the systemic velocity
of −255 km s−1 (Tenorio-Tagle et al. 2000). The abcissa scale is centered in the main shell
(seen as the two rather asymmetrical peaks in the flux distribution) and increases towards
the SW. At the SW edge of the shell the gas velocity suddenly jump to −270 km s−1. This
blue-shifted high-excitation ionized gas can be interpreted as further evidence that the shell
has been broken and that the shreds have been blown out onto the line of sight, as suggested
by Tenorio-Tagle et al. (2000).
|
0802.3146 | 1 | 0802 | 2008-02-21T16:15:54 | Infrared Properties Of AGB Stars: from Existing Databases to Antarctic Surveys | [
"astro-ph"
] | We present here a study of the Infrared properties of Asymptotic Giant Branch stars (hereafter AGB) based on existing databases, mainly from space-borne experiments. Preliminary results about C and S stars are discussed, focusing on the topics for which future Infrared surveys from Antarctica will be crucial. This kind of surveys will help in making more quantitative our knowledge of the last evolutionary stages of low mass stars, especially for what concerns luminosities and mass loss. | astro-ph | astro-ph | Title : will be set by the publisher
Editors : will be set by the publisher
EAS Publications Series, Vol. ?, 2018
8
0
0
2
b
e
F
1
2
]
h
p
-
o
r
t
s
a
[
1
v
6
4
1
3
.
2
0
8
0
:
v
i
X
r
a
INFRARED PROPERTIES OF AGB STARS: FROM EXISTING
DATABASES TO ANTARCTIC SURVEYS.
R. Guandalini 1 and M. Busso 1
Abstract. We present here a study of the Infrared properties of Asymp-
totic Giant Branch stars (hereafter AGB) based on existing databases,
mainly from space-borne experiments. Preliminary results about C and
S stars are discussed, focusing on the topics for which future Infrared
surveys from Antarctica will be crucial. This kind of surveys will help
in making more quantitative our knowledge of the last evolutionary
stages of low mass stars, especially for what concerns luminosities and
mass loss.
1
Introduction
Towards the end of their life, stars of low and intermediate mass (M < 8M⊙)
evolve along the Asymptotic Giant Branch (AGB) phase [see Busso et al. (1999)
and references herein for more details]. In this stage they experience extensive phe-
nomena of mass loss that affect deeply their evolution. Sedlmayr (1994) shows that
stellar winds are also fundamental for the enrichment of the Interstellar Medium
that is replenished by these stars with about 70% of all the matter returned after
stellar evolution. Moreover, AGB stars provide the starting conditions for the
formation of planetary nebulae.
The radiative emission of the cool dust in the infrared (IR) normally dominates
the energy distribution of AGB stars (particularly for the most evolved ones) and
this fact is mainly due to their strong stellar winds. Until recently the bolometric
magnitude of the most evolved AGB stars was difficult to derive, due to insufficient
photometric coverage of the mid-IR range of the electromagnetic spectrum (the
importance of mid-IR observations for AGB stars is clearly shown in Guandalini
et al. (2006), Figure 1). The availability of large IR databases from space-borne
telescopes like ISO, IRTS, MSX has substantially improved this situation. At the
same time, Hipparcos distances for AGB stars have been corrected from various
1 Dipartimento di Fisica, Universit`a di Perugia, Via A. Pascoli, 06123 Perugia, Italy; e-mail:
[email protected]
c(cid:13) EDP Sciences 2018
DOI: (will be inserted later)
2
Title : will be set by the publisher
biases in works like the one from Bergeat & Chevallier (2005) and the period-
luminosity relations found for Miras have been drastically improved as shown in
Whitelock et al. (2006). The study of fundamental physical parameters of these
stars (like luminosity, IR colors and mass loss) can now be be performed in a rather
quantitative way.
However, IR space-borne observations of AGB stars present some disadvan-
tages.
In particular, the duration of the operational period of the telescope is
quite limited, observations with a long time of integration are difficult and AGB
stars are generally observed at a single epoch. All these facts hinder our under-
standing of several basic physical parameters, which are fundamental in the study
of AGB stars.
In this respect a complementary role in solving these problems
could be played by ground-based observations at IR wavelengths, especially from
Antarctica.
The Antarctic Plateau (in particular Dome C) presents the best conditions
available on Earth from the point of view of infrared observations as shown in
several other contributions from this conference: therefore, Antarctica is the best
place where ground-based observations in the (mid-)IR can be performed. An
Antarctic IR Observatory might be crucial for clarifying the final stages of stellar
evolution by:
• observing properties of evolved stars in the Magellanic Clouds, at known
distance and metallicity different from the Milky Way;
• doing the same for the Galactic Center, where the metal blend is different
and the extreme O-enhancement prevents the formation of C stars as shown
in Uttenthaler et al. (2006);
• looking for mass loss calibrations in the IR.
Our efforts are addressed to the preparation for IR observations of AGB stars
from Antarctica through the IRAIT telescope, presented by Tosti et al. (2007)
in this conference. With this aim we are trying to understand which kind of
observations from Antarctica is the best and most promising from the point of
view of AGB stars. In Guandalini et al. (2006),(2007) we are making for these
sources an overview of the IR data available from existing catalogues that will
be completed in future works.
In this note we address some interesting issues
concerning AGB stellar variability from an IR point of view, and the relevance of
Antarctic observations for them.
2
Infrared Variability
AGB stars present strong variability at optical wavelengths and are roughly divided
into three subclasses according to their variability type: Miras, Semiregulars and
Irregulars. Effects due to variability at IR wavelengths are expected to be less
relevant when compared with the optical ones.
Figures 1 and 2 present the available information on the IR SEDs for two groups
of AGB sources that are discussed in detail in Busso et al. (2006). Figure 1 shows
Guandalini et al.: Infrared Properties Of AGB Stars . . .
3
ST Her
RY Dra
1400
1200
1000
800
600
400
200
)
y
J
(
y
t
i
s
o
n
i
m
u
L
0
0
10
300
250
200
150
100
50
)
y
J
(
y
t
i
s
o
n
i
m
u
L
0
0
10
20
30
Wavelength (um)
HD 56126
20
30
Wavelength (um)
600
500
400
300
200
100
)
y
J
(
y
t
i
s
o
n
i
m
u
L
40
50
0
0
10
600
500
400
300
200
100
)
y
J
(
y
t
i
s
o
n
i
m
u
L
40
50
0
0
10
ISO-SWS
MSX
IRAS-PSC
IRAS-LRS
TIRCAM2
2MASS
20
30
Wavelength (um)
Red Rectangle
20
30
Wavelength (um)
40
50
40
50
Fig. 1. The Spectral Energy Distributions (SEDs) of a few sources, as available from
IRAS PSC, IRAS LRS, ISO-SWS, MSX, TIRCAM2 and 2MASS. SEDs with maximum
emission in near-IR, as well as with maximum emission longward of 20µm all show a
constant flux in mid-IR.
distributions that share the property of being non-variable at IR wavelengths over
an elapse of time of almost 20 years. They include Semiregular sources with min-
imal IR excess, in which the SED is peaked in near-IR. They also include evolved
(post-AGB) objects in which the maximum emission is at very long wavelengths
(from 20 to more than 40 µm). Instead, Figure 2 shows a group of Mira variables,
in which the emission peaks near 10 µm: they present the common property of
a long-term mid-IR variability that seems to be restricted to this special class of
sources. Moreover, Figure 3 shows two of the few available AGB sources observed
several times by ISO-SWS and confirms the same behavior: Mira variables present
an IR variability even over a relatively short time interval.
There is not yet an agreement on the origin and properties of this phenomenon.
4
250
200
150
100
50
)
y
J
(
y
t
i
s
o
n
i
m
u
L
0
0
10
300
250
200
150
100
50
)
y
J
(
y
t
i
s
o
n
i
m
u
L
0
0
10
Title : will be set by the publisher
RAFGL 190
RAFGL 809
350
300
250
200
150
100
50
)
y
J
(
y
t
i
s
o
n
i
m
u
L
40
50
0
0
10
300
200
100
)
y
J
(
y
t
i
s
o
n
i
m
u
L
40
50
0
0
10
ISO-SWS
MSX
IRAS-PSC
IRAS-LRS
TIRCAM2
2MASS
20
30
Wavelength (um)
RU Vir
20
30
Wavelength (um)
40
50
40
50
20
30
Wavelength (um)
IRC +60144
20
30
Wavelength (um)
Fig. 2. SEDs of sources that show significant variability over the time elapsed from the
IRAS to the TIRCAM2 observations. Data available from IRAS PSC, IRAS LRS, ISO-
SWS, MSX, TIRCAM2 and 2MASS are included. Only sources with maximum emission
in the range 8 − 20µm appear to be variable.
It needs to be examined in more detail to understand its nature.
It could be
perhaps a variability induced by shock waves caused by dynamic events in the
photosphere or by magneto-hydrodynamical modes (and magnetic storms). Oth-
erwise, it could be a "simpler" mid-IR variability, originating in the emission of the
circumstellar envelopes and caused by modulations in the efficiency of dust forma-
tion. This would be expected to be more typical of the Mira sources; Semiregulars
have thinner circumstellar envelopes, while Post-AGB stars should have detached
shells not strongly influenced by the variability of the central star.
The best way to examine this variability is that of observing AGB stars at
different epochs also in mid-IR and this task can be performed by a ground-
based telescope placed in Antarctica. Moreover, simultaneous observations in the
Guandalini et al.: Infrared Properties Of AGB Stars . . .
5
V Cyg
1800
1600
1400
1200
1000
800
600
400
200
)
y
J
(
y
t
i
i
s
o
n
m
u
L
0
0
900
800
700
600
500
400
300
200
100
)
y
J
(
y
t
i
i
s
o
n
m
u
L
0
0
ISO-SWS
''
''
''
''
''
MSX
IRAS PSC
IRAS LRS1
IRAS LRS2
2MASS
10
20
30
40
50
Wavelength (um)
R Scl
ISO-SWS
''
''
''
''
''
IRAS PSC
2MASS
10
20
30
40
50
Wavelength (um)
Fig. 3. SEDs of two sources observed several times with ISO-SWS. Only the source with
maximum emission in the 10µm region appears to be variable.
near- and mid-IR and correlations with optical variability could be fundamental
to understand these phenomena.
3 Conclusions: Why Surveys from Dome C
Important IR studies for AGB stars can be made in the best way from ground-
based locations like Dome C and Antarctica. The study of the main features
regarding IR variability (and also optical variability) is one of them. It can be ful-
filled with, surveys through wide field (in the future) or small area (IRAIT) imag-
ing of interesting stellar systems, multiple observations of chosen AGB sources at
different epochs and wavelengths and observation of AGB sources from Magellanic
Clouds, whose distance is well-estimated.
In this way we could obtain:
6
Title : will be set by the publisher
• light curves and therefore good knowledge of variability;
• an accurate study of the luminosity variations over a wide region of the
electromagnetic spectrum, including optical and IR;
• finally, a reliable comparison between observations and models of AGB stars.
References
Bergeat, J., & Chevallier, L. 2005, A&A, 429, 235
Busso, M., Gallino, R., & Wasserburg, G. J. 1999, ARA&A, 37, 239
Busso, M., Guandalini, R., Persi, P., Corcione, L., & Ferrari-Toniolo, M. 2006,
AJ, submitted
Guandalini, R., Busso, M., Ciprini, S., Silvestro, G., & Persi, P. 2006, A&A, 445,
1069
Guandalini, R., Busso, M., & Cardinali, M. 2007, in preparation
Sedlmayr, E. 1994, Lecture Notes in Physics, 428, 163
Straniero, O., Dom´ınguez, I., Cristallo, S., & Gallino, R. 2003, PASA 20, 389
Tosti, G. et al. 2007, this conference
Uttenthaler, S., Hron, J., Lebzelter, T., Busso, M., Schultheis, M., & Kaeufl, H.
U. 2006, astro-ph/0610500
Whitelock, P. A., Feast, M. W., Marang, F., & Groenewegen, M. A. T. 2006,
MNRAS, 369, 751
|
0711.0230 | 1 | 0711 | 2007-11-01T22:53:57 | The Effect of the ISM Model on Star Formation Properties in Galactic Discs | [
"astro-ph"
] | Modelling global disc galaxies is a difficult task which has previously resulted in the small scale physics of the interstellar medium being greatly simplified. In this talk, I compare simulations of galaxies with different ISM properties to determine the importance of the ISM structure in the star formation properties of the disc. | astro-ph | astro-ph |
**FULL TITLE**
ASP Conference Series, Vol. **VOLUME**, **YEAR OF PUBLICATION**
**NAMES OF EDITORS**
The Effect of the ISM Model on Star Formation
Properties in Galactic Discs
Elizabeth J. Tasker1 and Greg L. Bryan2
Abstract. Modelling global disc galaxies is a difficult task which has previ-
ously resulted in the small scale physics of the interstellar medium being greatly
simplified.
In this talk, I compare simulations of galaxies with different ISM
properties to determine the importance of the ISM structure in the star forma-
tion properties of the disc.
1.
Introduction
One of the main problems in understanding galaxy formation is that star forma-
tion is extremely complex. The gas out of which stars form is a turbulent mix
of forces in which gravitational collapse, thermal pressure, magnetic fields, cos-
mic ray pressure and energy from supernovae all fight for dominance (MacLow
2004).
This leaves galaxy simulators with a choice; either to model the interstellar
gas in detail but restrict their study to a small patch of the galaxy (e.g. Slyz et al.
2005) or to simulate the entire galaxy but use a greatly simplified model for the
ISM (e.g. Li et al. 2005). The former approach allows the inclusion of many
more of the important physical processes but is unable to tell us anything about
the global evolution of the disc. The latter, meanwhile, reveals properties of the
whole galaxy including star formation histories and global structures, but it is
impossible to judge the effect the simple ISM model is having on these results.
However, recent numerical simulations are now able to include a more
complex multiphase ISM in these global models (Tasker & Bryan 2007, 2006;
Wada & Norman 2007). This allows us to test the impact of modelling the ISM
in a more realistic way in galaxy formation. This has particular baring on cos-
mological simulations, where the resolution of the ISM of individual galaxies is
still beyond our reach.
2. Numerical method
Using the AMR code, Enzo (Bryan & Norman 1997), we compared three models
of isolated galaxy discs where we varied the properties of the interstellar gas.
In all cases, the simulations started with a rotating Milky Way-sized exponen-
tial disc of gas sitting in a static NFW dark matter potential. The set-up is
1Department of Astronomy, University of Florida, Gainesville, FL 32611
2Department of Astronomy, Columbia University, New York, NY 10027
1
2
Tasker and Bryan
Figure 1. Gas density projections at 388 Myrs for (left to right) ISM #1,
ISM #2 and ISM #3, all without feedback. Scale is to the base-10 logarithm
with units M⊙Mpc−2 and each image is 60 kpc across.
described in more detail in Tasker & Bryan (2006, 2007).
In our first galaxy
model (ISM #1), the gas was allowed to radiatively cool to 300 K. In our second
model (ISM #2), an additional background photoelectric heating term was in-
cluded while in our last disc (ISM #3), the ISM was kept at a fixed temperature
of 10,000 K, in keeping with previous simulations. All simulations included star
formation and ISM models 1 and 2 also included feedback from Type II SNe
(this was impossible to include for the third, isothermal, disc).
3. The structure of the ISM
Figure 1 shows projections of the gas density of each galaxy disc 377 Myrs after
the start of the simulation. At this time, the disc has undergone its initial
fragmentation (and subsequent star burst) and the evolution is now slow.
In
Figure 1, all images are for runs without stellar feedback. With our first ISM
model, shown on the far left, we see that the gas has fragmented out to a well-
defined radius. Beyond this point, there is still a significant amount of gas,
but it is stable to gravitational fragmentation and does not collapse to form
stars. When we include photoelectric heating in ISM #2 (middle image) we see
a notably different gas structure. In particular, there exists large voids of hot,
low density gas. The porous nature of this ISM has been observed both in our
own galaxy and (perhaps most dramatically) in the HI map of the LMC. This
result agrees with recent simulations of Wada et al. (2000), who suggest these
holes are not the results of SNe remnants, but rather the product of gravitational
and thermal instabilities. Our final ISM model, where the temperature is fixed,
shows another distinct structure. Due to the fixed high temperature of 104 K,
the Jean's Length is higher than in the other discs, causing the star-forming
knots of gas to be much higher in mass. This has the effect of producing very
large star clusters that are confined to the densest, inner regions of the disc.
Their formation also produces voids in the gas distribution, but this is due to
gas deficit, not to instabilities.
The pressure distribution of these discs is also interesting. Without feed-
back, discs with ISM #1 and ISM #2 are largely in pressure equilibrium, as
Effect of ISM model in galaxies
3
100
ISM #1
ISM #2
ISM #3
No feedback
Feedback
n
o
i
t
c
a
r
F
e
m
u
l
o
V
10-2
10-4
1e-06
1e-03
ρ [Msolarpc-3]
1e+00
1e-06
1e-03
ρ [Msolarpc-3]
1e+00
1e-06
1e-03
ρ [Msolarpc-3]
1e+00
Figure 2.
PDF of the volume weighted gas density at 377 Myrs for the three
ISM types (running left to right). A log normal fit is shown as a solid line in
the first panel.
predicted by the analytical models of McKee & Ostriker (1977). The isothermal
disc cannot be, since fixing the temperature requires the pressure to be propor-
tional to the gas. When feedback is introduced, this situation changes. SNe
energy causing streams of gas to be blown both in the disc's plane and off its
surface, creating a galactic fountain.
The structure of these ISMs can be examined more quantitatively by looking
at their 1D PDFs, as shown in Figure 2. Comparing all three of the ISMs across
the panels, we see a significant amount of substructure in the low density gas, but
the high density end of the PDFs are surprisingly uniform. At these densities,
all profiles are well fitted by a lognormal distribution. This remains true even
when feedback is included.
4. Observational comparison
The left-hand plots in Figure 3 shows the star formation history for each galaxy
disc. Without feedback, our ISM #1 disc converts all the available gas into
stars, halting further star formation. By contrast, the addition of background
heating quenches star formation by raising the temperature of the coldest gas
and allowing the disc to show the beginning of self-regulation. The addition of
feedback, however, is a much stronger effect, with the added energy destroying
the star forming knots and thereby increasing the available gas at later times.
The isothermal disc also shows a flattening in the star formation rate, but this is
likely due to the confinement of the star formation to the denser parts of the disc
which slows down the rate of gas consumption. We can also compare our results
with the widely observed Kennicutt-Schmidt law (Kennicutt 1989), as shown
in the right-hand plot of Figure 3. Here, we see that the observed gradient is
reproduced well in both the first two ISM types, but less well in our isothermal
disc. We do however, persistently overestimate the rate of star formation, even
when feedback is included. This is likely due to our star formation recipe and
is something to address in later work.
4
)
1
-
r
y
n
u
s
M
(
R
F
S
)
1
-
r
y
n
u
s
M
(
R
F
S
)
1
-
r
y
n
u
s
M
(
R
F
S
1000
100
10
1
0.1
1000
100
10
1
0.1
1000
100
10
1
0.1
0
Tasker and Bryan
No feedback
Feedback
No feedback
Feedback
No feedback
)
2
-
c
p
k
1
-
r
y
n
u
s
(
M
R
F
S
Σ
)
2
-
c
p
k
1
-
r
y
n
u
s
(
M
R
F
S
Σ
)
2
-
c
p
k
1
-
r
y
n
u
s
(
M
R
F
S
Σ
No feedback
Feedback
No feedback
Feedback
No feedback
100
10
1
0.1
0.01
0.001
100
10
1
0.1
0.01
0.001
100
10
1
0.1
0.01
0.001
0.2
0.4
0.6
0.8
1
1.2
0.1
1
t (Gyrs)
10
gas(Msunpc-2)
Σ
100
1000
Figure 3.
law (right) for ISM models 1, 2 and 3 (top to bottom).
The star formation history (left) and global Schmidt-Kennicutt
5. Conclusions
One of the most surprising results in this research is that despite significant
structural differences in the ISM, the star formation properties remain largely
unaffected. The reason for this can be seen in Figure 2, where we see that the
biggest differences between our three ISM types occur in the low to medium
density gas. The high density, star-forming gas, meanwhile, forms a consistent
lognormal profile in all cases. This result is largely positive; it suggests that
a decent subgrid model for star formation can be used where a detailed ISM
model is not possible, such as large-scale cosmological simulations. However,
it also implies that a through understanding of star-formation cannot be found
from the Kennicutt-Schmidt law, which is largely insensitive to the input physics.
This work is presented in greater detail in Tasker & Bryan (2007).
References
Bryan, G. L. & Norman, M. L. 1997, ASP Conf. Ser. 123: Computational Astrophysics;
12th Kingston Meeting on Theoretical Astrophysics, 363
Kennicutt, R. C. 1989, ApJ, 344, 685
Li, Y., Mac Low, M.-M., & Klessen, R. S. 2005, ApJ, 626, 823
MacLow, M.-M. 2004, Ap&SS, 289, 323
McKee, C. F., & Ostriker, J. P. 1977, ApJ, 218, 148
Slyz, A. D., Devriendt, J. E. G., Bryan, G., & Silk, J. 2005, MNRAS, 356, 737
Tasker, E. J., & Bryan, G. L. 2006, ApJ, 641, 878
Tasker, E. J., & Bryan, G. L.
2007, ArXiv Astrophysics e-prints, arXiv:astro-
ph/0709.1972
Wada, K., Spaans, M., & Kim, S. 2000, ApJ, 540, 797
Wada, K., & Norman, C. A. 2007, ApJ, 660, 276
|
astro-ph/0302040 | 1 | 0302 | 2003-02-03T22:36:05 | Are Compact High-Velocity Clouds Extragalactic Objects? | [
"astro-ph"
] | Compact high-velocity clouds (CHVCs) are the most distant of the HVCs in the Local Group model and would have HI volume densities of order 0.0003/cm^3. Clouds with these volume densities and the observed neutral hydrogen column densities will be largely ionized, even if exposed only to the extragalactic ionizing radiation field. Here we examine the implications of this process for models of CHVCs. We have modeled the ionization structure of spherical clouds (with and without dark matter halos) for a large range of densities and sizes, appropriate to CHVCs over the range of suggested distances, exposed to the extragalactic ionizing photon flux. Constant-density cloud models in which the CHVCs are at Local Group distances have total (ionized plus neutral) gas masses roughly 20-30 times larger than the neutral gas masses, implying that the gas mass alone of the observed population of CHVCs is about 40 billion solar masses. With a realistic (10:1) dark matter to gas mass ratio, the total mass in such CHVCs is a significant fraction of the dynamical mass of the Local Group, and their line widths would exceed the observed FWHM. Models with dark matter halos fare even more poorly; they must lie within approximately 200 kpc of the Galaxy. We show that exponential neutral hydrogen column density profiles are a natural consequence of an external source of ionizing photons, and argue that these profiles cannot be used to derive model-independent distances to the CHVCs. These results argue strongly that the CHVCs are not cosmological objects, and are instead associated with the Galactic halo. | astro-ph | astro-ph | Are Compact High-Velocity Clouds Extragalactic Objects?
Philip R. Maloney1 & Mary E. Putman2,3
Center for Astrophysics and Space Astronomy, University of Colorado
Boulder, CO 80309-0389
ABSTRACT
Compact high-velocity clouds (CHVCs) are the most distant of the HVCs in the Lo-
cal Group model and, at d ∼ 1 Mpc, they have HI volume densities of ∼ 3 × 10−4 cm−3.
Clouds with these volume densities and the observed column densities NHI ∼ 1019 cm−2
will be largely ionized, even if exposed only to the extragalactic ionizing radiation field.
Here we examine the implications of this process for models of CHVCs. We have mod-
eled the ionization structure of spherical clouds (with and without dark matter halos)
for a large range of densities and sizes, appropriate to CHVCs over the range of sug-
gested distances, exposed to an extragalactic ionizing photon flux φi ∼ 104 phot cm−2
s−1. Constant-density cloud models in which the CHVCs are at Local Group distances
have total (ionized plus neutral) gas masses ∼ 20 − 30 times larger than the neutral
gas masses, implying that the gas mass alone of the observed population of CHVCs is
∼ 4 × 1010 M⊙. With a realistic (10:1) dark matter to gas mass ratio, the total mass
in such CHVCs is a significant fraction of the dynamical mass of the Local Group, and
their line widths would greatly exceed the observed ∆V . Self-consistent models of gas
in dark matter halos fare even more poorly; they must lie within approximately 200 kpc
of the Galaxy, and (for a given distance) are much more massive than the correspond-
ing uniform density models. We also show that exponential neutral hydrogen column
density profiles are a natural consequence of an external source of ionizing photons, and
argue that these profiles cannot be used to derive model-independent distances to the
CHVCs. These results argue strongly that the CHVCs are not cosmological objects,
and are instead associated with the Galactic halo.
Subject headings: ISM: clouds -- ISM: H I -- Galaxy: halo -- Local Group
3
0
0
2
b
e
F
3
1
v
0
4
0
2
0
3
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1.
Introduction
The anomalous velocity clouds of neutral hydrogen known as the high-velocity clouds (HVCs)
may represent the continuing infall of matter onto the Local Group (e.g., Oort 1966; Verschuur
1969; Blitz et al. 1999). The distances to the majority of these clouds remains unknown, but in the
[email protected]
2Hubble Fellow
[email protected]
-- 2 --
Local Group scenario the distances would range from a few kpc (for gas currently accreting onto
the Galaxy) to beyond Andromeda (∼ 1 Mpc). HVCs range in size from ∼ 0.1 − 100 deg2, and it
is possible that the sizes reflect their distance from the Galaxy. In particular, some of the HVCs
are both compact and isolated from extended emission, and show many similar HI properties to
dwarf galaxies. These are referred to as compact HVCs (CHVCs) and may be the pristine building
blocks of the Local Group at ∼ 1 Mpc.
The survival and composition of these CHVCs against photoionization depends on their volume
densities (and therefore distances) and the strength of the extragalactic ionizing photon flux. The
presence of an extragalactic ionizing radiation field is inferred from several sets of observations.
Maloney (1993) used a deep H I observation of NGC 3198 (van Gorkom 1993), which shows a
sharply truncated neutral hydrogen disk at NHI ≈ 5 × 1019 cm−2, to infer that the flux in the
energy range hv = 13.6 to ∼ 200 eV is approximately φi = 104 phot cm−2 s−1. Estimates from
the proximity effect at low redshift (Kulkarni & Fall 1993; Scott et al. 2002), upper limits from
sensitive Hα observations (Madsen et al. 2001; Weymann et al. 2001), and theoretical estimates
based on known sources (Haardt & Madau 1996; Shull et al. 1999) are in agreement with a value
φi ∼ 0.5 − 1.5 × 104 phot cm−2 s−1.
In this paper we examine the ionization state of the compact high-velocity clouds at various
distances, subject to an extragalactic ionizing radiation field. In the original Local Group model,
the CHVCs are confined by dark matter halos, with Mdm/Mgas ≈ 10. In a more recent variant of
this model, it is suggested that the CHVCs are condensations in an intragroup medium, bound by
external pressure (Blitz, private communication). We therefore consider both constant gas density
models and dark matter-dominated models. The H I properties of the CHVCs are briefly reviewed
in section 2 and the photoionization model is described in section 3. The results of the models are
presented in section 4, and in section 5 we discuss the implications of the results.
2. HI Data
The H I properties of the CHVCs are taken from the H I Parkes All-Sky Survey (HIPASS)
HVC catalog of Putman et al. (2002). This is a catalog of high-velocity clouds with declination
δ < +2◦ and vLSR ≤ 500 km s−1, excluding emission with vLSR < 90 km s−1 to avoid confusion
with Galactic emission. The spatial resolution of the survey is 15.5′ and the velocity resolution is
26 km s−1. The 5σ column density sensitivity of the HVC HIPASS data is 2 × 1018 cm−2, which
we adopt as a cutoff. Barnes et al. (2001) give a full description of HIPASS. The original CHVC
catalog of Braun & Burton (1999) was based on data of lower resolution and sensitivity, and 50%
of the CHVCs in the region of overlap between the two catalogs were reclassified based on the
HIPASS data. In order for a high-velocity cloud to be considered a compact and isolated CHVC
in the HIPASS catalog, it must have a diameter at 25% of the peak column density less than 2◦,
and must not be elongated in the direction of any extended emission. The typical parameters of
a HIPASS CHVC are an angular size of 0.36 deg2 (to the sensitivity limit), a total H I flux of
19.9 Jy km s−1, and a typical peak column density of NHI ∼ 1019 cm−2. There are a total of 179
-- 3 --
CHVCs cataloged from HIPASS and when this is combined with the northern sky data from the
Leiden-Dwingeloo survey there are approximately 250 CHVCs in total (de Heij, Braun & Burton
2002). The linewidth distribution shows a prominent peak at a FWHM ∆V = 25 km s−1 (de Heij,
Braun & Burton).
For most of the results in this paper, we have used the typical CHVC parameters quoted
above. This is a reasonable assumption, since the distributions in column density, angular size,
and velocity are fairly sharply peaked. An increase or decrease of less than a factor of three in
column density away from the peak value produces a drop by a factor of five or more in the number
of clouds4, and similarly for the cloud areas (cf. Figures 12 and 13 of Putman et al. [2002]). The
linewidth distribution of de Heij et al. (2002) is less sharply peaked than the column density or
angular size distributions, and is skewed to the low-velocity end: 70% of the observed clouds have
a FWHM ∆V = 25 km s−1 or less; less than 10% have ∆V in excess of 42 km s−1. Despite the
well-defined typical values, in order to verify that our conclusions are robust to uncertainties in
the cloud parameters, we have also included models in which the peak central column density is
up to an order of magnitude larger than the typical CHVC value, and the linewidth is doubled to
∆V = 50 km s−1 (see §4.2.2).
3. Model
The photoionization models of CHVCs have been calculated with the code described by Mal-
oney (1993), modified to allow for the spherical geometry of the CHVCs. (Modest -- tens of percent
-- departures from sphericity will have no significant impact on the model results.) Spherical geome-
try raises the ionization rates substantially over plane-parallel models with the same total hydrogen
density; integration of the ionization rates over angle is done using 10-point Gaussian quadrature.
The photoionization equilibrium of a CHVC exposed to an isotropic background radiation field is
calculated iteratively, including the effects of the diffuse radiation. Thermal equilibrium at T ∼ 104
K is assumed, as is appropriate for gas in the warm neutral phase. An ionizing photon flux φi ≈ 104
phot cm−2 s−1 has been assumed; a flux a factor of two smaller would reduce the total gas masses
by a factor of ≈ 1.5 (see below). A radial grid with 100 depth points (with logarithmic spacing)
was used; we have verified that this provides adequate accuracy (less than 1% error in the total
column density for a fixed neutral column density). In all cases we have assumed that the only
ionizing photon flux is due to the extragalactic background; depending on the fraction of ionizing
photons that escape from the Milky Way, the Galactic ionizing radiation field could dominate for
CHVC distances d <
∼ 100 kpc (Maloney & Bland-Hawthorn 1999).
4Sensitivity starts to affect the column density distribution for column densities not too far below the peak, and
so the drop-off with decreasing column below the peak is not reliably established. However, since clouds with column
neutral hydrogen column densities smaller than the peak value will be even more highly ionized than clouds with our
typical parameters, the actual number of clouds at the low column density end is irrelevant for the purposes of this
paper.
-- 4 --
4. Results
As noted in the introduction, we have considered both models with uniform gas (total hydro-
gen) density nH and models where the gas is confined by dark matter potential wells. We first
discuss the constant nH models, as these contain much of the relevant physics, and then the models
that include dark matter.
4.1. Uniform density models
If the CHVCs were at local group distances (∼ 1 Mpc), their H I volume densities would be
nHI ∼ 10−4 cm−3. We therefore consider total hydrogen densities nH ranging from 10−4 to 10−2
cm−3; as we will see, the low-density end of this range is untenable for reasonable HVC distances.
The model results (total hydrogen columns, gas masses, etc.) are given in Table 1.
In Figure 1, we plot the column density of total hydrogen that is required to produce a neutral
hydrogen column NHI = 1019 cm−2 as a function of the total hydrogen density, when the model
CHVCs are exposed to our fiducial ionizing photon flux. The solid curve is for our fiducial ionizing
photon flux and the dashed curve is for φi reduced by a factor of two. At the low density end of
the range, the ratio NH/NHI is enormous, ∼ 100 − 200, in consequence of the small neutral fraction
of gas at this density even if exposed only to the extragalactic ionizing background. At the highest
densities, this ratio has declined to ≈ 2.5 − 3, as the gas is substantially neutral over much of the
total column.
Since we know both the neutral and total hydrogen column densities as well as the volume
densities, we can calculate the gas masses directly from the models. However, in order to compare
the results to observations, we need to assign distances to the model clouds. To do this, we adopt
the following procedure. For each model, we determine the radius at which the projected H I
column drops to the HIPASS sensitivity limit of NHI = 2 × 1018 cm−2. We then set the distance
by requiring that the apparent angular size (using this radius) matches the typical value found for
the CHVCs in the HIPASS survey, for which the angular radius ∆θ ∼ 0.34 degrees. This gives the
result dHVC = 169R18.3 kpc, where R18.3 is the model radius (in kpc) at the NHI = 2 × 1018 cm−2
level.
In Figure 2 we show both the total gas masses Mgas and the apparent masses MHI, the latter
derived from the neutral hydrogen column densities and the apparent (as measured in H I) sizes,
as a function of the cloud distance and total gas volume density, for our fiducial ionizing photon
flux. As is evident from Table 1, models in which the total gas density is less than nH ∼ 3 × 10−3
are ruled out, as such clouds are so large that they are required to lie outside of the Local Group,
and the gas masses become absurdly large. At the maximum acceptable distance (d ∼ 1 Mpc) the
total gas mass is an order of magnitude larger than the apparent H I mass.
Figure 3 shows the same quantities as in Figure 2, but we now show the results only for clouds
within the Local Group (d ≤ 1 Mpc). In addition, this figure also shows the gas masses for φi a
factor of two smaller than our assumed value. For the lower photon flux, clouds of a fixed density
are at smaller radius compared to the fiducial model. At a fixed distance, the total gas masses are
-- 5 --
smaller by about a factor of 1.5 for the reduced flux models. For model CHVCs at Local Group
distances (d >
∼ 0.5 Mpc), the gas masses are Mgas ∼ 107 − 108 M⊙.
There is an additional constraint that we can apply to the CHVC models. The linewidth
distribution of CHVCs is sharply peaked at a FWHM ∆V ≈ 25 km s−1 (de Heij, Braun, & Burton
2002). We have therefore calculated the expected line FWHM for the models (i.e., using the velocity
dispersion for a self-gravitating system), both with and without a dark matter component (in the
former case the total mass has simply been scaled up by the assumed ratio of dark to baryonic
mass); ∆V has been calculated assuming a uniform spherical mass distribution. In Figure 4 we
show the predicted line FWHM as a function of distance, for dark matter to gas mass ratios of ten
and zero. In the former case, the models are consistent with the observations for a distance d ∼ 330
kpc; with no dark matter the clouds must lie at approximately 1.5 Mpc. For φi reduced by a factor
of two, these distances increase somewhat, to d ∼ 430 kpc and 2.1 Mpc for Mdm/Mgas = 10 and
no dark matter, respectively.
While these models rule out the possibility that the CHVCs sample a population of cosmological
objects (i.e., gas in dark matter-dominated potential wells) at distances characteristic of the Local
Group (d ∼ 1 Mpc), they suggest that they could represent such objects at distances of ∼ 300 − 400
kpc. However, as we will see in the next section, models of CHVCs in realistic dark matter potential
wells require that they lie much closer to the Galaxy, with an upper limit to the distance of about
200 kpc. This distance scale rules out models in which the CHVCs represent continuing infall onto
the Local Group.
4.2. Dark matter halo models
Motivated by the order-of-magnitude agreement between the properties of model clouds with
Mdm/Mgas ∼ 10 and the observations for characteristic distances d ∼ 350 kpc, we have constructed
models of CHVCs using realistic dark matter potential wells. Numerical simulations of halo forma-
tion in cold dark matter cosmologies predict a specific form for the density profile, as initially found
by Navarro, Frenk, & White (1996). However, there is considerable disagreement as to whether this
profile describes the halos of real galaxies, particularly dwarf and low-surface brightness galaxies
(e.g., Flores & Primack 1994; Moore 1994; Burkert 1995). A number of authors have argued that
these latter objects have nearly constant-density cores rather than the cusped power-law NFW
density profile, but this point is controversial. We have therefore adopted two different models
for the halo profile. The first is the the Navarro, Frenk, & White (1996) halo density profile, but
modified to allow for the possible presence of a density core:
ρdm(r) = ρcrit
δcrs
(r + ro)(1 + r/rs)2
(1)
where ρcrit = 3H 2/8πG is the closure density, the scale length rs is related to the virial radius of
the halo by the concentration parameter c = Rvir/rs, ro is the core radius, and δc is a characteristic
density contrast, determined by the requirement that the halo mass
Mh(Rvir) = ∆ρcrit(4π/3)R3
vir ,
(2)
-- 6 --
where ∆ is the overdensity parameter. For a flat (Ω + Λ = 1) universe, ∆ is given by (e.g., Eke,
Navarro, & Frenk 1998) ∆ = 178Ω0.45 ≈ 111 at z = 0 for the Ωdm = 0.35 cosmology adopted here.
For the relevant mass range of halos, c ∼ 10.
In the limit of no core, the density contrast is given by the standard result (e.g., Eke, Navarro,
& Steinmetz 2001)
δc =
∆
3
c3
ln(1 + c) − c/(1 + c)
.
(3)
Since we have no physical basis for choosing the core radius (this point is discussed further below)
we have kept the concentration parameter fixed (for a given halo mass) at the no-core value, and
calculated δc from the requirement that the integral of equation (1) to Rvir equal the mass given
by equation (2). The NFW halo parameters (c, Rvir) for a given mass have been calculated using
the results of Mo, Mao & White (1998) and Eke et al. (2001).
The second halo mass profile is that suggested by Burkert (1995) based on observations of
dwarf galaxies,
ρdm(r) =
ρor3
o
(r + ro)(r2 + r2
o)
(4)
where ρo is the core density and ro is again the core radius. Like the NFW profile, the density of
the Burkert halo falls as r−3 at large radius. Burkert found that the core density and radius were
tightly correlated for the observed galaxies, so that the dark matter profiles are described by only
one free parameter, which can be taken to be ro. Burkert also concluded that the halo virial radius
Rvir ≈ 3.4ro, so that these halos are much less centrally concentrated than the NFW halos.
We have calculated the gas density profile within these halos, including the effects of the gas
self-gravity. The dark matter density profile is fixed, however: we do not include the response of
the halo to the gas. (This would only strengthen our conclusions on the viability of models with
dark matter.) In most cases, the gas velocity dispersion is taken to be σg = 10.6 km s−1, which
produces a line FWHM equal to the typical observed value ∆V = 25 km s−1, and is the expected
value for gas at T ∼ 104 K, characteristic of the warm neutral medium. For NFW halos, in the
case of no core and negligible gas self-gravity, the gas density profile is given by
where r = r/rs and
ρg(r) = ρg(0)eCh[ln(1+r)/r−1]
Ch =
4πGρcritδcr2
s
σ2
g
= 7.34 × 10−3 δcr2
s
σ2
g
(5)
(6)
for rs in kpc and σg in km s−1. For Burkert halos the gas profile is not analytic even in the absence
of self-gravity.
We first discuss the results of models with NFW halos, and then the Burkert halo models.
The uniform density models that are consistent with a dark matter to baryon ratio of about
10 have total hydrogen volume densities of approximately 3 × 10−3 cm−3 and total gas masses of
4.2.1. NFW halo models
-- 7 --
about Mgas ≈ 1.4 × 107 M⊙. We have therefore calculated the gas distribution and the neutral
fraction for a halo with a mass Mdm = 1.4 × 108 M⊙. The virial radius of this halo is 13 kpc and
the concentration parameter c = 12.44. We have also allowed for tidal truncation of the halo due
to the Galaxy's tidal field; this introduces a cutoff at a radius of 10.2 kpc.
In Figure 5 we show the gas density profile for the model that satisfies the requirement that
the peak NHI = 1019 cm−2. This profile (which is characteristic of all the halo models) immediately
reveals the problem that plagues all of these models. Most of the gas mass in the halo is at large
radius, where the density is low and therefore the neutral fraction is very small. Hence, in order
to match the observed neutral column density, the density at the center of the halo must be raised
substantially over that in the uniform density model. As Figure 6 (which plots the cumulative dark
matter and gas masses as a function of radius) demonstrates, this large central gas density results
in a gas to dark matter mass ratio of approximately unity within the halo, which is obviously
untenable. This conclusion is unaltered by reducing the ionizing flux by a factor of two or by
allowing the dark matter profile to have a core. Although this latter modification results in gas
density profiles that are more nearly constant over a larger range in radius, and therefore produce
the same total hydrogen column density for a smaller central density, the slower drop-off of nH with
r due to the core means that these models always have a higher value of Mgas/Mdm at the outer
boundary than the no-core models. (In fact, the models with cores invariably require larger total
hydrogen columns than the models without cores, which makes the problem even worse.)
The only way to produce reasonable baryon to dark matter ratios in these models is to increase
the mass of the halo. In order to produce a model in which the ratio Mgas/Mdm is less than about
0.15 at the outer boundary prior to tidal truncation, the halo mass must be Mdm
∼ 5 × 108 M⊙.
>
The mass increase is smaller than the factor of ten one might expect from Figure 6 because higher
mass halos are both larger than lower mass ones and denser at the same physical radius, and
therefore the observed neutral hydrogen column can be produced with a smaller total gas mass
than is required in the lower mass halo.
In addition to the mass problem, to which we return below, these NFW dark matter-dominated
models suffer from another, severe problem for the cosmological hypothesis: the physical size of
the neutral hydrogen distribution is so small that the model CHVCs are restricted to lie less than
d ∼ 200 kpc from the Galaxy. This is illustrated in Figure 7, which shows the projected H I column
as a function of impact parameter for a variety of halo masses, core radii, and ionizing fluxes. In
no case does the model CHVC radius at the HIPASS threshold, R18.3, significantly exceed 1 kpc. As
discussed in §4.1, this constrains the clouds to distances of no more than about 200 kpc from the
Galaxy. In fact, except for the lowest mass (Mdm = 1.4 × 108 M⊙) model shown in Figure 7, which,
as discussed above, has an unacceptably large gas to dark matter mass ratio, all of the models have
<
∼ 0.5 kpc unless we include a core in the halo density distribution. As we noted earlier, we
R18.3
have no physical basis for choosing the halo core parameters, and have included a core simply to
5Assuming that galaxy clusters provide a fair sample of the universe, the baryon to dark matter mass ratio is
estimated to be 0.13 for Ho = 70 km s−1 Mpc−1 (Mohr, Mathiesen, & Evrard 1999).
-- 8 --
see whether, given the uncertainties in the actual halo dark matter profiles, this could alleviate the
problems arising from the small distance scale mandated by the small physical size of the model
CHVCs. As is evident from Figure 7, it does not. The reason is simply that if the halo is massive
enough that the observed neutral hydrogen column density can be reproduced with a reasonable
baryon to dark matter mass ratio, the velocity dispersion of the baryonic component (which is
fixed by the observed linewidth) is only a fraction of the velocity dispersion characterizing the dark
matter potential well, with the result that the gas is confined to the core of the halo. Increasing
the size of the core beyond ro ≈ rs does not lead to any significant further increase in R18.3, as is
seen in the right-hand panel of Figure 7.
4.2.2. Burkert halo models
In Figure 8 we show results for models with a Burkert halo of mass Mdm = 108 M⊙. For
these models we have also explored a larger region of parameters for the CHVCs, considering
neutral hydrogen columns up to ten times larger than the typical value seen in the HIPASS survey
and linewidths up to a factor of two larger than the typical value. The left-hand panel assumes
the standard velocity dispersion of σg = 10.6 km s−1. From left to right, we plot the neutral
hydrogen column as a function of impact parameter for peak neutral hydrogen column densities
NHI = 1019 cm−2 (solid line), NHI = 2.5 × 1019 cm−2 (long-dashed line), NHI = 5.0 × 1019 cm−2
(dotted line), NHI = 7.5 × 1019 cm−2 (short-dashed line) and NHI = 1020 cm−2 (solid line). This
Burkert halo model does not suffer from the unacceptably high baryon to dark matter mass ratio
problem that afflicts the NFW halos of comparable mass, because the Burkert halo is physically
much smaller (virial radius of 1.8 kpc compared to 11.7 kpc for the same mass NFW model) and
therefore its mean density is much higher than the corresponding NFW halo. However, this small
size scale becomes a serious problem just as it does for the NFW halos. Even for a peak column
NHI = 1020 cm−2, the projected size at the HIPASS sensitivity threshold barely reaches 1 kpc,
again implying that the halos are no more than 200 kpc distant. The right-hand panel shows the
same models, only for a doubled velocity dispersion, so that the line FWHM ∆V = 50 km s−1.
Even with this assumption, the physical size of the clouds remains small: raising the peak column
density to NHI = 5.0 × 1019 - five times the typical HIPASS value - only increases the cloud radius
to about 1.6 kpc, which would push the clouds out to no more than 270 kpc. (Lower mass halos
- e.g., 107 M⊙ - with Burkert profiles are so small that they can lie no further than d ∼ 100 kpc
from the Milky Way, independent of total column density.)
In Figure 9 we present similar results, but for a halo with Mdm = 109 M⊙. In the left-hand
panel, with a velocity dispersion corresponding to the typical observed value, the results scarcely
differ from those for the 108 M⊙ halo, and the cloud size at the HIPASS threshold again never
exceeds 1 kpc. The results for the models with twice the linewidth are also very similar to those
for the lower mass Burkert halo. The chief difference is that, since the 109 M⊙ halo is physically
larger (by a factor of 2.7) than the 108 M⊙ halo, the highest column density models are larger in
the higher mass model, since the column density cutoff is not set by the physical edge of the halo
-- 9 --
as it is in the 108 M⊙ model.
Hence the CHVC models with Burkert halos suffer from the same distance problem as the
NFW halo models: the sizes at the HIPASS sensitivity threshold are simply too small to allow
them to be placed at distances much larger than d ≈ 200 kpc from the Milky Way. Our results are
in excellent agreement with those of Kepner, Babul & Spergel (1997) and Kepner et al. (1999), who
calculated the chemical, thermal, and ionization equilibrium of gas in dark matter minihalos with
Burkert density profiles. They found that the characteristic radius at the NHI ≈ 1018 cm−2 level
never exceeded r ≈ 1 kpc, even when the central hydrogen column density reached values much
larger than any we consider here.
The combination of these two factors (the large mass that is required in order to have plausible
gas to dark matter ratios for NFW halos, and the close distances required by the small physical
size of the clouds for either NFW or Burkert halos) is fatal for any model in which the CHVCs
represent a population of cosmological objects. In Figure 10 we show the halo mass distribution
for a cosmological simulation of the Local Group (Moore et al. 2001). The ∼ 2000 halos were
created by analyzing the formation of a binary pair of massive halos in a hierarchical universe
dominated by cold dark matter. The masses, separation and relative velocities of the binary halos
are close to those observed for the Milky Way and Andromeda. The binary system was chosen from
a large cosmological simulation such that a nearby massive cluster similar to Virgo was present.
The simulation is described in detail in Moore et al. (2001). The filled circles show all of the halos
(aside from those representing the Milky Way and Andromeda galaxies) within the simulated Local
Group volume and the open circles show the halos that lie within 200 kpc of the Galaxy. As we
showed above, the minimum acceptable mass for CHVC halos is Mdm ∼ 5 × 108 M⊙. From this
mass limit up to Mdm = 2 × 109 M⊙, there are 16 halos in the simulation6 To explain the CHVCs
as cosmological, dark matter-dominated objects, we need approximately 250. Hence unless the
hierarchical, cold dark matter-dominated cosmology grossly under-predicts the number of halos in
this mass range, our results do not allow CHVCs to represent such a population.
4.3. CHVC HI distribution
There have now been several detailed studies of the HI structure of CHVCs. Wakker & Schwarz
(1991) investigated two CHVCs and were the first to notice the core/halo structure of the clouds.
Braun & Burton (2000) imaged six CHVCs at 1′ resolution and found compact cores with linewidths
as low as 2 - 10 km s−1, as well as the large diffuse halo modeled here with linewidths of ∼ 25 km
s−1. The warm (∼ 104 K) halos were imaged at 3′ resolution with Arecibo by Burton, Braun &
6For a lower mass limit of Mdm ∼ 108 M⊙, as allowed for Burkert profile halos, the number of halos would be a few
times larger. However, except for extreme assumed CHVC parameters (i.e., peak column densities and linewidths
much larger than the typical values), the CHVC models with Burkert halos will be physically even smaller than those
with NFW halos. Hence these objects would have to lie closer than 200 kpc, and so the total number predicted would
be about the same as for the NFW-halo models.
-- 10 --
Chengalur (2001; hereafter BBC).7 They argued that the column density distribution at the edges
of the clouds (for NHI < 1018.5 cm−2) drops off as an exponential with radius, indicating a spherical
exponential distribution of neutral hydrogen volume density as a function of radius. (Figures 11
and 12 of BBC demonstrate the quality of the exponential fits.) Kinematically the cores and halos
of the CHVCs do not appear to be related. BBC derived distances of typically a few hundred kpc
for the clouds in their sample, based on the following assumptions:
• The neutral gas density distribution is reasonably well-described by an exponential with
radius, so that the central density can be inferred from the peak column density and the fit
scalelength;
• The gas neutral fraction is close to unity in the center of the CHVC, and the temperature
T ∼ 104 K;
• The total gas pressure at the center is P/k ∼ 100 cm−3 K, which Braun & Burton (2000)
argued is the characteristic pressure at which the cold cores can co-exist with the warm neutral
medium.
With these assumptions, the angular scale length of the exponential can be converted to a physical
scale length, thereby yielding a cloud distance.
In this section we show that the characteristic exponential NHI profile arises as a simple
consequence of the physics of photoionization, and in the absence of fine-tuning of conditions, is
very unlikely to provide reliable distance estimates.
In Figure 11 we plot the projected neutral hydrogen column density as a function of impact
parameter (normalized to the cloud radius) for five different uniform density models, with total
densities nH between 10−4 and 10−2 cm−2. For all of these models the NHI profile resembles
an exponential over most of the cloud radius, with the exception of a core which becomes more
pronounced with increasing density, and of a cutoff as the cloud outer boundary is approached. The
exponential appearance of the H I column density distribution indicates that the neutral hydrogen
volume density distribution is also approximately an exponential, at least over a substantial fraction
of the cloud volume. This is simply a consequence of the photoionization physics, and reflects the
attenuation of the ionizing radiation (and therefore the increase in the neutral fraction) with depth
into the clouds; the total gas volume densities in all these models are uniform.
7Of the ten CHVCs imaged by BBC at Arecibo, 8 are covered by the northern extension of the HIPASS survey. We
have examined the HIPASS data for these clouds to see whether the HIPASS results have been seriously compromised
by resolution effects. With the exception of the faintest cloud in the sample (for which the Arecibo peak column is
about 4.5 times the HIPASS value: however, the column density peaks are due to probably unresolved, non-centered
structure, and the entire western half of the cloud has a typical column extremely close to the HIPASS value), the
Arecibo and HIPASS peak column densities differ by at most a factor of two. Hence there is no reason to think that
the HIPASS observations systematically underestimate the cloud column densities significantly enough to affect the
results of this paper.
-- 11 --
BBC fit NHI profiles of the form
NHI(p) = 2noh(cid:16) p
h(cid:17) K1(p/h)
(7)
to their observations, where p is the impact parameter, no is the central density, h is the scale-
length, and K1 is a modified Bessel function. This is the form expected for an exponential volume
density distribution of infinite extent (van der Kruit & Searle 1981). This is modified for a cloud
of finite extent, however, and so instead we use the finite-cloud profile
NHI(p) = 2noh(cid:16) p
h(cid:17)he−R/h(cid:0)R2/p2 − 1(cid:1)1/2
+ k1 (p/h, R/p)i
where R is the cloud radius and k1 is an incomplete Bessel function,
k1(ρ, x) = ρZ x
1
e−ρt(cid:0)t2 − 1(cid:1)1/2
dt
(8)
(9)
which goes to the usual K1(ρ) as x → ∞.
In Figure 12 we show two examples of fits of equation (8) to the projected neutral hydrogen
column density distributions for the CHVC models described in §4.1. These are not formal fits to
the distributions, and are intended solely to show that these uniform density models are very well
described over a large range in radius (in fact, everywhere except for the cores) by an exponential
nHI distribution. Although BBC argued based on observations of dwarf galaxies that there is a
characteristic value of h, that conclusion does not hold for these models, as h depends nontrivially
on the volume density. For the pair of models shown in Figure 12, which differ in nH by a factor of
two, the fitted values of h differ by nearly a factor of four.
The BBC distance method generally yields erroneous results when applied to these models,
for two reasons: (1) many of the CHVC models have substantial ionized fractions even in their
cores, so that the central neutral density - even if determined with reasonable accuracy from the
exponential fit - can be substantially below the actual total gas density, and (2) the central pressure
is generally not equal to P/k ∼ 100 cm−3 K. 8 It is not at all obvious that the value of P/k ∼ 100
quoted by Braun & Burton (2000) for the co-existence of cold gas with the warm neutral medium
is generally applicable. This conclusion was based on a phase diagram calculated by Wolfire et al.
(priv. comm.) for "Local Group conditions", which are not otherwise specified. In the calculations
of Wolfire et al. (1995a,b), which presumably use similar physics, gas heating is dominated by
grain photoelectric heating, and the only source of ionizing photons is the extragalactic soft X-ray
background, which produces an ionizing flux about an order of magnitude smaller than assumed
here. Furthermore, the resulting phase diagram is rather sensitive to the adopted gas parameters
8For the nH = 0.01 cm−3 model, for which P/K does equal 100 at cloud center, the BBC distance method gives
a result within about 25% of the value assigned to the cloud (D = 67 kpc, Table 1) in order to make its angular
size match the typical HIPASS size, provided that the true central column density is used, rather than the fitted
column density, since in these relatively high-density models the presence of a core means that the exponential fit
overestimates the central column density - cf. Figures 11 and 12.
-- 12 --
(see Figures 1 and 6 in Wolfire et al. 1995b, in which the range of pressures over which a two-phase
medium is allowed can extend over three orders of magnitude).
In fact, in the phase diagram
presented by Braun & Burton (their Figure 13), the minimum pressure Pmin for which the cold
phase exists is approximately five times larger (Pmin/k ≈ 300 cm−3 K) for a CHVC warm gas
column density NHI = 1019 cm−2, as appropriate for the HIPASS CHVCs discussed here, than it
is for NHI = 1020 cm−2 (for which Pmin/k is actually about 60), as adopted by Braun & Burton
for their sample. This would reduce the inferred distances by a factor of three. It is also not yet
clearly established that the presence of cold gas is a generic property of CHVCs: there are only
six clouds in the Braun & Burton (2000) sample, and while all of these exhibit cold cores, a much
larger sample is clearly needed. The Magellanic Stream shows no evidence for cold gas (Mebold et
al. 1991).
Interestingly, the models in which the gas is confined by a dark matter potential are not well
described by profiles of the form (7) or (8). We show one such NFW model in Figure 13. This
is the Mdm = 5 × 108 M⊙, no-core model shown as the solid line in the center panel of Figure 7.
Inspection of that figure shows that the shape of the neutral hydrogen column density distribution
in these dark matter-dominated models is generic. As is evident from Figure 13, these models
cannot be described by a single component of the form (7) or (8). This is because the neutral
hydrogen column density always exhibits a core with an extended wing, in consequence of the gas
density profiles in these halos (cf. Figure 5). Hence the profile fit to the core of the NHI distribution
(we have used equation (7) for these models, since the models are actually much larger than the
plotted region, and hence the cutoff is unimportant here), shown by the dashed line in Figure 13,
grossly under-predicts NHI outside r ∼ 0.5 kpc, while the fit to the column density distribution
at larger radius (dotted line) falls an order of magnitude short of NHI at r = 0. The results for
Burkert profile halos are similar; in Figure 14 we show the neutral column density distribution
for the largest column density model (NHI(0) = 1.0 × 1020 cm−2) shown in the lefthand panel of
Figure 8 (Mdm = 108 M⊙, σg = 10.6 km s−1). As for the NFW halo model, there is no reasonable
fit with a single component of the form (7) or (8); in fact, the core-halo structure is even more
pronounced than in the NFW halo. As before, fits to the core of the NHI distribution (here we
had to use equation [8]) severely underestimate NHI beyond r ∼ 0.8 kpc, while fits to NHI at large
radius underestimate the column density at small radii by an order of magnitude.
5. Summary and Discussion
Models in which the compact high-velocity clouds (CHVCs) lie at distances of hundreds of
kpc to ∼1 Mpc from the Galaxy have such low volume densities of hydrogen that exposure to the
extragalactic ionizing background radiation alone will largely ionize them. For models of uniform
density clouds (§4.1), we find that the total hydrogen column needed to produce the typical neutral
hydrogen column of NHI= 1019 cm−2 ranges between 3 and 200 times NHI (Figure 1) as the total
hydrogen density varies between 10−2 and 10−4 cm−3. The total gas masses of these clouds are
therefore much larger than the apparent neutral gas masses. To match the typical angular size
-- 13 --
seen for CHVCs in the HIPASS survey, low-density clouds must be at large distances and have
correspondingly large masses. Models in which the cloud hydrogen densities are less than nH ≈
2 − 3 × 10−3 cm−3 are ruled out, as such clouds would have to lie outside the Local Group (cf.
Figures 2 and 3).
At Local Group distances, d ∼ 0.7 − 1 Mpc, the gas mass alone of the individual CHVCs would
be Mgas ∼ 108 M⊙, and the mass of the observed population would be M ∼ 4 × 1010 M⊙. Such
CHVCs could have dark matter/gas mass ratios of only ∼ 2 − 3 without producing line widths
larger than the observed value. Models in which the CHVCs are at smaller distances, d ∼ 350
kpc, with gas masses Mgas ∼ 107 M⊙, appear to be more reasonable, and could be dynamically
compatible with plausible (∼ 10 to 1) dark matter to gas mass ratios.
∼ 1 kpc), which requires that the CHVCs are at distances d <
A closer examination of this scenario, however, using realistic dark matter halos, shows that
it runs into severe difficulties (§4.2). This is because: (1) the sizes of the detectable H I clouds are
too small (R <
∼ 200 kpc (this is true
whether we use NFW profiles or Burkert profiles for the halo dark matter density distribution);
and (2) the halo masses required for reasonable ratios of Mdm/Mgas for NFW halos are too large
(Mdm ∼ 109 M⊙). Hence the predicted number of such objects in a hierarchical CDM structure
formation scenario is more than an order of magnitude smaller than the observed number of CHVCs,
and CHVCs cannot represent a population of infalling, cosmological objects.
We have also examined the suggestion that the approximately exponential NHI profiles seen
in high resolution studies of a sample of CHVCs could provide an estimate of the cloud distances
(§4.3). Such profiles are a natural consequence of an external source of ionizing photons for clouds
with uniform total density. The derived scale lengths are not constant and depend on the cloud
density. We argue that this technique cannot be used to derive distances to the CHVCs in a model-
independent way, as the scale lengths are not fixed and the alternative assumption of a constant
pressure for the CHVC population is debatable. The CHVC NHI profiles can be explained by
photoionization even if the clouds are near to the Galaxy (d <∼ 100 kpc).
Our models have assumed that the gas is smoothly distributed, and therefore do not include
substantial clumping. This is arguably a plausible assumption for the CHVC envelopes that we
have modeled in this paper, as the gas in these envelopes presumably is in the warm (T ∼ 104 K,
as is indicated by the observed linewidths) neutral phase, and is supported by the generally smooth
distributions and reflection symmetry seen in high-resolution studies of the envelopes (BBC). The
only way in which this assumption can be substantially in error, and the gas highly clumped, is
if the clumps are in the cold neutral phase, with temperatures of order 100 K. In this case the
observed linewidths must reflect bulk motion rather than thermal velocities. Since the (thermal)
linewidths of individual clumps would be ∆V ∼ 2 km s−1, a large number of clumps would need
to be present along any given sightline in order to explain the smoothness of the velocity profiles.
Such clumps would also have to be substantially smaller than the spatial resolution of the high
resolution imaging studies, which frequently do find (and resolve) small numbers of cores with
narrow linewidths (e.g., BBC). If pressures P/k ∼ 100 cm−3 K are necessary to drive the gas into
the cold neutral phase (BBC), then the hydrogen densities must be at least nH ∼ 1 cm−3.
It
-- 14 --
is rather difficult to see how to make such a density structure compatible with the observations.
If the clumps are self-gravitating, then they must have column densities that are much larger
than the beam-averaged column densities measured observationally, which requires that the areal
filling factor is small even though the filling factor in velocity space must be large to reproduce
the observed line profiles. If they are not self-gravitating, we are left with the question of what
confines them (or generates them, if they are transient objects). Hence we consider it unlikely that
the CHVC envelopes that we have considered in this paper actually consist of dense, cold clumps
rather than relatively smooth warm neutral gas.
What possibilities does this leave us with for the nature of CHVCs? Clouds without dark
matter located within the confines of the Local Group are not massive enough to be self-gravitating,
and must be either transient or confined by some external medium. Either of these scenarios would
seem to argue for the CHVCs being associated with the Galaxy rather than the Local Group,
as both the relatively short dynamical timescale (t ∼ 108 years) and the presence of a confining
medium are much more easily understood if the CHVCs are connected to the flux of mass and energy
through the Galactic halo. This suggestion is supported by the overall similarities of the CHVCs
to the remainder of the HVC population (Putman et al. 2002). The total and neutral hydrogen
gas masses of the entire observed CHVC population (∼ 250 objects) would be Mgas ∼ 7 × 108 M⊙
and MHI ∼ 7 × 107 M⊙, respectively, at d ≈ 200 kpc, and scale nearly as d2 for smaller distances.
We are grateful to several colleagues, especially Nick Gnedin and Jessica Rosenberg, for help-
ful discussions. Nick kindly provided the software (written by Oleg Gnedin) for calculating the
properties of NFW halos. PRM is supported by the National Science Foundation under grant AST
99-00871 and by NASA through HST grant AR-08747.02-A. MEP is supported by NASA through
Hubble Fellowship grant HST-HF-01132.01-A awarded by the Space Telescope Science Institute,
which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA,
under contract NAS 5-26555.
Barnes D., Staveley-Smith L., et al. 2001, MNRAS, 322, 486
REFERENCES
Blitz L., Spergel D., Teuben P., Hartmann D., & Burton W.B. 1999, ApJ, 514, 818
Braun R. & Burton W.B., 1999, A&A, 341, 437
Braun R. & Burton W.B. 2000, A&A, 354, 853
Burton W.B., Braun R. & Chengalur J. 2001, A&A, 369, 616
de Heij V., Braun R. & Burton W.B. 2002, A&A, 392, 417
Haardt, F., & Madau, P. 1996, ApJ, 461, 20
Kulkarni, V.P., & Fall, S.M. 1993, ApJ, 413, L63
-- 15 --
Madsen, G.J., Reynolds, R.J., Haffner, L.M., Tufte, S.L., & Maloney, P.R. 2001, ApJ, 560, L135
Maloney P., 1993, ApJ, 414, 41
Maloney, P.R., & Bland-Hawthorn, J. 1999, in Stromlo Workshop on High-Velocity Clouds, ed.
B.K. Gibson & M.E. Putnam (San Francisco: Astronomical Society of the Pacific), 199
Mebold, U., Herbstmeier, U., Kalberla, P.M.W., Greisen, E.W., Wilson, W., & Haynes, R.F. 1991,
A&A, 251, L1
Moore, B., Calc´aneo-Rold´an, C., Stadel, J., Quinn, T., Lake, G., Sebastiano, G., & Governato, F.
2001, PhysRevD, 64, 063508
Mohr, J.J., Mathiesen, B., & Evrard, A.E. 1999, ApJ, 517, 627
Oort, J.H., 1966, Bull. Astron. Inst. Netherlands, 18, 421
Putman M.E., et al., 2002, AJ, 123, 873
Scott, J., Bechtold, J., Morita, M., Dobrzycki, A., & Kulkarni, V.P. 2002, ApJ, in press.
Shull, J.M., Roberts, D., Giroux, M.L., Penton, S.V., & Fardal, M.A. 1999, AJ, 118, 1450
van der Kruit, P.C., & Searle, L. 1981, A&A, 95, 105
van Gorkom, J.H. 1993, in The Environment and Evolution of Galaxies, ed. J.M. Shull & H.A.
Thronson (Dordrecht: Kluwer), 343
Verschuur, G.L. 1969, ApJ, 156, 771
Wakker, B.P. & Schwarz, U. 1991, A&A, 250, 484
Weymann, R.J., Vogel, S.N., Veilleux, S., & Epps, H. 2001, ApJ, 561, 559
Wolfire, M.G., Hollenbach, D., McKee, C.F., Tielens, A.G.G.M., & Bakes, E.L.O. 1995a, ApJ, 443,
152
Wolfire, M.G., McKee, C.F., Hollenbach, D., & Tielens, A.G.G.M. 1995b, ApJ, 453, 673
This preprint was prepared with the AAS LATEX macros v5.0.
-- 16 --
Fig. 1. -- The total hydrogen column density NH required to obtain a neutral hydrogen column
density NHI= 1019 cm−2, as a function of the total hydrogen density. The solid curve is for an
ionizing flux φ ∼ 104 phot cm−2 s−1; the dashed curve shows the effect of reducing φ by a factor
of two. In the latter case the values of NH are smaller by a factor of two at the low-density end
compared to the fiducial model.
-- 17 --
Fig. 2. -- The total (neutral plus ionized) and apparent (derived from the neutral column density
and apparent size) gas masses, plotted vs. distance and total H density. For each model (i.e.,
assumed total hydrogen density), the distance d is determined from the requirement that the
apparent angular size (determined using the radius for which the projected neutral hydrogen column
drops to the HIPASS sensitivity limit, NHI= 2 × 1018 cm−2) matches the typical size seen in the
HIPASS sample of CHVCs, for which the angular radius ∆θ ∼ 0.34 degrees. The masses as a
function of density (top axis) are plotted as dashed lines, and as a function of distance (bottom
axis) as solid lines; the true gas mass Mgas is always greater than the apparent mass MHI.
-- 18 --
Fig. 3. -- The total and apparent gas masses, plotted vs. distance and total H density, as in Figure
2. The left panel shows the results for our fiducial ionizing flux φi ∼ 104; the right panel is for φi
reduced by a factor of two.
-- 19 --
Fig. 4. -- Predicted line FWHM vs. distance, for dark matter to gas mass ratios of 10 and 0, for the
model CHVCs. As in Figure 2, the distance is determined by requiring that the apparent angular
size equals the typical CHVC angular size seen in the HIPASS survey. The observed line width,
∆V ∼ 25 km s−1, is shown as the long-dashed line. The models are consistent at d ∼ 330 kpc with
Mdm/Mgas = 10 (dotted line) and at 1.5 Mpc with no dark matter (solid line). For the reduced
φi models (not shown), the intercepts are at d ∼ 430 kpc and d ∼ 2.1 Mpc for 10:1 and no dark
matter, respectively.
-- 20 --
Fig. 5. -- The gas density distribution in a NFW dark matter halo of mass Mdm = 1.4 × 108 M⊙
(solid line); the dashed line shows the density profile in the limit of negligible gas self-gravity.
-- 21 --
Fig. 6. -- The cumulative dark matter and gas masses for the halo shown in Figure 5. The ratio
Mgas/Mdm ≈ 1 for this model.
-- 22 --
Fig. 7. -- The projected neutral hydrogen column as a function of impact parameter, for several
different NFW halo models. From left to right, the panels are for halo dark matter masses Mdm =
1.4×108 M⊙, 5.0×108 M⊙, and 1.0×109 M⊙, respectively. The dotted line running across the panels
shows the HIPASS column density threshold of NHI = 2 × 1018 cm−2. Several models are shown in
each panel. In the lefthand panel (Mdm = 1.4 × 108 M⊙), the solid curve is for our fiducial ionizing
flux φo, and the long-dashed curve is for φi = φo/2. In the center panel (Mdm = 5.0 × 108 M⊙),
the solid line is for φi = φo, the short-dashed line is for φi = φo/2, and the long-dashed line is
for φi = φo and a halo core radius ro = rs. In the righthand panel (Mdm = 1.0 × 109 M⊙), all
models have φi = φo, and the solid, long-dashed, and short-dashed lines are for ro = 0, rs, and 2rs,
respectively.
-- 23 --
Fig. 8. -- The projected neutral hydrogen column as a function of impact parameter, for Burkert
halo models. The halo mass is Mdm = 1.0 × 108 M⊙. From left to right, the curves are for central
neutral hydrogen column densities NHI(0) = 1.0 × 1019, 2.5 × 1019, 5.0 × 1019, 7.5 × 1019, and
1.0 × 1020 cm−2, respectively. In the lefthand panel a velocity dispersion of σg = 10.6 km s−1 has
been assumed, while in the righthand panel σg = 21.2 km s−1. The dotted line running across the
panels shows the HIPASS column density threshold of NHI = 2 × 1018 cm−2.
-- 24 --
Fig. 9. -- The projected neutral hydrogen column as a function of impact parameter, for Burkert
halo models. The halo mass is Mdm = 1.0 × 109 M⊙. From left to right, the curves are for central
neutral hydrogen column densities NHI(0) = 1.0 × 1019, 2.5 × 1019, 5.0 × 1019, 7.5 × 1019, and
1.0 × 1020 cm−2, respectively. In the lefthand panel a velocity dispersion of σg = 10.6 km s−1 has
been assumed, while in the righthand panel σg = 21.2 km s−1. The dotted line running across the
panels shows the HIPASS column density threshold of NHI = 2 × 1018 cm−2.
-- 25 --
Fig. 10. -- The halo mass distribution for a realization of the Local Group in a cold dark matter-
dominated cosmological simulation. The filled circles show all of the halos (other than those
representing the Milky Way and Andromeda) within the Local Group, and the open circles show
all those that lie within 200 kpc of the Milky Way.
-- 26 --
Fig. 11. -- The neutral hydrogen column density as a function of normalized impact parameter
(i.e., the radial offset divided by the cloud radius) for different values of total hydrogen density.
From top to bottom, the curves are for nH = 1.0 × 10−2, 5.0 × 10−3, 2.5 × 10−3, 1.0 × 10−3, and
1.0 × 10−4 cm−3. The profiles generally resemble exponentials except for the presence of a core,
which becomes more prominent as the density is increased. The profiles also show a cutoff close
to the outer boundary, but the latter occurs at neutral hydrogen columns below the detection
threshold of the HIPASS survey.
-- 27 --
Fig. 12. -- The neutral hydrogen column density (solid lines) as a function of impact parameter b
for uniform density clouds with nH = 2.5 × 10−3 (left panel), and 5.0 × 10−3 cm−3 (right panel).
Also plotted are fits for projected exponential volume density distributions of neutral hydrogen, as
discussed in the text (dashed lines). In both cases the cloud radius is identical to the maximum
impact parameter. For the lefthand panel, the scale length h = 1.5 kpc and the central column
density NHI(0) = 1.1 × 1019 cm−2, while for the righthand panel h = 0.42 kpc, NHI(0) = 1.55 ×
1019 cm−2.
-- 28 --
Fig. 13. -- The neutral hydrogen column density (solid lines) as a function of impact parameter b
for a CHVC model in which the gas is confined by a dark matter halo with a NFW profile. This
halo has a mass Mdm = 5 × 108 M⊙, and is shown as the solid line in the center panel of Figure
7. Also plotted are fits for projected exponential volume density distributions of neutral hydrogen,
as discussed in the text (dashed and dotted lines). The NHI distribution cannot be fit by a single
profile as given by equation (7) or (8). The inner and outer regions of the column density profile
are fit by infinite gas distributions (equation [7]) with h = 0.17 kpc and central column density
NHI(0) = 1.0 × 1019 cm−2 (dashed line) and h = 0.55 kpc, NHI(0) = 9.8 × 1017 cm−2 (dotted line),
respectively.
-- 29 --
Fig. 14. -- The neutral hydrogen column density (solid lines) as a function of impact parameter b
for a CHVC model in which the gas is confined by a dark matter halo with a Burkert density profile.
This halo has a mass Mdm = 108 M⊙, and is shown as the rightmost solid line in the lefthand panel
of Figure 8. Also plotted are fits for projected exponential volume density distributions of neutral
hydrogen, as discussed in the text (dashed and dotted lines). The NHI distribution cannot be fit
by a single profile as given by equation (7) or (8). The inner and outer regions of the column
density profile are fit by finite gas distributions (equation [8]) with h = 0.4 kpc and central column
density NHI(0) = 1020 cm−2 (dashed line) and h = 0.3 kpc, NHI(0) = 1.1 × 1019 cm−2 (dotted line),
respectively. The cutoff radii for the two components are 0.83 kpc and 1.83 kpc, respectively.
-- 30 --
Table 1: Results of CHVC models for an ionizing photon flux φi ∼ 104 phot cm−2 s−1.a
nH
NH
(10−3 cm−3)
(1019 cm−2)
0.25
0.50
0.75
1.0
2.5
5.0
7.5
10.0
81.61
40.60
27.17
20.50
8.829
5.074
3.834
3.227
rH
(kpc)
524
130
58
33
5.7
1.6
0.82
0.52
rHI
(kpc)
278
71
32
18.7
3.6
1.14
0.61
0.40
nHI
(10−3 cm−3)
0.0058
0.023
0.051
0.087
0.45
1.4
2.7
4.1
D
(kpc)
47000
12000
5400
3200
610
190
100
67
MHI
Mgas
Mdm/Mgas
(106 Modot)
17000
1100
220
76
2.8
0.28
0.082
0.035
(106 Modot)
4800000
150000
20000
4800
62
2.8
0.56
0.19
--
--
--
--
4
25
64
120
aThe columns give the volume density of total hydrogen (nH), the total hydrogen column (NH) needed to produce
a neutral column density of 1019, the cloud radius (rH ), the radius at the HIPASS sensitivity limit (rHI), the mean
neutral hydrogen volume density (nHI), the distance (D) required to match the typical HIPASS angular size, the
apparent mass that would be derived from HI observations (MHI), the total gas mass (Mgas), and the ratio of dark
matter to gas mass needed to produce the typical observed line width of 25 km s−1 for gravitational confinement of
the gas. (No entries are given for the nH ≤ 10−3 cm−3 models, as they violate the line width constraint with no dark
matter.)
|
astro-ph/0310871 | 1 | 0310 | 2003-10-30T16:42:20 | Dissipationless collapse of a set of N massive particles | [
"astro-ph"
] | The formation of self-gravitating systems is studied by simulating the collapse of a set of N particles which are generated from several distribution functions. We first establish that the results of such simulations depend on N for small values of N. We complete a previous work by Aguilar and Merritt concerning the morphological segregation between spherical and elliptical equilibria. We find and interpret two new segregations: one concerns the equilibrium core size and the other the equilibrium temperature. All these features are used to explain some of the global properties of self-gravitating objects: origin of globular clusters and central black hole or shape of elliptical galaxies. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 29 October 2018
(MN LATEX style file v2.2)
Dissipationless collapse of a set of N massive particles
Fabrice Roy(1)⋆ and J´erome Perez(1,2)†
(1) ´Ecole Nationale Sup´erieure de Techniques Avanc´ees, Unit´e de Math´ematiques Appliqu´ees, 32 Bd Victor, 75015 Paris, France
(2)Laboratoire de l'Univers et de ses TH´eories, Observatoire de Paris-Meudon, 5 place Jules Janssen, 92350 Meudon, France
29 October 2018
ABSTRACT
The formation of self-gravitating systems is studied by simulating the collapse of a set of
N particles which are generated from several distribution functions. We first establish that
the results of such simulations depend on N for small values of N. We complete a previous
work by Aguilar & Merritt concerning the morphological segregation between spherical
and elliptical equilibria. We find and interpret two new segregations: one concerns the
equilibrium core size and the other the equilibrium temperature. All these features are
used to explain some of the global properties of self-gravitating objects: origin of globular
clusters and central black hole or shape of elliptical galaxies.
Key words: methods: numerical, N-Body simulations -- galaxies: formation -- globular
clusters: general.
1 INTRODUCTION
It is intuitive that the gravitational collapse of a set of N
masses is directly related to the formation of astrophysical
structures like globular clusters or elliptical galaxies (the pres-
ence of gas may complicate the pure gravitational N -body
problem for spiral galaxies). From an analytical point of view,
this problem is very difficult. When N is much larger than 2,
direct approach is intractable, and since Poincar´e results of
non analyticity, exact solutions may be unobtainable. In the
context of statistical physics, the situation is more favorable
and, in a dissipationless approximation1, leads to the Colli-
sionless Boltzmann Equation (hereafter denoted by CBE)
∂f
∂t
+ p.
∂f
∂r
+ m
∂ψ
∂r
.
∂f
∂p
= 0
(1)
where f = f (r, p, t) and ψ = ψ (r, t) are respectively the
distribution function of the system with respect to the canon-
ically conjugated (r, p) phase space variables and the mean
field gravitational potential. As noted initially by H´enon
(1960), this formalism holds for such systems if and only
if we consider N identical point masses equal to m. This
problem splits naturally into two related parts: the time de-
pendent regime and the stationary state. We can reasonably
think that these two problems are not completely understood.
The transient time dependent regime was investigated mainly
⋆ [email protected]
† [email protected]
1 The dissipationless hypothesis is widely accepted in the context
of gravitational N -body problem because the ratio of the two-body
relaxation time over the dynamical time is of the order of N . For
a system composed of more than ∼ 104 massive particles a study
during a few hundreds dynamical times can really be considered
as dissipationless, the unique source of dissipation being two-body
encounters.
c(cid:13) 0000 RAS
solutions
considering self-similar
(Lynden-Bell & Eggleton
(1980), Henriksen & Widrow (1995), Blottiau et al. (1988)
and Lancellotti & Kiessling (2001)). These studies conclude
that power law solutions can exist for the spatial dependence
of the gravitational potential (with various powers). Never-
theless, there is no study which indicates clearly that the
time dependence of the solutions disappears in a few dynami-
cal times, giving a well defined equilibrium-like state. On the
other hand, applying Jeans theorem (e.g. Binney & Tremaine
(1987) hereafter BT87, p. 220), it is quite easy to find a sta-
tionary solution. For example, every positive and integrable
function of the mean field energy per mass unit E is a potential
equilibrium distribution function for a spherical isotropic sys-
tem. Several approaches are possible to choose the equilibrium
distribution function. Thermodynamics (Violent Relaxation
paradigm: Lynden-Bell (1967), Chavanis (2002), Nakamura
(2000)) indicate that isothermal spheres or polytropic systems
are good candidates. Stability analyses can be split into two
categories. In the CBE context (see Perez & Aly (1998) for
a review), it is well known that spherical systems (with de-
creasing spatial density) are generally stable except in the case
where a large radial anisotropy is present in the velocity space.
This is the Radial Orbit Instability, hereafter denoted by ROI
(see Perez & Aly (1998), and Perez et al. (1998) for a detailed
analytic and numeric study of these phenomena) which leads
to a bar-like equilibrium state in a few dynamical times. In
the context of thermodynamics of self-gravitating systems, in
a pioneering work by Antonov (1962), it was shown that an
important density contrast leads to the collapse of the core of
system (see Chavanis (2003) for details).
In all these studies there is no definitive conclusion, and the
choice of the equilibrium distribution remains unclear. Intro-
ducing observations and taking into account analytical con-
straints, several models are possible: chronologically, we can
cite (see for example BT87, p. 223-239) the Plummer model
2
F.Roy and J.Perez
(or other polytropic models), de Vaucouleurs r1/4 law, King
and isochrone H´enon model or more recently the very sim-
ple but interesting Hernquist model (Hernquist (1990)) for
spherical isotropic systems. In the anisotropic case, Ossipkov-
Merritt or generalized polytropes can be considered. Finally
for non spherical systems, there also exists some models re-
viewed in BT87 (p. 245-266). Considering this wide vari-
ety of possibilities, one can try to make accurate numeri-
cal simulations to clarify the situation. Surprisingly, such a
program has not been completely carried on. In a pioneer-
ing work, van Albada (1982) remarked that the dissipation-
less collapse of a clumpy cloud of N equal masses could
lead to a final stationary state that is quite similar to el-
liptical galaxies. This kind of study was reconsidered in an
important work by Aguilar & Merritt (1990), with more de-
tails and a crucial remark concerning the correlation between
the final shape (spherical or oblate) and the virial ratio of
the initial state. These authors explain this feature invoking
ROI. Some more recent studies (Cannizzo & Hollister (1992),
Boily et al. (1999) and Theis & Spurzem (1999)) concentrate
on some particularities of the preceding works. Finally, two
works (Dantas et al. (2002) and Carpintero & Muzzio (1995))
develop new ideas considering the influence of the Hubble flow
on the collapse. However, the problem is only partially de-
picted.
The aim of this paper is to analyse the dissipationless collapse
of a large set of N Body systems with a very wide variety of
'realistic' initial conditions. As we will see, the small number
of particles involved, the numerical technique or the specificity
of the previous works did not allow their authors to reach a
sufficiently precise conclusion. The layout of this paper is as
follows. In section 2 we describe in detail the numerical proce-
dures used in our experiments. Section 3 describes the results
we have obtained. These results are then interpreted in section
4, where some conclusions and perspectives are also proposed.
2 NUMERICAL PROCEDURES
2.1 Dynamics
The Treecode used to perform our simulations is a modified
version of the Barnes & Hut (1986) Treecode, parallelised by
D. Pfenniger using the MPI library. We implemented some
computations of observables and adapted the code to suit our
specific problems. The main features of this code are a hi-
erarchical O(N log(N )) algorithm for the gravitational force
calculation and a leap-frog algorithm for the time integration.
We introduced an adaptative time step, based on a very simple
physical consideration. The time step is equal to a fraction nts
of the instantaneous dynamical time Td (2) , i.e. ∆t = Td/nts
. The simulations were run on a Beowulf cluster (25 dual CPU
processors whose speed ranges from 400MHz to 1GHz).
2 The fraction nts is adapted to the virial parameter η and ranges
roughly from nts = 300 when η = 90 to nts = 5000 when η = 08.
The dynamical time we used is given by
N
Td =
Xi=1 px2
Xi=1 pvx2
N
i + y2
i + z2
i
i + vy2
i + vz2
i
2.2 Initial Conditions
The initial virial ratio is an important parameter in our sim-
ulations. The following method was adopted to set the virial
ratio to the value Vinitial. Positions ri and velocities vi are
generated. We can then compute
Vp =
2K
U
where
N
Xi=1
1
2
miv2
i
K =
and
U = −
G
2
N
Xi6=j
mimj
(max((ri − rj )2, ǫ2))1/2 .
(2)
(3)
(4)
In this relation ǫ is a softening parameter whose value is dis-
cussed in section 2.3.2. As the potential energy depends only
on the positions, we obtain a system with a virial ratio equal
to Vinitial just by multiplying all the particle velocities by the
factor (Vinitial/Vp)1/2. For convenience we define
η = Vinitial × 102
(5)
2.2.1 Homogeneous density distribution (Hη)
As we study large N -body systems, we can produce a ho-
mogeneous density by generating positions randomly. These
systems are also isotropic. We produce the isotropic velocity
distribution by generating velocities randomly.
2.2.2 Clumpy density distribution (Cn
η )
A type of inhomogeneous systems is made of systems with a
clumpy density distribution. We first generate n small homoge-
neous spherical systems with radius Rg. Centers of these sub-
systems are uniformly distributed in the system. The empty
space is then filled using a homogeneous density distribution.
In the initial state, each clump contains about 1% of the total
mass of the system and have a radius which represents 5%
of the initial radius of the whole system. These systems are
isotropic.
2.2.3 Power law r−α density distribution (Pα
η )
We first generate the ϕ and z cylindrical coordinates using
two uniform random numbers u1 and u2:
(z, ϕ) = (2u1 − 1, 2πu2) .
Using the inverse transformation method, if
r = RF −1 (u) with F (r) =
1
S Z r
ι
x2−αdx
(6)
(7)
where R is the radius of the system, u is a uniform random
number, ι ≪ 1 and
S = Z 1
x2−αdx ,
(8)
0
then the probability density of r is proportional to r2−α, and
the mass density ρ is proportional to r−α. Finally, one gets
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Dissipationless collapse of a set of N massive particles
3
r =
r√1 − z2 cos ϕ
r√1 − z2 sin ϕ
rz
These systems are isotropic.
(9)
2.2.4 Gaussian velocity distribution(Gσ
η )
Most of the systems we use have a uniform velocity distribu-
tion. But we have also performed simulations with systems
presenting a gaussian initial velocity distribution. These sys-
tems are isotropic, but the x-, y- and z-components of the
velocity are generated following a gaussian distribution. Using
a standard method we generate two uniform random numbers
u1 and u2, and set
The number of particles whose mass is M 6 m 6 M + dM
is n(M )dM . In some models, the value of α and β depends
on the range of mass that is considered. We have used several
types of mass functions, among them the initial mass function
given by Kroupa (2001) (k = I), the one given by Salpeter
(1955) (k = II) and an M −1 mass function (k = III). In
order to generate masses following these functions, we first
calculate αk to produce a continuous function. We then can
calculate the number of particles whose mass is between M
and M + dM . We generate n(M ) masses
mi = M + udM
1 6 i 6 n(M )
(16)
where 0 6 u 6 1 is a uniform random number. In the initial
state, these systems have a homogeneous number density, a
quasi homogeneous mass density and they are isotropic.
vi = p−2σ2 ln u1 cos(2πσ2u2)
where σ is the gaussian standard deviation.
i = x, y, z
(10)
2.2.7 Nomenclature
We indicate below the whole set of our non rotating initial
conditions.
2.2.5 Global rotation (Rf
η )
Some of our initial systems are homogeneous systems with a
global rotation around the Z-axis. The method we choose to
generate such initial conditions is the following. We create a
homogeneous and isotropic system (an H-type system). We
then compute the average velocity of the particles.
¯v =
1
N
N
Xi=1
kvik
(11)
C20
67, C20
65, C20
61, C20
• Homogeneous Hη models: H88, H79, H60, H50, H40, H30,
H20, H15 and H10
29, C20
• Clumpy Cn
η models: C20
14,
07 and C03
10, C20
10
• Power Law Pα
• Gaussian velocity profiles Gσ
• Mass spectra Mk
η models: MI
15 and MI
07
η models: P2.0
η models: G1
50, MIII
35, MII
25,
50 , P0.5
50 , P1.0
50 , P2.0
09 , P1.0
48, C20
39, C20
and P1.5
40
50, G2
50, G3
50, G4
12
50, MII
51 , MI
and G5
50
MIII
10 , P1.5
08
We project the velocities on a spherical referential, and add a
fraction of ¯v to vφ with regard to the position of the particle.
We set
For all these models we ran the numerical simulations
with 30 000 particles (see § 3.1)
vi,φ = vi,φ + f
ρi¯v
R
(12)
2.3 Observables
where f is a parameter of the initial condition, ρi is the dis-
tance from the particle to the Z-axis and R the radius of the
system. The amount of rotation induced by this method can
be evaluated through the ratio:
µ = Krot/K,
(13)
where K is
whereas Krot
Navarro & White (1993):
total kinetic
the
energy defined above,
is the rotation kinetic energy defined by
2.3.1 Units
the
are not
commonly used ones
Our units
(see
Heggie & Mathieu (1986)). We did not set the total en-
ergy E of the system to −0.25 because we wanted to prescribe
instead the initial virial ratio Vinitial, the size R of the system
and its mass M . We thus have M = 1, R = 10 and G = 1,
and values of Vinitial and E depending on the simulation. We
can link the units we have used with more standard ones. We
have chosen the following relationships between our units of
length and mass and common astrophysical ones:
Krot =
1
2
N
Xi=1
mi
(Ji · Jtot)2
i − (ri · Jtot)2]
[r2
Above, Ji is the specific angular momentum of particle i, Jtot
is a unit vector in the direction of the total angular momentum
of the system. In order to exclude counter-rotating particles,
the sum in equation (14) is actually carried out only over those
particles which verify the condition (Ji · Jtot) > 0.
(14)
M = 106M⊙ and R = 10 pc
Our unit of time ut is given by:
1ut = r R3
cGsMs
sGcMc ≈ 4.72 1011 s = 1.50 104 yr
R3
where variables Xs are expressed in our simulation units and
variables Xc in standard units.
(17)
(18)
2.2.6 Power-law initial mass function(Mk
η)
2.3.2 Potential softening and energy conservation
Almost all the simulations we made assume particles with
equal masses. However, we have created some initial systems
with a power-law mass function, like
n(M ) = αM β
(15)
The non conservation of the energy during the numerical evo-
lution has three main sources.
The softening parameter ε introduced in the potential calculus
(cf. equation 4) is an obvious one. This parameter introduces
a lower cutoff Λ in the resolution of length in the simulations.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4
F.Roy and J.Perez
Following Barnes & Hut (1989), structural details up to scale
Λ . 10ε are sensitive to the value of ε. Moreover, in order to
be compatible with the collisionless hypothesis, the softening
parameter must be greater than the scale where important
collisions can occur. Still following Barnes & Hut (1989), this
causes
&
(19)
G hmi
hv2i
ε
10
In our collapse simulation with 3 · 104 particles, this results
in εη & 2/3. The discretization of time integration introduces
inevitably another source of energy non conservation, particu-
larly during the collapse. The force computation also generates
errors. The choice of the opening angle Θ, which governs the
accuracy of the force calculation of the Treecode, is a compro-
mise between speed and accuracy. For all these reasons, we
have adopted ε = 0.1. This choice imposes η & 6 (for 30 · 104
particles). This trade-off allowed to perform simulations with
less than 1% energy variation without requiring too much com-
puting time. For each of our experiments, the total CPU time
ranges between 3 to 24 hours for 3000 ut and 3 · 104 particles.
The total agregated CPU time of all our collapse experiments
is approximately 6 months.
We have tested two other values of the softening parameter
(ε = 0.03 and ε = 0.3) for several typical simulations. These
tests did not reveal significant variations of the computed ob-
servables.
2.3.3 Spatial indicators
As indicators of the geometry of the system, we computed
axial ratios, radii containing 10% (R10), 50% (R50) and 90%
(R90) of the mass, density profile ρ (r) and equilibrium core
radius. The axial ratios are computed with the eigenvalues λ1,
λ2 and λ3 of the (3x3) inertia matrix I, where λ3 6 λ2 6 λ1
and, if the position of the particle i is ri = (x1,i ; x2,i ; x3,i)
Iµν = −
mi xµ,i xν,i
for µ 6= ν = 1, 2, 3
Iµµ =
mi(r2
i − x2
µ,i)
for µ = 1, 2, 3
The axial ratios a1 and a2 are given by a2 = λ1/λ2 > 1.0 and
a1 = λ3/λ2 6 1.0.
The density profile ρ, which depends only on the radius r,
together with the Rδ (δ = 10, 50, 90) have a physical mean-
ing only for spherical or nearly spherical systems. For all the
spatial indicators computations we have only considered par-
ticles whose distance to the center of mass of the system is less
than 6×R50 of the system. This assumption excludes particles
which are inevitably 'ejected' during the collapse3.
After the collapse a core-halo structure forms in the system.
In order to measure the radius of the core, we have com-
puted the density radius as defined by Casertano and Hut
(see Casertano & Hut (1985)). The density radius is a good
estimator of the theoretical and observational core radius.
We have also computed the radial density of the system. The
N
Pi=1
Pi=1
N
3 The number of excluded particles ranges from 0% to 30% of the
total number of particles, depending mostly on η. For example, the
number of excluded particles is 0% for H80, 3% for C20
67, 5% for H50,
22% for C20
10 and 31% for H10.
density is computed by dividing the system into spherical bins
and by calculating the total mass in each bin.
2.3.4 Statistical indicators
When the system has reached an equilibrium state, we com-
pute the temperature of the system
T =
2 hKi
3N kB
(21)
where K is the kinetic energy of the system, kB is the Boltz-
mann constant (which is set to 1) and the notation hAi denotes
the mean value of the observable A, defined by
hAi =
1
N
N
Xi=1
Ai
(22)
In order to characterise the system in the velocity space we
have computed the function
(23)
κ (r) =
2(cid:10)v2
i,rad(cid:11)r6ri<r+dr
(cid:10)v2
i,tan(cid:11)r6ri<r+dr
where vi,rad is the radial velocity of the ith particle, and vi,tan
its tangential velocity. For spherical and isotropic systems
(a1 ≃ a2 ≃ 1 and κ (r) ≃ 1), we have fitted the density by
1- a polytropic law
ρ = ρ0ψγ
which corresponds to a distribution function
f (E) ∝ Eγ−3/2
2- an isothermal sphere law
ρ = ρ1eψ/s2
which corresponds to a distribution function
f (E) =
ρ1
(2πs2)3/2 e
E
s2
(24)
(25)
(26)
(27)
3 DESCRIPTION OF THE RESULTS
We have only studied systems with an initial virial ratio cor-
responding to η ∈ [7, 88]. In such systems, the initial velocity
dispersion cannot balance the gravitational field. This pro-
duces a collapse. After this collapse, in all our simulations the
system reaches an equilibrium state characterised by a tempo-
ral mean value of the virial ratio equal to −1, i.e η = 100, and
stationary physical observables. These quantities (defined in
section 2) are presented in a table of results in the appendix
of this paper. The following results will be discussed and in-
terpreted in section 4.
3.1 Relevant number of particles
In all previous works on this subject (van Albada (1982),
Aguilar & Merritt (1990), Cannizzo & Hollister (1992) and
Boily et al. (1999)), the authors did not really consider the
influence of the number of particles on their results. In the
first two and more general works, this number is rather small
(not more than a few thousands in the largest simulations).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
(20)
Using the least square method in the ln(ρ) − ln(ψ) plane we
get (γ, ln (ρ0)) and (cid:0)s2, ln (ρ1)(cid:1).
Dissipationless collapse of a set of N massive particles
5
Figure 2. Axial ratios of equilibrium states reached from Homoge-
neous, Clumpy, Gaussian velocity dispersion, Power law and Mass
spectrum initial conditions.
of particles used in each case ranges from 102 to 105. We can
see in Figure 1 that some observables are N-dependent when
N < 3.104. In particular, R50 and Rd present a monotonic
variation larger than 50% when N varies from 102 to 105 and
the ellipticity of the final state is overestimated for small val-
ues of N . As a conclusion of this preliminary study, we claim
that the relevant number of particles for collapse simulations is
N > 3.104. As all simulations have been completed with a to-
tal energy loss smaller than 1%, we state moreover that this re-
sult is independent of the numerical scheme used (Treecode or
Direct N-Body). As a consequence, the simulations presented
hereafter have been performed using N = 3.104 particles.
3.2 Morphological segregation
An important study by Aguilar & Merritt (1990) shows that,
in the case of an initial density profile ρ ∝ r−1, the shape of
the virialized state depends on η: a very small η leads to a
flattened equilibrium state, when a more quiet collapse pro-
duces a spherical one. Our investigations concern a wide range
of different initial conditions (homogeneous, clumpy,...) and
show that the influence of η depends on the properties of the
initial system4. Figure 2 shows the axial ratios of the equi-
librium state reached by our simulations. In fact, only a few
simulations produced a final state with an ellipticity greater
than E1. Every homogeneous initial condition (i.e. Hη, Gσ
η and
Mk
η) resulted in a spherical equilibrium state independently of
the values of η we tested. Cold clumpy systems have a weakly
flattened equilibrium state. The only final systems with an el-
lipticity significantly greater than E1 are those produced by
the collapse of cold Pα
η .
Previous studies invoked ROI to explain this morphological
segregation. However, it seems that ROI requires inhomo-
geneities near the center to be triggered.
4 The particular case of rotating initial conditions is discussed in a
special section.
Figure 1. Influence of the number of particles on the physical ob-
servables of a collapsing system. Axial ratios are on the top panel,
radius containing 50% of the total mass and density radius are on
the middle pannel and the best s2 and γ fit for respectively isother-
mal and polytropic distribution function are on the bottom panel.
All cases are initially homogeneous with η = 10 (solid line), η = 50
(dotted line) and η = 80 (dashed line). The number N of particles
used is in units of 103.
The two other studies are more specific and use typically 104
and, in a few reference cases, 2.104 particles. In order to test
the influence of the number of particles on the final results, we
have computed several physical observables of some collapsing
systems with various numbers of particles. The results are pre-
sented in Figure 1. In order to check the influence of N in the
whole phase space, we have studied positions and velocities
related observables: a1, a2, R10, R50 and R90 and parame-
ters of isothermal and polytropic fit models namely γ and s2.
Moreover, in order to be model-independent, we have studied
three representative initial conditions: H80, H50 and H10, i.e.
respectively initially hot, warm and cold systems. The number
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
6
F.Roy and J.Perez
Figure 3. Density profile for Hη models plotted in units of R50 .
Figure 5. Density profile for Pα
η models plotted in units of R50.
Figure 4. Density profile for Cn
The dashed line corresponds to the C03
10 model
η models plotted in units of R50.
3.3 Characteristic size segregation
In addition to the morphological segregation, presented in the
previous section, we discovered a finer phenomenon.
In the Figures 3, 4, 5, 6 and 7, we have plotted the mass
density of all equilibrium states produced by the collapse of
our initial conditions as a function of the ratio r/R50. These
plots represent the density at the end of our simulations (after
about 100 crossing times). These functions do not significantly
evolve after the collapse except for MI
07. For this special case,
a comparative plot is the subject of Figure 8.
All equilibrium states we obtain clearly fall into two cate-
gories:
• Flat Core Systems
All these systems present a core halo structure, i.e. a large
central region with a constant density and a steep envelope.
Figure 6. Density profile for Gσ
η models plotted in units of R50.
• Small Core Systems
These systems are typically such that ln (R50/Rd) < 0.5 and
ln (R10/Rd) < −0.05.
For such systems, the central density is two order of magnitude
larger than for Flat Core systems. There is no central plateau
and the density falls down regularly outward. These systems
are typically such that ln (R50/Rd) > 0.7 and ln (R10/Rd) >
0.1.
The diagram ln (R10/Rd) vs ln (R50/Rd) is the subject
of Figure 9. One can see in this figure that each equilibrium
state belongs to one or the other family except in a few par-
ticular cases. In the Flat Core family we found all Hη, Gσ
η and
η systems except MI
Mk
η systems namely P0.5
50 and
P1
50. These systems are all initially homogeneous or slightly
inhomogeneous (e.g. P0.5
50 systems). In the Small Core
family, we found all C20
09 systems. These
systems are all initially rather very inhomogeneous. Finally,
50 and P1
η and the P2.0
7, and two Pα
50 and P2.0
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Dissipationless collapse of a set of N massive particles
7
Figure 7. Density profile for Mk
η models plotted in units of R50.
Figure 9. The core size segregation: ln (R10/rc) vs ln (R50/rc) is
plotted for all non rotating systems.
10 and MI
there are 5 systems in-between the two categories: C03
10, P1.5
40 ,
P1.5
08 , P1
07. This last model is the only one which mi-
grates from Flat core set (when t ≃ 10T d) to the edge of the
small core region (when t ≃ 100T d).
3.4 Equilibrium Distribution Function
In order to compare systems in the whole phase space, we
fitted the equilibrium state reached by each system with two
distinct isotropic models, e.g. polytropic and isothermal (see
equations 24-27 or BT87, p.223-232). Figure 10 shows these
two fits for the P0.5
50 simulation. The technique used for the
fit is described in section 2 of this paper. The result obtained
for this special study is the following: the equilibrium states
reached by our initial conditions can be fitted by the two
models with a good level of accuracy. As long as η < 70,
the polytropic fit gives a mean value γ = 4.77 with a stan-
dard deviation of 2.48 10−1. This deviation represents 5.1%
of the mean value. This value corresponds typically to the well
known Plummer model for which γ = 5 (see BT87 P.224 for
details). When the collapse is very quiet ( typically η > 70 )
polytropic fit is always very good but the value of the index is
much larger than Plummer model, e.g. γ = 6.86 for H79 and
γ = 7.37 for H88. The corresponding plot is the subject of the
Figure 12. All the data can be found in appendix. As we can
see on the example plotted in Figure 10, the isothermal fit is
generally not as good as the polytropic one. On the whole set
of equilibria, isothermal fits give a mean value s2 = 2.5 10−2
with a standard deviation of 1.6 10−2 (60%). The correspond-
ing plot is the subject of Figure 11. All the data can be found
in appendix.
In fact both isothermal and polytropic fits are reasonable:
as long as the model is able to reproduce a core halo structure
the fit is correct. The success of the Plummer model, which
density is given by
ρ (r) =
3
4πb3 (cid:20)1 +(cid:16) r
b(cid:17)2(cid:21)−5/2
can be explained by its ability to fit a wide range of models
with various ratio of the core size over the half-mass radius.
Figure 8. Comparison between the evolution of the mass density
with respect to time for C20
07. For each case,
plotted times are 10,20,30,40,50,75 and 100 Td
10 (top-panel) and MI
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
8
F.Roy and J.Perez
Figure 10. Polytropic and isothermal fit for the P 0.5
50 simulation.
Figure 12. Best fit of the γ parameter for a polytropic model for
all non rotating systems studied. The error bars correspond to the
least square difference between the data and the model.
Figure 11. Best fit of the s2 parameter for an isothermal model
for all non rotating systems studied. The error bar correspond to
the least square difference between the data and the model.
The adjustment of this ratio is made possible by varying the
free parameter b. We expect that other core halo models like
King or Hernquist models work as well as the Plummer model.
As a conclusion of this section, let us say that as predicted
by theory there is not a single universal model to describe the
equilibrium state of isotropic spherical self-gravitating system.
3.5 Influence of rotation
We saw in section 3.2 a source of flattening for self-gravitating
equilibrium. Let us now show the influence of initial rotation,
which is a natural candidate to produce flattening. The way
we have added a global rotation and the significance of our
rotation parameters f and µ are explained in section 2.2.5.
The set of simulations performed for this study contains 31
different elements. The initial virial ratio ranges from η =
10 to η = 50, and the rotation parameter from f = 0 (i.e.
µ = 0) to f = 20 (i.e. µ = 0.16 when η = 50). As a matter
of fact, equilibrium states always preserve a rather important
Figure 13. Axial ratios as for different values of η as a function of
the initial solid rotation parameter f
part of the initial rotation5 and, observed elliptical gravitating
systems generally possess very small amount of rotation (see
e.g. Combes et al. (1995)). We thus exclude large values of f
.
Our experiments exhibit two main features (see Figure 13): on
the one hand, rotation produces a flattened equilibrium state
only when f exceeds a triggering value (typically f = fo ≃ 4).
On the other hand, we have found that for a given value of
η, the flatness of the equilibrium is roughly f−independent,
provided that f > fo.
5 We observed that µ is always smaller in the equilibrium state than
in the initial one, typically each rotating systems conserves 65% of
the initial µ
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Dissipationless collapse of a set of N massive particles
9
Figure 14. Evolution of the temperature as a function of time
Figure 15. Energy-Temperature diagram
3.6 Thermodynamical segregation
As we study isolated systems, the total energy E contained in
the system is constant during the considered dynamical evo-
lution . This property remains true as long as we consider
collisionless evolutions. For gravitational systems, this means
that we cannot carry out any simulation of duration larger
than a few hundred dynamical times. We have obviously taken
this constraint into account in our experiments. All systems
which experience a violent relaxation reach an equilibrium
state which is stationary in the whole phase space. Spatial be-
haviour like morphological segregation produced by ROI was
confirmed and further detailled thanks to our study. A new
size segregation was found in section 3.3. Now let us consider
another new segregation appearing in the velocity space. Each
equilibrium state is associated with a constant temperature T ,
calculated using equation (21). More precisely, we have calcu-
lated the temporal mean value 6 of the temperatures, evalu-
ated every one hundred time units. As we can see in Figure
14, after the collapse and whatever the nature of the initial
system is, the temperature is a very stable parameter.
Figure 15 shows the E − T diagram of the set of all non rotat-
ing simulations. It reveals a very interesting feature of post-
collapse self-gravitating systems.
On the one hand, the set of systems with a total energy
E > −0.054 can be linearly fitted in the E − T plane. We call
this set Low Branch 1 (hereafter denoted by LB1, see Figure
15). On the other hand, the set of the systems with a total
energy E < −0.054 splits into two families. The first is an
exact continuation of LB1. Hence we named it Low Branch 2
(hereafter denoted by LB2). The second can also be linearly
fitted, but with a much greater slope (one order of magnitude).
We label this family High Branch (hereafter denoted by HB).
In LB1 or LB2, we find every H, G, and M systems with
η > 25, every P and every C with n > 10. In HB, we find C03
10
and every H, G and M systems with η < 25. This segregation
thus affects violent collapses (cold initial data): on the one
hand, when η > 25 all systems are on LB1, on the other hand
for η < 25, initially homogeneous or quasi-homogeneous (e.g.
C03
10) systems reach HB when initially inhomogeneous systems
stay on LB2 instead.
4
INTERPRETATIONS, CONCLUSIONS AND
PERSPECTIVES
Let us now recapitulate the results we have obtained and pro-
pose an interpretation:
(i) The equilibrium state produced by the collapse of a set
of N gravitating particles is N−independent provided that
N > 3.0 104.
(ii) Without any rotation, the dissipationless collapse of a
set of gravitating particles can produce two relatively distinct
equilibrium states:
large core and a steep envelope.
• If the initial set is homogeneous, the equilibrium has a
• If the initial set contains significant inhomogeneities
(n > 10 for clumpy systems or α > 1 for power law sys-
tems), the equilibrium state has only a small core around of
which the density falls down regularly.
The explanation of this core size segregation is clear: it is
associated to the Antonov core-collapse instability occurring
when the density contrast between central and outward re-
gion of a gravitating system is very big. As a matter of fact,
if the initial set contains inhomogeneities, these collapse much
more quickly than the whole system7 and fall quickly into the
central regions. The density contrast becomes then very large
and the Antonov instability triggers producing a core collapse
phenomenon. The rest of the system then smoothly collapses
around this collapsed core. If there are no inhomogeneities in
the initial set, the system collapses as a whole, central den-
sity grows slowly without reaching the triggering value of the
Antonov instability. A large core then forms. Later evolution
can also produce core collapse: this is what occurs for our MI
07
system (see Figure 8). This is an initially homogeneous system
6 The temporal mean value is computed from the time when the
equilibrium is reached until the end of the simulation
7 Because their Jeans length is much more smaller than the one of
the whole system.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
10
F.Roy and J.Perez
with Kroupa mass spectrum which suffers a very strong col-
lapse. As the mass spectrum is not sufficient to bring quickly
enough a lot of mass in the center of the system, Antonov
instability does not trigger and a large core forms. As the col-
lapse is very violent, an increasing significant part of particles
are progressively "ejected" and the core collapse takes pro-
gressively place. This is the same phenomenon which is gener-
ally invoked to explain the collapsed core of some old globular
clusters (e.g. Djorgowski et al. (1986)): during its dynamical
evolution in the galaxy, some stars are tidally extracted from
a globular cluster, to compensate this loss the cluster concen-
trate its core, increasing then the density contrast, triggering
sooner or later the Antonov instability.
(iii) Without any rotation, the collapse (violent or quiet)
of an homogeneous set of gravitating particles produces an
E0 (i.e. spherical) isotropic equilibrium state. There are two
possible ways to obtain a flattened equilibrium:
• Introduce a large amount of inhomogeneity near the
center in the initial state, and make a violent collapse (η <
25).
• Introduce a sufficient amount (f > 4) of rotation in the
initial state.
These two ways have not the same origin and do not produce
the same equilibrium state.
In the first case, one can reasonably invoke the Radial Orbit
Instability: as a matter of fact, as it is explained in a lot of
works (see Perez et al. (1998) for example) two features are
associated to this phenomenon. First of all, it is an instability
which needs an equilibrium state from which it grows. Sec-
ondly, it triggers only when a sufficient amount of radial or-
bits are present. The only non rotating flattened systems we
observed just combine this two conditions: sufficient amount
of radial orbits because the collapse is violent and something
from which ROI can grow because we have seen in the previous
point that inhomogeneities collapse first and quickly join the
center. The fact that cold Pα
η systems are more flattened than
Cα
η ones is in complete accordance with our interpretation:
as a matter of fact, by construction, power law systems have
an initial central overdensity, whereas clumpy systems create
(quickly but not instantaneously) this overdensity bringing the
collapsed clumps near the center. The ROI flattening is oblate
(a2 ≃ 1 and a1 < 1).
The rotational flattening is more natural and occurs when
the centrifugal force overcomes the gravitational pressure. The
rotational flattening is prolate ( a2 > 1 and a1 ≃ 1). We no-
tice that initial rotation must be invoked with parsimony to
explain the ellipticity of some globular clusters or elliptical
galaxies. As a matter of fact, these objects are very weakly
rotating systems and our study has shown that the amount of
rotation is almost constant during the collapse.
(iv) Spherical equilibria can be suitably fitted by both
isothermal and polytropic laws with various indexes. It sug-
gests that any distribution function of the energy exhibiting
an adaptable core halo structure (Polytrope, Isothermal, King,
Hernquist,...) can suitably fit the equilibrium produced by the
collapse of our initial conditions.
(v) There exists a temperature segregation between equilib-
rium states. It concerns only initially cold systems (i.e. systems
which will suffer a violent collapse): for such systems when η
decreases, the equilibrium temperature T increases much more
for initially homogeneous systems than for initially inhomoge-
neous systems. On the other hand, whatever their initial ho-
mogeneity, quiet collapses are rather all equivalent from the
point of view of the equilibrium temperature: T increases in
the same way for all systems as η decreases. This feature may
be the result of the larger influence of the dynamical friction
induced by the primordial core on the rapid particles in a vi-
olent collapse.
All these properties may be directly confronted to phys-
ical data from globular clusters (see Harris catalogue Harris
(1996)) or galaxies observations.
As a matter of fact, in the standard "bottom-up" scenario of
the hierarchical growth of structures, galaxies naturally form
from very inhomogeneous medium. Our study then suggests
for the equilibrium state of such objects a potential flattening
and a collapsed core. This is in very good accordance with the
E0 to E7 observed flatness of elliptical galaxies and may be a
good explanation for the presence of massive black hole in the
center of galaxies (see Schodel et al. (2002)).
On the other hand Globular Clusters observations show that
these are spherical objects (the few low flattened clusters all
possess a low amount of rotation), and that their core is gen-
erally not collapsed (the collapsed core of almost 10% of the
galactic Globular Clusters can be explained by their dynami-
cal evolution through the galaxy). Our study then expect that
Globular Clusters form from homogeneous media.
These conclusions can be tested using the E − T plane. As
a matter of fact, we expect that an E − T plane build from
galactic data would not present any High Branch whereas the
same plane build from Globular Clusters data would.
ACKNOWLEDGMENTS
We thank the referee for the relevance of his remarks and sug-
gestions. We thank Joshua E. Barnes, who wrote the original
Treecode. We also particularly thank Daniel Pfenniger for the
use of the parallel Treecode. The simulations were done on
the Beowulf cluster at the Laboratoire de Math´ematiques Ap-
pliqu´ees from the ´Ecole Nationale Sup´erieure de Techniques
Avanc´ees.
This is a preprint of an article accepted for publication in
Monthly Notices of The Royal Astronomical Society c(cid:13)2003
The Royal Astronomical Society.
REFERENCES
Aguilar, L., and Merritt, D., 1990, ApJ, 354, 33
Antonov, V.A., Vest. Len. Univ., 7, 135, 1962 in russian
translated in english in Dynamics of star clusters, Ed. Good-
man J. & Hut P. (IAU 113, Reidel 1985)
Barnes, J. and Hut, P., 1986, Nature, 324, 446
Barnes, J. and Hut, P., 1989, ApJS, 70, 389
Binney, J., and Tremaine, S., 1987, Galactic dynamics,
Princeton University Press
Blottiau, P., Bouquet, S., Chi`eze,J.-P., 1988,A&A, 207, 24
Boily, C.M., Clarke, C.J., Murray, S.D., 1999,M.N.R.A.S.,
302, 399
Cannizzo,J.K., and Hollister, T.C., 1992, ApJ, 400, 58
Carpintero,D.D., Muzzio, J.C., 1995, ApJ, 440, 5
Casertano, S., Hut,P., 1985, ApJ, 298, 80
Chavanis, P.-H.,2002,Submitted
to Phys. Rev. E,
cond-mat/0107345
Chavanis, P.-H., A& A,401, 15, 2003, astro-ph/0207080
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Dissipationless collapse of a set of N massive particles
11
Combes, F., Boiss´e, P., Mazure, A., Blanchard, A., 1995,
Galaxies and cosmology, Springer Verlag
Dantas, C.C., Capelato, H.V., de Carvalho, R.Ribiero, A.L.B,
2002, A&A, astroph-0201247
Djorgovski, S., King, I. R., Vuosalo, C., Oren, A., Penner,
H.,IAU Symp.118, Ed. J.B. Hearnshaw, P.L. Cottrell, 1986
Harris, W.E. 1996, AJ, 112, 1487
Heggie, D.C., and Mathieu R.D., 1986, in The Use of Su-
percomputers in Stellar Dynamics, ed. P. Hut & S. L. W.
McMillan, Springer Verlag.
H´enon, M. , 1960, An.Astro, 23, 472
Henriksen, R.N., and Widrow, L.M., 1995, M.N.R.A.S, 276,
679, astro-ph/9412047
Hernquist, 1990, ApJ, 356, 359
Kroupa, P., 2001, M.N.R.A.S., 322, 231, astro-ph/0009005
Lancellotti, C.,Kiessling, M., 2001,ApJ, 549, L93
Lynden-Bell, D., and Eggleton, P.P., 1980, M.N.R.A.S., 191,
483
Lynden-Bell, D., 1967, M.N.R.A.S, 136, 601
Nakamura, T.K., 2000, ApJ, 531, 739
Navarro, J.E., White, S.D.M. 1993, M.N.R.A.S., 265, 271
Perez, J. and Aly, J.-J., 1998, M.N.R.A.S., 280, 689,
astro-ph/9511103
Perez, J. Alimi, J.-M., Aly, J.-J. and Scholl, H., 1998,
M.N.R.A.S., 280, 700, astro-ph/9511090
Salpeter, E.E., 1955, ApJ, 121, 161
R. Schodel, T. Ott, R. Genzel, R. Hofmann, M. Lehnert, A.
Eckart, N. Mouawad, T. Alexander, M.J. Reid, R. Lenzen,
M. Hartung, F. Lacombe, D. Rouan, E. Gendron, G. Rous-
set, A.-M. Lagrange, W. Brandner, N. Ageorges, C. Lidman,
A.F.M. Moorwood, J. Spyromilio, N. Hubin, K.M. Menten
,17 Oct 2002 ,Nature, astro-ph/0210426
Theis, C. and Spurzem, R., 1999, A&A, 341, 361,
astro-ph/9810023
van Albada, T.S., 1982, M.N.R.A.S., 201, 939
APPENDIX A: TABLES OF RESULTS
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
12
F.Roy and J.Perez
Table A1. Homogeneous Initial Conditions: Hη
η
88
79
60
50
40
30
20
15
10
∆ E
( %)
η
(end)
a1
a2
R10
R50
R90
Rd
γ
Σ2
γ
(× 102)
s2
(× 102 )
Σ2
s2
(× 103)
T
−E
(× 1011 )
(× 102)
0.0
0.0
0.0
0.0
0.1
0.0
0.0
0.0
1.4
98
99
96
96
96
96
101
108
120
1.02
1.00
1.01
1.01
1.01
1.02
1.01
1.01
1.02
0.99
1.00
0.98
0.98
0.99
0.99
0.99
0.99
0.98
3.39
3.11
2.41
2.04
1.72
1.36
0.95
0.74
0.52
6.53
6.01
4.73
4.09
3.51
2.88
2.22
1.89
1.59
10.6
9.8
11.5
15.1
25.7
253.2
874.1
1143.0
1448.0
5.02
4.59
3.41
2.81
2.30
1.75
1.17
0.88
0.60
7.37
6.86
5.05
4.73
4.72
4.66
4.68
4.66
4.59
-6
-4
-1
-2
-2
-2
-1
-1
-1
1.12
1.37
2.04
2.49
2.79
3.27
4.03
4.73
5.76
-3
-3
-2
-2
-2
-3
-5
-6
-9
2.30
2.20
2.89
3.04
3.39
3.74
4.78
5.83
7.71
3.30
3.60
4.20
4.50
4.80
5.10
5.39
5.53
5.66
Table A2. Clumpy Initial Condition: Cn
η
η
10
67
65
61
48
39
29
14
10
7
n ∆ E
( %)
η
(end)
a1
a2
R10
R50
R90
Rd
γ
Σ2
γ
(× 102)
s2
(× 102)
Σs2
(× 103 )
T
−E
(× 1011 )
(× 102 )
3
20
20
20
20
20
20
20
20
20
0.1
0.8
0.9
0.6
0.6
0.5
1.2
0.3
1.4
0.3
97
95
95
95
94
94
94
93
97
94
1.03
1.01
1.02
1.05
1.03
1.04
1.04
1.09
1.13
1.14
0.96
0.94
0.96
0.98
0.99
0.99
0.99
0.98
0.99
0.92
0.55
0.93
0.93
0.86
0.81
0.76
0.72
0.64
0.61
0.57
1.85
6.14
5.63
5.15
4.24
3.84
3.42
2.72
2.54
2.44
1241.0
16.1
13.5
12.3
11.8
12.8
15.4
39.5
345.5
224.6
0.58
0.66
0.63
0.58
0.59
0.40
0.56
0.48
0.54
0.45
4.61
5.00
5.27
5.43
5.15
4.72
4.42
4.56
4.70
4.74
-1
-1
-2
-3
-3
-1
-1
-1
-9
-1
5.21
1.98
2.00
2.09
2.61
3.20
3.81
4.27
4.25
4.41
-11
-10
-13
-11
-13
-10
-6
-7
-15
-12
6.99
2.86
3.07
3.32
3.79
4.06
4.24
4.66
4.98
5.36
5.82
3.96
4.29
4.64
5.31
5.65
5.97
6.50
6.66
6.76
Table A3. Power Law Initial Conditions: Pα
η
η
α
∆ E
( %)
η
(end)
a1
a2
R10
R50
R90
Rd
γ
Σ2
γ
(× 102)
s2
(× 102)
Σs2
(× 103 )
T
−E
(× 1011 )
(× 102 )
50
50
10
8
50
40
9
0.5
1
1
1.5
2
1.5
2
0.0
0.0
0.1
0.1
1.7
0.1
1.6
95
94
96
96
93
96
96
1.01
1.01
1.00
1.01
1.02
1.00
1.01
0.99
0.99
0.80
0.71
0.99
0.99
0.78
1.84
1.56
0.69
0.62
0.53
0.97
0.38
3.92
3.77
2.71
2.63
3.20
3.31
2.51
13.5
12.1
382.2
25.1
9.2
11.0
10.6
2.53
2.01
0.70
0.61
0.34
1.03
0.18
4.66
4.77
4.61
4.63
5.30
4.71
4.68
-1
-6
-8
-7
-6
-8
-10
2.65
2.78
4.05
4.42
3.35
3.44
5.21
-2
-3
-8
-9
-9
-6
-20
3.54
3.69
4.85
5.52
5.30
4.42
6.73
4.69
5.00
6.32
7.18
7.32
5.99
9.31
Table A4. Mass Spectrum Initial Conditions: Mi
η
η
i
∆ E
(% )
η
(end)
a1
a2
R10
R50
R90
Rd
γ
Σ2
γ
(× 102)
s2
(× 102)
Σs2
(× 103 )
T
−E
(× 1011 )
(× 102 )
Krou
7
1/M
15
25
Salp
35 Krou
51
1/M
50 Krou
50
Salp
5.0
0.6
0.4
0.2
0.2
0.1
0.1
132
101
99
98
95
96
96
1.02
1.01
1.01
1.01
1.01
1.02
1.01
0.99
0.98
0.98
0.99
0.98
0.98
0.98
0.25
0.68
1.18
1.55
1.79
1.93
2.03
2.04
2.04
2.51
3.18
4.19
4.09
4.13
1721.0
366.8
225.5
39.8
14.2
15.1
15.1
0.37
0.97
1.54
2.09
2.83
2.87
2.87
4.26
4.65
4.66
4.66
4.67
4.60
4.70
-1
-1
-2
-2
-9
-1
-2
5.16
4.35
3.59
2.97
2.39
2.45
2.45
-15
-10
-3
-3
-4
-3
-2
9.97
5.72
4.08
3.51
3.15
3.19
3.08
5.62
5.53
5.23
4.95
4.48
4.50
4.49
Table A5. Gaussian Velocity Dispersion Initial Conditions: Gσ
η
η
σ
∆ E
(% )
η
(end)
a1
a2
R10
R50
R90
Rd
γ
Σ2
γ
(× 102)
s2
(× 102)
Σ2
s2
(× 103 )
T
−E
(× 1011 )
(× 102 )
48 G1
49 G2
50 G3
12 G4
50 G5
0.0
0.0
0.0
0.4
0.0
95
95
95
118
96
1.00
1.02
1.00
1.03
1.01
1.00
0.99
1.00
0.98
0.99
1.72
1.74
1.90
0.56
1.98
3.98
4.02
4.08
1.77
4.11
14.6
14.4
14.1
1312.0
14.8
2.23
2.28
2.56
0.64
2.72
4.66
4.65
4.72
4.56
4.71
-5
-6
-8
-1
-1
2.64
2.61
2.52
5.33
2.50
-3
-3
-2
-10
-2
3.13
3.13
3.09
6.08
3.19
4.50
4.50
4.50
5.63
4.50
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
astro-ph/0702502 | 1 | 0702 | 2007-02-20T00:17:14 | Bounds on the mass and abundance of dark compact objects and black holes in dwarf spheroidal galaxy halos | [
"astro-ph"
] | We establish new dynamical constraints on the mass and abundance of compact objects in the halo of dwarf spheroidal galaxies. In order to preserve kinematically cold the second peak of the Ursa Minor dwarf spheroidal (UMi dSph) against gravitational scattering, we place upper limits on the density of compact objects as a function of their assumed mass. The mass of the dark matter constituents cannot be larger than 1000 solar masses at a halo density in UMi's core of 0.35 solar masses/pc^3. This constraint rules out a scenario in which dark halo cores are formed by two-body relaxation processes. Our bounds on the fraction of dark matter in compact objects with masses >3000 solar masses improve those based on dynamical arguments in the Galactic halo. In particular, objects with masses $\sim 10^{5}$ solar masses can comprise no more than a halo mass fraction $\sim 0.01$. Better determinations of the velocity dispersion of old overdense regions in dSphs may result in more stringent constraints on the mass of halo objects. For illustration, if the preliminary value of 0.5 km/s for the secondary peak of UMi is confirmed, compact objects with masses above $\sim 100$ solar masses could be excluded from comprising all its dark matter halo. | astro-ph | astro-ph |
Draft version September 19, 2017
Preprint typeset using LATEX style emulateapj v. 11/12/01
BOUNDS ON THE MASS AND ABUNDANCE OF DARK COMPACT OBJECTS AND BLACK
HOLES IN DWARF SPHEROIDAL GALAXY HALOS
F. J. S´anchez-Salcedo1 and V. Lora1
Draft version September 19, 2017
ABSTRACT
We establish new dynamical constraints on the mass and abundance of compact objects in the halo
of dwarf spheroidal galaxies. In order to preserve kinematically cold the second peak of the Ursa Minor
dwarf spheroidal (UMi dSph) against gravitational scattering, we place upper limits on the density of
compact objects as a function of their assumed mass. The mass of the dark matter constituents cannot
be larger than 103 M⊙ at a halo density in UMi's core of 0.35 M⊙ pc−3. This constraint rules out
a scenario in which dark halo cores are formed by two-body relaxation processes. Our bounds on the
fraction of dark matter in compact objects with masses & 3× 103 M⊙ improve those based on dynamical
arguments in the Galactic halo. In particular, objects with masses ∼ 105 M⊙ can comprise no more than
a halo mass fraction ∼ 0.01. Better determinations of the velocity dispersion of old overdense regions
in dSphs may result in more stringent constraints on the mass of halo objects. For illustration, if the
preliminary value of 0.5 km/s for the secondary peak of UMi is confirmed, compact objects with masses
above ∼ 100 M⊙ could be excluded from comprising all its dark matter halo.
Subject headings: dark matter -- galaxies: halos -- galaxies: individual (Ursa Minor dwarf spheroidal)
-- galaxies: kinematics and dynamics -- galaxies: structure
1.
introduction
The composition of dark halos around galaxies is a diffi-
cult problem. Many of the baryons in the universe are dark
and at least some of these dark baryons could be in galac-
tic halos in the form of very massive objects (VMOs), with
masses above 100 M⊙. Astrophysically motivated candi-
dates include massive compact halo objects (MACHOs)
and black holes either of intermediate mass (IMBHs; 101.3
to 105 M⊙) or massive (& 105 M⊙).
IMBHs are an in-
tringuing possibility as they could contribute, in principle,
to all the baryonic dark matter and may be the engines
behind ultraluminous X-ray sources recently discovered in
nearby galaxies.
A successful model
in which VMOs are the dom-
inant component of dark matter halos could resolve
some long-standing problems (Lacey & Ostriker 1985;
Tremaine & Ostriker 1999; Jin et al. 2005).
If the halos
of dSphs are comprised by black holes of masses between
∼ 105 and 106 M⊙, they evolve towards a shallower inner
profile in less than a Hubble time, providing an explana-
tion for the origin of dark matter cores in dwarf galaxies,
and the orbits of globular clusters (GCs) do not shrink to
the center by dynamical friction (Jin et al. 2005). Very few
observational limits on VMOs in dSph halos have been de-
rived so far. This Letter is aimed at constraining the mass
and abundance of VMOs in the halos of dSphs by the dis-
ruptive effects they would have on GCs and cold long-lived
substructures.
2. constraints from the survival of fornax gcs
GCs should be known. Since the three-dimensional dis-
tances of the GCs to the center of Fornax are unknown,
the density of VMOs at a mean distance of ∼ 1 kpc will
be adopted. By f we will denote the halo mass fraction
in VMOs, i.e. f ≡ ρh/ρdm, where ρdm is the halo density;
ρdm = 0.02 -- 0.05 M⊙ pc−3 at 1 kpc (Walker et al. 2006a).
Klessen & Burkert (1996) give the "survival diagram"
of GCs, considering different encounter histories for ρh =
0.026 M⊙ pc3 and a velocity dispersion of halo particles
σh = 120 km s−1. The survival diagram establishes the
range of mass and concentration such that GCs with cen-
tral densities between 103 M⊙ pc−3 and 104 M⊙ pc−3
(or, equivalently, core radii between 0.5 and 2 pc) have a
probability of less than 1% to survive after 10 Gyr. Since
the dissolution timescale for a certain GC (and a given
Mh) scales as ∝ ρ−1
h σh, and σh ≈ 20 km/s in Fornax, it
will be a factor of (5 -- 12)f less in Fornax as compared to
the Galactic case considered by Klessen & Burkert (1996).
Therefore, the survival diagram for Fornax's GCs can be
derived at once (see Fig. 1). GCs in the region above the
thick line have a probability of less than 1% to survive in
a dark halo with ρhMh = 103 M2
⊙ pc−3. For such a ρhMh
value and if the present parameters of the GCs are repre-
sentative of their parameters at the time they formed, we
would be observing the lucky survivors of an initial popu-
lation of & 500 GCs of ∼ 2 × 105 M⊙, which turns out to
be very unrealistic for a galaxy with a V-band luminosity
of 1.5 × 107 L⊙.
It is possible to estimate the probability that we
are observing only the survivors of a larger original
population that are in the process of quick disruption
(Tremaine & Ostriker 1999). The distribution of the disso-
lution age of GCs is expected to follow a scale-free power
law F = C(tdis + tH )−q, where tdis is the characteristic
dissolution timescale and tH is the age of the cluster pop-
ulation (Gnedin & Ostriker 1997). Fig. 1 shows that at
Fornax is a dark dominated dSph galaxy with an unusu-
ally high GC frequency for its dynamical mass. In order
to place dynamical constraints on the mass of VMOs by
requiring that not too many GCs are disrupted, the den-
sity of VMOs ρh with mass Mh along the orbits of the
1 Instituto de Astronom´ıa, Universidad Nacional Aut´onoma de M´exico, Ciudad Universitaria, Aptdo. 70 264, C.P. 04510, Mexico City, Mexico
([email protected], [email protected])
1
2
least 4 GCs are above the thick line; this indicates that
their present disruption timescales are < 0.22tH.
If the
exponent of distribution of lifetimes is q ∼ 2, as derived
for Galactic GCs (Gnedin & Ostriker 1997), the probabil-
ity to have 4 out of 5 GCs with lifetimes less than 0.22tH
is ∼ 1%, whereas the probability that the lifetimes of all
the GCs are less then 0.5tH is 0.4%. Hence the probabil-
ity that the whole dark halo is comprised of objects with
masses > 5× 104(ρh/0.02 M⊙pc−3)−1 M⊙ is less than 1%.
If only a fraction f of the dark mass is in compact objects
of mass Mh, then f < 5 × 104 M⊙/Mh.
3. persistence of dynamically-cold
subpopulations
Localized regions with enhanced stellar density and,
where data permit, extremely cold kinematics have been
detected in some dSphs (e.g., Olszewski & Aronson 1985;
Kleyna et al. 2003, hereafter K03; Coleman et al. 2004;
Walker et al. 2006b).
In particular, UMi dSph has re-
ceived the most attention. Collecting the velocity of stars
in 6′ radius aperture, K03 found that a two-Gaussian pop-
ulations, one representing the underlying 8.8 km/s Gaus-
sian and the other with velocity dispersion σs = 0.5 km/s,
representing a subpopulation of fraction 0.7, is > 3 × 104
times more likely than the default 8.8 km/s. The best-fit
σs is ill-determined as it is much smaller than the median
velocity errors (5 km/s). Nevertheless, even with these
fiducial errors, we can be certain that the velocity dis-
persion is < 2.5 km/s at ∼ 95% confidence level. The
stars which form the secondary density peak are not dis-
tinguished in colour and magnitude from the remainder of
the UMi population (Kleyna et al. 1998). In fact, UMi star
formation history indicates that its stars have been formed
in a single burst earlier than 10 Gyr ago (Carrera et al.
2002).
Although UMi has long been suspected of experienc-
ing ongoing tidal disruption, regions with enhanced vol-
ume density and cold kinematics cannot be the result
of tidal
interactions because the coarse-grained phase-
density, ∼ ρ/σ3, in collisionless sytems must be constant
or even decrease, thus implying that overdensity regions
should appear dynamically hotter. This suggests that the
clump is long-lived. An alternative explanation is that
the density peak is a projection effect and that what we
are seeing is a cold, low-density tidal tail. However, nu-
merical experiments have shown that this scenario is very
unlikely (Read et al. 2006). The most plausible interpre-
tation is that the clump is a disrupted stellar cluster, now
surviving in phase-space because the underlying gravita-
tional potential is harmonic (K03). Within this potential,
gravitational encounters with the hypothetical VMOs will
dominate the orbital diffusion of the stars once they be-
come unbound from the progenitor cluster. The integrity
of the cold clump may impose useful upper limits on the
mass of VMOs. The fact that the subpopulation is orbiting
within the dark matter core of UMi will greatly simplify
its dynamical description.
Clump's stars will undergo a random walk in momen-
tum space by the collisions with the population of VMOs.
Here we are interested in the velocity change induced in a
star relative to the clump center of mass. The mean-square
velocity change of a star in an encounter with a VMO of
mass Mh, and impact parameter b ≥ 5r1/2, with r1/2 the
clump's median radius, in the impulsive approximation is:
∆v2 =
2
b2V (cid:19)2
3 (cid:18) 2GMh
r2,
(1)
where V is the maximum relative velocity between the
clump and the perturber and r2 is the mean-square po-
sition of the stars in the clump (Spitzer 1958).
In the
opposite case of a head-on collision (b = 0), the mean
change is comparable to that predicted by the tidal ap-
proximation when b ≃ 1.4r1/2. The usual way to proceed
is to integrate ∆v2 given in Eq. (1) for impact parameters
b from 1.4r1/2 to infinity and correct for the encounters
in which the tidal approximation fails by a factor g ≈ 3
(e.g., Binney & Tremaine 1987; Gieles et al. 2006). Doing
so, and for a distribution of clumps and VMOs with a rel-
ative one-dimensional velocity dispersion σrel, we obtain:
∆v2 =
16√πgG2ρhMhr2∆t
9σrelr2
1/2
.
(2)
For UMi, the persistence of the clump for a large frac-
tion of a Hubble time indicates a core of the dark halo
of at least 2-3 times the size of the orbit of the clump,
which is & 150 pc.
In terms of the stellar core radius
(∼ 200 pc), this makes a halo core 1.5 -- 2 times the stellar
core and, consequently, the velocity dispersion of the halo
particles in the core is, at least ∼ 1.5 -- 2 times the stellar
velocity dispersion, corresponding to σh ∼ 15 -- 20 km/s.
The impulsive approximation is valid for bω ≤ V , where
ω = σs/r1/2 and σs is the internal one-dimensional velocity
dispersion of the subpopulation. For the encounters with
b . 5r1/2, responsible for most of the velocity impulse,
this condition is well satisfied for σh ≫ 5σs. Therefore,
for σs ∼ 1 km/s, this requirement is fulfilled within the
isothermal dark core of UMi.
Since stars in the clump are unbound, the self-gravity
of the clump in a first approximation can be ignored con-
sidering only orbit diffusion in the large-scale harmonic
potential of the parent galaxy. In a one-dimensional har-
monic potential, a velocity impulse ∆v2 produces a change
in the velocity dispersion ∆σ2 ≡ ∆(cid:10)v2(cid:11) = ∆v2/2, where
the brakets h...i refer to the mean value after averaging
over one orbit. Combining this relation with Eq. (2), we
find the change of σs in a time ∆t:
∆σ2
s =
8√πgG2ρhMhr2∆t
27σhr2
1/2
,
(3)
where σrel ≈ σh is assumed, since the population of clumps
is expected to have a velocity dispersion similar to the stel-
lar background, ∼ 9 km/s in UMi. The ratio η ≡ r2/r2
1/2
depends on the model:
for a Plummer cluster η = 4,
whereas η = 1.5 for both a King profile with a dimension-
less potential depth of W0 = 9 (e.g., Gieles et al. 2006) and
a Gaussian density distribution. In order to take a conser-
vative value and to facilitate comparison with photometric
and theoretical analysis that assume Gaussian models, we
will adopt η = 1.5 constant in time.
By t2.5 we will indicate the time required for a very cold
group of unbound stars σs ≃ 0, to acquire a velocity dis-
persion of 2.5 km/s. From Eq. (3) with g = 3 we then
obtain:
t2.5 = 5 Gyr(cid:18)
ρh
0.1 M⊙/pc3(cid:19)−1(cid:18) Mh
5 × 103 M⊙(cid:19)−1(cid:18) σh
20 km/s(cid:19) .
(4)
Since many of the recent dynamical models suggest
mean dark matter densities within the stellar core ra-
dius of UMi & 0.1 M⊙ pc−3, corresponding to a cen-
tral mass-to-light ratio of & 15 M⊙/L⊙ (Lake 1990;
Pryor & Kormendy 1990; Irwin & Hatzidimitriou 1995;
Mateo 1998; Wilkinson et al. 2006), ρdm = 0.1 M⊙ pc−3
will be accepted as a conservative reference value.
0 +
GMh
r2 = r2
in time with σs, according to the relation ¯r2 ≃ σ2
The mean-square radius of the subpopulation increases
s /Ω2,
where Ω is the orbital frequency in the constant-density
core. If the core is dominated by the dark matter compo-
nent ρdm/Ω2 = (4πG/3)−1, then
f
√π
If σs = 2.5 km/s, the mean radius √ ¯r2 ≃ σs/Ω ≈ 60 pc.
This corresponds to a 1 σ radius of 25.5 pc for a Gaussian
density profile, implying an angular size of almost 2′ at
the distance of UMi (∼ 66 kpc). This value is comparable
to but slightly larger than the observed 1 σ radius of the
secondary peak ≃ 1.6′. As a consequence σs . 2.5 km/s;
otherwise, the stellar subpopulation would appear more
extended and diffuse than it is observed. This upper value
is in agreement with the observations of K03.
(5)
σh
t.
Our estimates for the size and velocity dispersion of the
subpopulation are independent of the eccentricity of the
orbit because there is little variation of the macroscopic
properties of the halo, ρh, σh or Ω, within the core where
the clump is orbiting. Nevertheless, the dark halo could
have suffered significant evolution due to two-body pro-
cesses and tidal stirring. For Mh . 105 M⊙ relaxation
processes induce an insignificant change in the internal
properties of the halo (e.g., Jin et al. 2005). Tidal heating
can lead to a reduction of both the density of dark mat-
ter particles and its velocity dispersion (e.g., Mayer et al.
2001; Read et al. 2006). Since the phase-space density,
ρh/σ3
h, for collisionless systems is nearly constant or de-
creases with time, the rate of energy gained by the clump
due to encounters with VMOs should have been more in-
tense in the past. The inclusion of evolution of halo prop-
erties by tidal effects would lead to a stringent upper mass
limit.
If the stellar progenitor cluster became unbound imme-
diately after formation when supernovae expel the gas con-
tent (Goodwin 1997; K03), t2.5 should be greater than the
age of the cluster tH ∼ 10 Gyr. This requirement com-
bined with Eq. (4) implies:
Mh . 2.5 × 103 M⊙(cid:18)
0.1 M⊙/pc3(cid:19)−1(cid:18) σh
20 km/s(cid:19) .
(6)
ρh
If initially the stellar cluster is gravitationally bound,
the increase in the internal energy gained by encounters
with VMOs will eventually exceed its binding energy af-
ter a time tbe. Let us estimate tbe, the time at which
the cluster becomes unbound. K03 infer a total mass
of the cluster, Mcl, of 3 × 104 M⊙.
If this cluster fol-
lowed the recently observed relation between radius and
mass of Larsen (2004), r1/2 ≈ 4 pc. Adopting the refer-
ence values of the dark matter halo (ρdm = 0.1 M⊙ pc−3
3
and σh = 20 km/s) and rescaling the survival diagram of
Klessen & Burkert (1996) (their figure 11) for the param-
eters of UMi, we infer that more than 95% of the clusters
with mass 3 × 104 M⊙ will become unbound after tbe ≈ 3
Gyr if Mh ≥ 3.5× 103f −1 M⊙. Therefore, in order to have
a dynamically cold subpopulation with σs < 2.5 km/s, as
that observed in UMi at the present time tH ∼ 10 Gyr,
we need t2.5 ≥ tH − tbe, which implies the following upper
limit for Mh
Mh . 3.5 × 103 M⊙(cid:18)
0.1 M⊙/pc3(cid:19)−1(cid:18) σh
20 km/s(cid:19) .
(7)
ρh
Other corrosive effects such as mass loss by stellar evolu-
tion or tidal heating may also accelerate the dissolution of
the cluster. Therefore, our estimates for tbe and, hence,
for Mh, are upper limits.
The survival probability after collisions with VMOs in-
creases for progenitors that are more compact. For in-
stance, if the progenitor were a supercluster with a core
radius rc ≈ 0.5 pc and central density ∼ 3× 104 M⊙ pc−3,
the probability of its remaining gravitationally bound after
6 Gyr is ∼ 25%, for Mh = 6.5× 103f −1 M⊙. Hence, there
may be a non-negligible probability that such a superclus-
ter has survived bound for 6 Gyr and that during the sub-
sequent 4 Gyr it is dynamically heated by 6.5 × 103f −1
M⊙ VMOs to reach σs = 2.5 km/s at the present time.
Unfortunately, the evaporation time for this supercluster,
setting the scale for dynamical dissolution by internal pro-
cesses, is very short.
In fact, for such a stellar cluster,
the evaporation time is ∼ 20trh, with trh the half-mass
relaxation time (Gnedin & Ostriker 1997). The resulting
evaporation time is . 1 Gyr for an average stellar mass
≥ 1 M⊙ and, hence, internal processes would have pro-
duced a fast desintegration of such a cluster. We conclude
that the upper limit given in Eq. (7) is realistic and robust.
Our approximations break down when the halo only con-
tains a few VMOs; at least 5 objects within a radius of 600
pc are required, implying that our analysis is restricted to
masses Mh < 2 × 107f M⊙. In Fig. 2, the observational
limits on VMOs over a wide range of masses and dark
matter fractions are shown.
4. discussion and conclusions
The analysis of the survival of Fornax's GCs rules out
the mass range that would be interesting for explaining
the origin of dark matter cores in dwarf galaxies (K03;
Goerdt et al. 2006; S´anchez-Salcedo et al. 2006), because
the relaxation timescale for VMOs of mass ≤ 5 × 104 M⊙
exceeds the Hubble time. Moreover, it was found that
the integrity of cold small-scale clustering seen in some
dSphs imposes more stringent constraints on the mass of
VMOs. A source of uncertainty is the mean density of
dark matter within the core of dSphs. In the particular
case of UMi and according to the scaling relations com-
piled in Kormendy & Freeman (2004), the corresponding
central dark matter density is 0.35 M⊙ pc−3. A slightly
larger value has been derived from its internal dynam-
ics (Wilkinson et al. 2006). At a density ρdm = 0.35
M⊙ pc−3, Eq. (7) implies that Mh . 1000f −1 M⊙. We
strongly encourage better determinations of the velocity
dispersion of cold density aggregates (bound or unbound)
in dSphs. For instance, if the preliminary quoted value of
0.5 km/s for the secondary peak of UMi were confirmed,
4
our upper limit for Mh would be immediately reduced by
a factor of 25, implying a very tight bound Mh . (40 --
120)f −1 M⊙, depending on the adopted dark matter den-
sity (0.3 M⊙ pc−3 to 0.1 M⊙ pc−3).
In a unified scheme such as the 'stirring scenario' by
Mayer et al. (2001), the composition of the dark halos of
low-surface brightness and dSph galaxies should be the
same. Rix & Lake (1993) found an upper limit of 104 M⊙
by examining the dynamical heating of the stellar disk
of GR 82. Tremaine & Ostriker (1999) warn about the
weakness of the argument of Rix & Lake (1993) because
the rotation curve of GR 8 decays in a Keplerian fashion
suggesting that GR 8 does not host a halo of dark matter,
as expected if the dark halo has evaporated by two-body
collisions. However, there exist dwarf galaxies with flat ro-
tation curves. From the sample of Hidalgo-G´amez (2004),
we have selected dwarf galaxies with high inclination an-
gles and estimated the maximum permitted value of Mh
consistent with the thickness of the old stellar disk. Per-
haps one of the most pristine cases is the edge-on galaxy
UGCA 442. This galaxy shows the typical flat rotation
curve, implying the existence of a halo of dark matter with
a central density of 0.07 M⊙ pc−3 and a velocity dispersion
σh = vc/√2 ≃ 35 km/s (Cot´e et al. 2000). The vertical
stellar velocity dispersion has been derived using vertical
hydrostatic equilibrium σ2
z = πGΣh, with Σ the total cen-
tral surface density of the disk ∼ 65 M⊙ pc−2 and h the
scale height of the old disk derived photometrically to be
350 pc. This yields σz = 18 km/s. By requiring that
∆σz < σz for ∆t > 1.5 Gyr, which is the characteristic
age of the old stellar population, the disk-heating argu-
ment establishes Mh . 8 × 104f −1 M⊙ (Lacey & Ostriker
1985).
Fig. 2 contains the most relevant dynamical constraints
on VMOs.
In the conservative case σs = 2.5 km/s, the
most stringent bound for masses between 102 M⊙ and 103
M⊙ comes from studies of the distribution of separations of
wide stellar binaries in the Galactic halo (Yoo et al. 2004).
Still, masses in that range are permitted if halo objects are
slightly extended (sizes & 0.05 pc). Other sources of un-
certainty in this approach are the orbital distribution of
the binaries and the duration of perturbations they are
subjected to (Jin et al. 2005).
The valuable suggestions by an anonymous referee
greatly improved the paper. We are indebted to
J. A. Garc´ıa Barreto for his help.
REFERENCES
Binney, J., & Tremaine, S. 1987, Galactic Dynamics, (Princeton
University Press: New York)
Bissantz, N., Englmaier, P., & Gerhard, O. 2003, MNRAS, 340, 949
Carrera, R., Aparicio, A., Mart´ınez-Delgado, D., & Alonso-Garc´ıa,
J. 2002, AJ, 123, 3199
Coleman, M., Da Costa, G. S., Bland-Hawthorn, J., Mart´ınez-
Delgado, D., Freeman, K. C., & Malin, D. 2004, AJ, 127, 832
Cot´e, S., Carignan, C., & Freeman, K. C. 2000, AJ, 120, 3027
Gieles, M. et al. 2006, MNRAS, 371, 793
Gnedin, O. Y., & Ostriker, J. P. 1997, ApJ, 474, 223
Goerdt, T., Moore, B., Read, J. I., Joachim, S., & Zemp, M. 2006,
MNRAS, 368, 1073
Goodwin, S. P. 1997, MNRAS, 286, 669
Hidalgo-G´amez, A. M. 2004, RevMexAA, 40, 37
Irwin, M., & Hatzidimitriou, D. 1995, MNRAS, 277, 1354
Jin, S., Ostriker, J. P., & Wilkinson, M. I. 2005, MNRAS, 359, 104
Klessen, R., & Burkert, A. 1996, MNRAS, 280, 735
Kleyna, J. T., Geller, M. J., Kurtz, M. J., & Thorstensen, J. R. 1998,
AJ, 115, 2359
Kleyna, J. T., Wilkinson, M. I., Gilmore, G., & Evans, N. W. 2003,
ApJ, 588, L21
Kormendy, J., & Freeman, K. C. 2004, in Dark Matter in Galaxies,
ed. S. D. Ryder et al., IAU Symp. 220 (San Francisco: ASP), 377
Lacey, C. G., & Ostriker, J. P. 1985, ApJ, 299, 633
Lake, G. 1990, MNRAS, 244, 701
Larsen, S. S. 2004, A&A, 416, 537
Mackey, A. D., & Gilmore, G. F. 2003, MNRAS, 340, 175
Mateo, M. 1998, ARA&A, 36, 435
Mayer, L. et al. 2001, ApJ, 559, 754
Olszewski, W. W., & Aronson, M. 1985, AJ, 90, 2221
Pryor, C., & Kormendy, J. 1990, AJ, 100, 127
Read, J. I., Wilkinson, M. I., Evans, N. W., Gilmore, G., & Kleyna,
J. T. 2006, MNRAS, 367, 387
Rix, H.-W., & Lake, G. 1993, ApJ, 417, L1
S´anchez-Salcedo, F. J. 1999, MNRAS, 303, 755
S´anchez-Salcedo, F. J., Reyes-Iturbide, J., & Hernandez, X. 2006,
MNRAS, 370, 1829
Spitzer, L. Jr. 1958, ApJ, 127, 17
Tremaine, S., & Ostriker, J. P. 1999, MNRAS, 306, 662
Walker, M. G., Mateo, M., Olszewiski, E. W., Bernstein, R. A.,
Wang, X., & Woodroofe, M. 2006a, AJ, 131, 2114
Walker, M. G., Mateo, M., Olszewiski, E. W., Pal, J. K., Sen, B., &
Woodroofe, M. 2006b, ApJ, 642, L41
Wilkinson, M. I. et al. 2006,
in Mass Profiles and Shapes of
Cosmological Structures, eds. G.A. Mamon, F. Combes, C.
Deffayet, B. Fort, (EAS Publications Series), Volume 20, 105
Yoo, J., Chanam, J., & Gould, A. 2004, ApJ, 601, 311
2 The value has been updated according to the discussion in S´anchez-Salcedo (1999), §1.
5
Fig. 1. -- Survival diagram for GCs in Fornax's halo for ρhMh = 103 M2
⊙ pc−3. If GCs had core radii in the range from 0.5 to 2 pc, their
probability of survival after 10 Gyr would be less than 1% in the region left and above the thick line. The core radii of Fornax 3, 4 and 5 lie
within the mentioned range, but Fornax 1 and 2 present core radii of 10 and 5.6 pc, respectively (Mackey & Gilmore 2003). Therefore, the
probability of survival for the latter GCs will be much less than 1%.
6
Fig. 2. -- Observational constraints on VMOs from MACHO microlensing experiments, the distribution of wide binaries in the thoroughly
validated Bassel mass model of the Milky Way (Bissantz, Englmaier & Gerhard 2003), the heating of the Galactic disk, the heating of the
stellar disk of UGCA 442, and the survival of UMi's dynamical fossil in UMi with ρdm = 0.35 M⊙ pc−3.
|
astro-ph/0411780 | 1 | 0411 | 2004-11-30T00:34:17 | Identification of an Extended Accretion Disk Corona in the Hercules X-1 Low State: Moderate Optical Depth, Precise Density Determination, and Verification of CNO Abundances | [
"astro-ph"
] | We identify an accretion disk atmosphere and corona from the high resolution X-ray spectrum of Hercules X-1, and we determine its detailed physical properties. More than two dozen recombination emission lines (from Fe XXVI at 1.78 A to N VI at 29.08 A) and Fe K-alpha, K-beta fluorescence lines were detected in a 50 ks observation with the Chandra High-Energy Transmission Grating Spectrometer (HETGS). They allow us to measure the density, temperature, spatial distribution, elemental composition, and kinematics of the plasma. We exclude HZ Her as the source of the recombination emission. We compare accretion disk model atmospheres with the observed spectrum in order to constrain the stratification of density and ionization, disk atmosphere area, elemental composition, and energetics. The atmospheric spectrum observed during the low state is photoionized by the main-on X-ray continuum, indicating that the disk is observed edge-on during the low state. We infer the mean number of scatterings N of Ly-alpha and Ly-beta line photons from H-like ions. We derive N < 69 for O VIII Ly_alpha_1, which rules out the presence of a mechanism modeled by Sako (2003) to enhance N VII emission via a line overlap with O VIII. The line optical depth diagnostics are consistent with a flattened atmosphere. Our spectral analysis, the disk atmosphere model, and the presence of intense N VII and N VI lines (plus N V in the UV), confirm the over-abundance of nitrogen relative to other metals, which was shown to be indicative of CNO cycle processing in a massive progenitor. | astro-ph | astro-ph |
Accepted to ApJ
Identification of an Extended Accretion Disk Corona in the Hercules X-1 Low
State: Moderate Optical Depth, Precise Density Determination, and
Verification of CNO Abundances
M. A. Jimenez-Garate
MIT Center for Space Research, 70 Vassar St (NE80-6009), Cambridge, MA 02139
[email protected]
J. C. Raymond
Center for Astrophysics, 60 Garden St., Cambridge, MA 02138
[email protected]
D. A. Liedahl
Lawrence Livermore National Laboratory, Department of Physics and Advanced Technologies, 7000 East
Ave., L-41, Livermore, CA 94550
[email protected]
and
C. J. Hailey
Columbia Astrophysics Laboratory, 538 W. 120th St., New York, NY, 10027
[email protected]
ABSTRACT
We identify an accretion disk atmosphere and corona from the high resolution X-ray spectrum
of Hercules X-1, and we determine its detailed physical properties. More than two dozen recom-
bination emission lines (from Fe XXVI at 1.78 A to N VI at 29.08 A) and Fe Kα,Kβ fluorescence
lines were detected in a 50 ks observation with the Chandra High-Energy Transmission Grating
Spectrometer (HETGS). They allow us to measure the density (ne = 2±1×1013 cm from Mg XI),
temperature (kT = 7 ± 3 eV from Ne IX), spatial distribution, elemental composition, and kine-
matics (∆v . 260 km s−1 ) of the plasma. We exclude HZ Her as the source of the recombination
emission. We compare accretion disk model atmospheres with the observed spectrum in order
to constrain the stratification of density and ionization, elemental composition, energetics, and
thermal stability. The derived disk atmosphere and corona radii are 8 × 1010 . r . 1 × 1011 cm,
in agreement with previously measured eclipse ingress light curves. The atmospheric spectrum
observed during the low state is photoionized by the main-on X-ray continuum, indicating that
the disk is observed edge-on during the low state. We infer the mean number of scatterings hN i
of Lyα and Lyβ line photons from H-like ions. We derive hN i . 69 for O VIII Lyα1, which
rules out the presence of a mechanism modeled by Sako (2003) to enhance N VII emission via
-- 2 --
a line overlap with O VIII. The line optical depth diagnostics are consistent with a flattened
atmosphere. Our spectral analysis, the disk atmosphere model, and the presence of intense N VII
and N VI lines (plus N V in the UV), confirm the over-abundance of nitrogen relative to other
metals, which was shown to be indicative of CNO cycle processing in a massive progenitor. The
spectral signatures of a thermal instability in the photoionized plasma are not evident, but the
measured density is in the stable regime of the models.
Subject headings: X-rays: binaries -- line: formation -- line: identification -- pulsars: individual
(Her X-1) -- accretion, accretion disks -- binaries: eclipsing
1.
Introduction
High-resolution spectroscopic observations of X-ray binaries have shown in some cases the presence of
extended X-ray emitting plasmas or outflows surrounding the accretion disk. The detection of these compo-
nents and their kinematic properties has been made possible by the observation of emission or absorption
lines from very highly ionized metals. In the case of Hercules X-1, a bright intermediate-mass X-ray binary
with Porb = 1.7 day orbital period and a Ppulse = 1.24 s X-ray pulsar (Tananbaum et al. 1972; Wilson,
Scott, & Finger 1997), observations with the XMM-Newton Reflection Grating Spectrometer (RGS) showed
a strong 35 d phase dependence in the high-resolution spectrum (Jimenez-Garate et al. 2002). The PΨ ∼ 35 d
pseudo-periodic cycle (Giacconi et al. 1973) has been associated with a tilted and precessing accretion disk
(Petterson 1975; Petterson, Rothschild, & Gruber 1991; Scott, Leahy, & Wilson 2000). Through the 35 d
cycle, the X-ray light-curve is asymmetric and contains two maxima: a state of ∼ 8 d duration reaching
the peak flux Fmax named the main-on, and a secondary high state of ∼ 4 d duration reaching ∼ 1/3 Fmax
named the short-on. A low state with ∼ 1/20 Fmax ensues at other epochs. The main-on spectrum exhibits
a strong continuum, with evidence for weak and very broad emission lines, while the low state shows bright
and narrow emission lines dominating over a very weak continuum. The short-on is intermediate, with a
moderate continuum flux and more intense emission lines than the low state (Jimenez-Garate et al. 2002).
The Fe Kα line follows a similar pattern. XMM-Newton EPIC data show a 6.4 keV Fe Kα line which is
practically unresolved during the low and short-on states, and a broad line at 6.5 keV with 330 ± 20 eV
FWHM during the main-on (Ramsay et al. 2001). Matter close to the magnetosphere and the pulsar appears
to be observable during the main-on only. This behavior supports the picture of a precessing accretion disk,
since the spectroscopic appearance of the disk atmosphere and the pulsar should be highly dependent on
inclination. Analysis of the RGS spectrum showed a clear over-abundance of N with respect to Ne, as well
as a moderate under-abundance of O and C with respect to Ne. The N over-abundance was proposed as a
signature of CNO cycle nucleosynthesis occurring at an early phase in the binary system, and it is indicative
of a star twice or more times more massive than the secondary/companion (Jimenez-Garate et al. 2002).
This is indicative of a massive progenitor of the system and a period of extreme mass loss, as had been pre-
dicted by evolutionary models of LMXBs (Podsiadlowski, Rappaport, & Pfahl 2002). Sako (2003) suggested
that the N over-abundance is not real, but that it is due instead to an X-ray Bowen fluorescence effect which
occurs at large line optical depths. In this article, we quantify the relevant line optical depths to distinguish
between the X-ray Bowen effect or the abundance anomaly scenarios.
Hercules X-1 has been observed extensively. Optical light-curves (Gerend & Boynton 1976) and X-ray
eclipses (Tananbaum et al. 1972) yield a Porb = 1.7 day orbital period. The 1.5 ± 0.3 M⊙ neutron star has a
2.3 ± 0.3 M⊙ companion, HZ Her, which changes from A to B spectral-type over the orbital period, due to
the strong X-ray illumination on its surface (Reynolds et al. 1997). The unabsorbed luminosity of Her X-1 is
-- 3 --
L = 3.8 × 1037 ergs−1, using a distance of D = 6.6 ± 0.4 kpc (Reynolds et al. 1997). The Her X-1 broadband
X-ray spectrum during the main-on state consists of a blackbody component with temperature kT ∼ 90 eV,
plus a power-law component with a 24 keV exponential cutoff, and a 42 keV cyclotron feature (dal Fiume et
al. 1998). The X-ray light-curve, the variations in the pulse profiles (Deeter et al. 1991), and the variability
of the dips (Giacconi et al. 1973; Scott & Leahy 1999), are fit by a geometric model of a precessing, warped
accretion disk, with an 85◦ inclination with respect to the line of sight, a 20◦ precession opening angle for
the outermost disk, and an 11◦ precession angle for the innermost disk (Scott, Leahy, & Wilson 2000). The
ultraviolet (UV) spectrum exhibits line emission from C V, N V, and O V, which have two separate velocity
components. A narrow UV line component is thought to originate on the illuminated face of HZ Her. The
broad UV line region likely originates in a prograde accretion disk of ∼ 1011 cm radius (Boroson et al. 2000).
One of the questions to be resolved is whether the X-ray line emission has a similar origin to the UV. The
broadening of the X-ray line emission indicates that it can originate at radii similar to the UV. For the case
of an irradiated disk atmosphere, the UV line region is denser and deeper than the X-ray line region.
In this work, we interpret the broadband (1.5 A to 30 A) high-resolution X-ray spectrum of Her X-1 and
investigate the nature of the line emitting region(s). The Chandra grating spectrum allows us to detect a
wealth of spectral features which had not been previously observed in Her X-1. Our data reduction procedure
is described in §2. To interpret the spectrum, we use two complementary approaches: 1) an analytic approach
for constraining the physical parameters of the plasma, based solely on the photoionized plasma physics;
and 2) a synthetic approach, by comparing the spectrum with that of an astrophysical model. In §3, discuss
the origin of the Fe Kα line and we perform spectral diagnostics on density, kinematics, and location from
the recombination lines. In §4, we utilize standard photoionized plasma models and radiative recombination
rates to perform an emission measure analysis, and to quantify the ionization distribution of the plasma, the
emitting volume, and the ratios among the metal abundances. We also compare the model temperatures
with those measured from radiative recombination continua (RRC). In §5, we quantify the mean number
of scatterings of selected lines, which shows that the optical depth is not high enough for the X-ray Bowen
fluorescence effect to take place in Her X-1. Line depth diagnostics also provide evidence for an anisotropic
plasma distribution. In §6, we identify the nature of the X-ray emission region on the basis of a comparison
of an accretion disk and corona model with the observed spectrum. In §7, we discuss the thermal stability
of the plasma. Finally, in §8 we compare the Her X-1 spectrum with that of Accretion Disk Corona (ADC)
sources, and we conclude in §9.
2. Chandra Observations
Hercules X-1 was observed on 2002 May 5 at 10:15 UT with the Chandra High Energy Transmission
Grating Spectrometer (Canizares et al. 2000, HETGS). The observation lasted 49.4 ks. Data from the All-Sky
Monitor (ASM) instrument onboard the Rossi X-ray Timing Explorer (RXTE), indicated that Her X-1 was
in the low state (Ψ = 0.44 − 0.46) while at mid-orbit (φ = 0.33 − 0.67) during the Chandra observation. From
the RXTE ASM data, we determined that first main-on start signal immediately preceding the Chandra
observation occurred on MJD 52384.22. Scott & Leahy (1999) observed that main-on turn-ons occur only
orbital phases 0.23 or 0.68, possibly due to a disk-orbit locking mechanism. The closest, most likely main-on
start time occurred at orbital phase 0.23, at MJD 52383.956, ∼ 10 hr earlier than the RXTE ASM detection.
We processed the HETGS data with the CIAO analysis software version 3.0. We obtained Medium
Energy Grating (MEG) and High Energy Grating (HEG) spectra, which have energy resolution FWHM of
∆λ = 0.023 A and ∆λ = 0.012 A, respectively. The ACIS-S CCD was set in FAINT mode since the low
-- 4 --
state does not produce pile-up in the image, and this minimizes the background in the spectrum. We use
the destreak tool to clean the ACIS-S image. The mkrmf and mkarf tools are used to produce the response
matrices and effective area applicable to our observation. We use the calibration data CALDB version 2.26,
which was updated 2004 Feb 2 to correct for the contamination layer on ACIS-S. We used the ISIS spectral
analysis software in our model fitting and flux measurements (Houck & Denicola 2000). The lightcurve
routine was used to extract the time variability of the flux. Her X-1 was placed at a Y Offset of 0.33 arcmin,
which is practically on-axis.
3. Spectral Analysis
The HEG and MEG spectra of Her X-1 exhibit emission lines throughout the X-ray band. Together,
the spectra exhibit bright lines from the H-like and He-like ions of N, O, Ne, Mg, and Si, plus the lines from
H-like ions of S and Fe. The MEG and HEG counts spectra in Figure 1 show a strong Fe Kα fluorescence
line, as well as numerous recombination emission lines. The HEG spectrum in Figure 2 highlights the Fe
Kα line from low ionization states and a line from the high ionization state Fe XXVI. The spectra of HEG
and MEG are added for display purposes in Figure 3, and the MEG spectrum at the high wavelength end
is shown in Figure 4. The detection of RRCs, as well as the fact that weak Fe L lines are observed, are
characteristic of photoionized plasmas (Liedahl & Paerels 1996; Liedahl 1999). We measure a 0.5-7 keV flux
of F0.5−7 = 3.3 × 10−11 erg cm−2 s−1 , and the fluxes of individual spectral features are listed in Table 1. We
use Cash statistics to fit the individual spectral features and obtain the statistical errors. Continuum fit
parameters are shown in Table 2.
3.1. Fe fluorescence
The data allow us to fit independently the Kα1 and Kα2 lines, which are merged in a single feature.
We also detect the Kβ line. The measured Fe K fluorescence line energies, as well as the line ratios among
Kα1, Kα2, and Kβ, all show that the iron is neutral, based on the Kaastra & Mewe (1993) calculations.
The Kα/Kβ ratio is 6.2 ± 1.6 (compared to 7.99 for Fe I from Kaastra & Mewe), while the Kα1/Kα2 ratio
is 2.28 ± 0.45 (compared to 2.00, idem). Any iron ion with an L-shell electron can fluoresce with high yields
as well. However, the line energy observed corresponds to Fe I -- Fe IX (Palmeri et al. 2003), and likely up to
Fe XIII. We approximate the 6.4 keV fluorescence line flux (in ph cm−2 s−1 ) with
F ≈ 1
2 YFe f T (cid:16) Ω
4π(cid:17)Z 200 keV
7.1 keV
FE (1 − e−τFe) dE
(1)
where we take FE to be the continuum flux (in ph cm−2 s−1 keV−1) observed during the main-on by dal
Fiume et al. (1998), τFe is the optical depth of M-shell iron, YFe = 0.34 is the fluorescence yield of Fe, Ω is
the solid angle subtended by the fluorescing plasma from the vantage point of the X-ray pulsar, 0 < T < 1
is a transmission coefficient due to absorption from species other than Fe I -- Fe XIII and Compton scattering,
and f is the fraction of the fluorescing region which is visible to us. Roughly half the fluorescent photons
are emitted downward into the optically thick gas. The 0.1 to 100 keV continuum flux during the main-on
is R FEdE = 6.7 × 10−9 erg cm−2 s−1 , and it is taken from the Model 2 fit by dal Fiume et al. (1998), which
consists of an absorbed blackbody plus broken power law, with an exponential cutoff and cyclotron absorption
line. Using τFe ∼ 4, which is an estimate of the depth from which an Fe K photon can escape, we obtain
f T Ω ∼ 0.22. The solid angle subtended by the companion star is Ω ∼ 0.50, and the solid angle subtended by
-- 5 --
the disk photosphere is Ω ∼ 1.0, as obtained from the illuminated disk model (Jimenez-Garate, Raymond, &
Liedahl 2002). These Ω are quite similar. Both cases produce f T ∼ 0.2 -- 0.4. Given 0.33 < φ < 0.67 during
the observations, we estimate that f ∼ 2/3 of the illuminated star is visible. For the disk, f . 1/2 for the
scenario of a self-shielding edge-on disk (see next section). In both cases reasonable transmission T & 0.5
can be obtained. Thus, we need to employ the orbital variability to distinguish amongst these components
(§3.4). The Fe Kα fluorescence line flux and broadening are much larger during the main-on than during
the short-on, and the low state line flux measured with XMM-Newton, of (2 -- 12) × 10−4 ph cm−2 s−1 (Zane
et al. 2004), brackets the Fe Kα line flux we measure with HETGS (see Table 1).
3.2. Unbroadened Lines
The gratings do not resolve any Doppler width in the emission lines. The Doppler broadening (∆v)σ is
measured by fitting Gaussian profiles to the brightest lines in the spectra, such that (∆v)FWHM ≃ 2.35(∆v)σ.
Table 1 shows that the best upper limits on ∆v are placed on the lines with largest λ. The observed line
broadening is entirely attributed to their doublet nature. We set 90 % confidence limits of (∆λ)σ = 0 -- 8 mA
for the O VII i line with MEG, (∆λ)σ = 8 -- 17 mA for the O VIII Lyα line with MEG, and (∆λ)σ = 0 -- 3 mA
for the Ne IX i line with HEG. The remaining bright lines have similar (∆λ)σ. The intrinsic line broadening,
plus any residual instrumental broadening for MEG and HEG not accounted for in the response matrix,
is quantified using the spectra of a bright star from which no velocity broadening is expected. We choose
the star HR 1099 as reference, which yields (∆λ)σ = 4.8 ± 0.4 mA with HEG and (∆λ)σ = 5.2 ± 0.6 mA
with MEG at 12.13 A (99% confidence limits). By fitting a plasma model, we verify that this broadening
can be fully accounted for by the Ne X Lyα-doublet, which is separated by 5 mA and should show a 2:1
intensity ratio. The doublet fully accounts for the measured (∆λ)σ. Other lines in the HR 1099 spectrum
are consistent with this. The upper limits on the velocity broadening on Table 1 are based on either the
measured or reference (∆λ)σ, whichever is largest. At the slightly lower resolving power of the XMM-Newton
RGS, the short-on and low state emission lines of Her X-1 were unresolved (Jimenez-Garate et al. 2002).
The absence of a Doppler velocity broadening constrains the dynamics of the disk atmosphere and the
geometry of the disk. Since the disk inclination is i ∼ 85◦, the orbital velocity of the disk is given by
v ∼ pGM/R cos Φ, where G is the gravitational constant, M the neutron star mass, R the disk radius,
and Φ is the azimuth on the disk. If the lines produced by a Φ-symmetric disk, translating the measured
(∆v)σ into a physical size of the disk requires modeling the line profiles with a calculated (or assumed)
emissivity versus radius. The disk lines would have double-peaked profiles. The profiles from a centrally
illuminated disk of 108.5 < r < 1011 cm radius in full view calculated with the Jimenez-Garate, Raymond,
& Liedahl (2002) model can be fit with a (∆v)σ ∼ 750 km s−1 Gaussian. Clearly the disk cannot be in
full view in the Her X-1 low state. This is an indication that we are observing the outer rim of the disk
atmosphere and corona. For a Φ-symmetric, edge-on, and optically thin disk, the emission region radius
would be r & GM/[4(∆v)2
σ]. For the X-ray lines observed at highest resolving power, the inferred lower
limits with this equation are r & (2 -- 3) × 1011 cm, larger than the neutron star Roche lobe and the disk size
r ∼ 1.4 × 1011 cm deduced from the eclipse ingress light curves of the He II, C IV, Si IV, and O V emission
lines in the UV (Chiang 2001; Boroson et al. 2000). Therefore, the disk atmosphere is likely asymmetric. One
interpretation is that due to the flared disk geometry, the disk shields itself, such that we only observe the
far side of the disk, which causes the cos Φ factor to reduce ∆v. This does not require an asymmetric disk,
just one that partially shields itself. Another interpretation is that the disk is asymmetric or warped, which
can similarly reduce the observed ∆v. Evidence of shielding in a warped disk was found by Chiang (2001)
-- 6 --
in the UV emission line profiles. A third possibility is that the disk is Φ-symmetric, but has an atmosphere
which is orbiting at sub-Keplerian velocities. The case of the asymmetric disk implies a modulation of the
X-ray line fluxes with orbital phase (see §3.4).
3.3. A location and density diagnostic from He-like ions
The Heα line triplets are shown in Fig. 5. At the HETGS resolution, each Heα complex consists of the
intercombination (i), forbidden (f ), and resonance (r) lines. The He-like triplet diagnostics were performed
with the XMM-Newton RGS spectrum for N VI, O VII, and Ne IX (Jimenez-Garate et al. 2002). Both RGS
and HETGS spectra show that the R = f /i flux ratio is zero for the latter ions. The HETGS spectrum
reveals that the Mg XI and Si XIII Heα lines behave differently: the R ratio is nonzero and increasing with
Z (see Table 3). The value of the R ratio depends on the atomic kinetics in each ion. The f line can get
converted into the i line by a process in which the metastable 1s2s 3S1 level is excited to 1s2p 3P1,2 (Porquet
& Dubau 2000). This occurs when the excitation rate wf→i is larger than the decay rate wf of 1s2s 3S1 to
the ground state. We note that the particular 1s2s 3S1 → 1s2p 3P2 transition does not modify the R ratio
for the low and mid-Z ions, for which 1s2p 3P2 decays back quickly to 1s2s 3S1, instead of decaying to the
ground state. An R ∼ 0 ratio can be produced by either collisional excitation above a critical electron density
ncrit
(Porquet & Dubau 2000; Bautista & Kallman 2000) or by photoexcitation above a critical UV-photon
flux F crit
(Mewe & Schrijver 1978). We discuss both as limiting cases for which R → 0. If both the plasma
density and the ambient UV flux are below their thresholds, the R ratio reaches an asymptotic value of
R & 2.2, which increases with T and depends weakly on Z. For Mg XI at T = 3 × 105 K, the maximum
R = 2.8 (Porquet & Dubau 2000).
ν
e
We first discuss the limiting case in which the density is below threshold and the UV flux is above
threshold. From the measured UV flux Fν, we constrain the distance (d) between the X-ray line emission
region and the UV source. The dominant source of UV flux is the surface of the companion (see below). A
similar diagnostic was used by Kahn et al. (2001). Jimenez-Garate et al. (2002) set a limit of d < 7 × 1011 cm
for Her X-1. With the newly detected lines, d is shown to be equal or larger than the binary separation. To
measure d, we quantify the photoexcitation rate by using the UV fluxes that were observed with the Hubble
Space Telescope (HST ) an the Far Ultraviolet Spectroscopic Explorer or FUSE (Boroson et al. 1996, 2000,
2001), and with the Hopkins Ultraviolet Telescope (Boroson et al. 1997). The relevant photoexcitation rate
is:
wf→i =
Fνf→i fosc,
(2)
πe2
mec
where e, me are the electron charge and mass, c is the speed of light, fosc is the oscillator strength, and Fνf→i
is the flux (in photons s−1 cm−2 Hz−1) at frequencies resonant with the 1s2s 3S1 → 1s2p 3P1 transition
in the UV. We use the fosc calculated with the HULLAC atomic code (Klapisch et al. 1977), and those
calculated by Cann & Thakkar (1992). Since Fνf→i is a function of d, we constrain d by relating wf→i to wf .
In the second limiting case, the UV flux is below threshold and the density is above threshold. In this
case, the lower limit of ne can be obtained from the upper limit on R using the Porquet & Dubau (2000)
calculations. The d calculated with Eqs. 2 and 3 in the first limiting case, as well as the inferred ne in the
second limiting case, are shown in Table 3.
A valuable new constraint is inferred from the Mg XI Heα lines, which show a critical R ratio, that is,
the value of R = 0.52 ± 0.13 is intermediate between the asymptotic values of R = 0.0 and R = 2.8. The
measured R ratio implies that
-- 7 --
wf→i ∼ wf .
(3)
For He-like ions other than Mg XI, Eq. 3 becomes an inequality. We interpret the Mg XI R ratio in two
limits and in the general case. In the low density limit, the UV flux is at a critical value, and Fν is measured.
In the low Fν limit, the density is at a critical value, and the measurement of ne is applicable. In the general
case, the same numbers for Mg XI become a lower limit for d and an upper limit for ne. A confidence region
in (d,ne) space requires new atomic kinetics calculations. Since in the general case d ≥ 2 × 1011 cm, we
conclude that the X-ray recombination line emission cannot possibly originate in the illuminated atmosphere
of HZ Her. This does not preclude an Fe Kα fluorescence line to be produced on HZ Her. If the recombination
lines did originate within the strong UV field of the companion, the Mg XI and Si XIII would have shown no
f line and a strong i line, like the rest of the He-like ions. The caveat for the d-limits is the possibility that
the UV from HZ Her are shielded on their way to the X-ray emission region, but not on their way to our
line of sight. The 864 A photons which photoexcite Si XIII are particularly vulnerable since they are above
the Lyman edge. Such a shielding material could be the accretion disk itself, in which case the origin of the
recombination lines in the disk atmosphere would be proven as well. Previously, no distinction could be made
between the case of disk coronal emission and emission from the illuminated companion (Jimenez-Garate et
al. 2002).
We base our arguments above in the fact that the primary source of UV emission in the binary system
is HZ Her and not the accretion disk. This is known from the orbital phase variability. The UV continuum
increases gradually from φ ∼ 0.2 until it peaks at orbital phase φ ∼ 0.5, and then decreases gradually until
the illuminated face of HZ Her comes out of view at φ ∼ 0.9 (Boroson et al. 2000). The intermediate mass
(∼ 2.3 M⊙) and size of HZ Her causes it to produce copious UV emission by reprocessing of the neutron star
X-rays. The accretion disk emission in the 1260 -- 1630 A band, which is identifiable from eclipse ingresses and
egresses at φ = ±0.1, represents a small fraction (∼ 1/10 to 1/20) of the peak HZ Her emission at that band
(Boroson et al. 2000). However, the question remains of how much UV continuum from the disk is visible
to the disk atmosphere. Although the observed UV from the disk is fainter than that from the companion,
that is partly because the disk is seen edge-on.
In sum, the weakening of any photoexcitation effects in the R ratios in Mg XI and Si XIII places a lower
limit on the distance from HZ Her to the X-ray emitting region, or a requirement on shielding from the UV.
These limits show that the X-ray lines are not produced on the illuminated surface of the companion, but
instead arise in a high density region that is likely associated with the accretion disk.
3.4. Search for orbital variability
The X-ray light curve is flat at mid-orbit. The light curves for the total and Fe Kα line fluxes are
shown in Figure 6. A Bayesian block analysis (Scargle 2003) on these light curves showed no deviations from
constant flux with 99.9% confidence. The coverage is limited to 1/3 of the orbit, so we cannot entirely rule
out orbital variability in the spectrum. After all, Her X-1 exhibits eclipses in the low state (Scott & Leahy
1999). Clearly, the variability of the Her X-1 emission is distinct from that observed in the ADC source
4U 1822-37, the latter likely arising on an illuminated disk bulge (Cottam et al. 2001). The UV continuum
emission from HZ Her shows a ∼ 25% flux variation from φ = 0.5 to φ = 0.65, peaking at mid-orbit, and
the lines of O V and N VI decrease in strength ∆φ = 0.2 away from eclipses (Boroson et al. 2000). With
better phase coverage than ours, XMM-Newton UV flux and Fe Kα flux measurements show a clear orbital
modulation (Zane et al. 2004, see their Fig. 9), which suggests that at least some of the Fe Kα flux originates
-- 8 --
on the face of the companion. Our data show that the Fe Kα flux was unusually flat during our observation.
4. XSTAR Photoionized Plasma Models and Emission Measure
The differential emission measure (DEM) maps the ionization distribution of the plasma. It serves as a
basis for testing astrophysical emission models and measuring elemental abundance ratios. A DEM analysis
is valid insofar as the spectrum is dominated by recombination emission, which occurs when the optical
depth is such that the plasma does not destroy or enhance the lines above their recombination values. The
luminosity of a line from level u to level l is given by
dLul = nenz,i+1EulηulαRRdV
(4)
where nz,i+1 is the number density of an ion with atomic number Z and charge +(i + 1), Eul is the transition
energy, dV is the differential volume, αRR = αRR(T ) is the total radiative recombination rate in units of
cm3 s−1, and ηul is the fraction of all the recombination photons that produce line emission through the
transition u → l. This can be related to the DEM = R d(EM )/d(log ξ) through the d(EM ) = n2
edV
available at each log ξ, since:
dLul =
nz,i+1
ne
EulηulαRRd(EM ) =
nz,i+1
ne
EulSuld(EM )
(5)
where the ionic density nz,i+1 = AZ nH fi+1 is often expressed in terms of the elemental abundance AZ , the
proton density nH , and the charge state fraction of the recombining ion fi+1. The specific line power is
defined as Sul = ηulαRR. Thus, the total line luminosity is
Lul =
AZ nH
ne
Eul Z fi+1Sul
d(EM )
d(log ξ)
d(log ξ)
(6)
This sort of DEM procedure was illustrated by Liedahl (1999) and applied by Sako, Liedahl, Kahn, & Paerels
(1999), and it is analogous to what has been used in collisionally ionized plasmas as d(EM )/dT (Kaastra et
al. 1996).
In the simplest analysis, we estimate the total emission measure EM = R n2
edV for each ion, assuming
each is formed at a single ξo and To. The EM for every prominent line in the spectrum is shown in Figure
7, calculated with the Solar abundances from Grevesse & Sauval (1998), with the updated C and O values
from Allende Prieto, Lambert, & Asplund (2002). We estimate the elemental abundance ratios by enforcing
the continuity of the EM as a function of ξo. The N VII and N VI line fluxes show a N over-abundance with
respect to C, O, Ne, Mg, Si, S, and Fe. The C VI line flux is obtained from the XMM-Newton RGS spectra
obtained by Jimenez-Garate et al. (2002). From Equation 6, it is useful to adopt ǫ ≡ fi+1Sul (in units of
photons cm3 s−1) as a measure of the line emissivity from now on. This ǫ is a function of ξ. We calculate
the fi+1 in a grid of models with 0.0 < log ξ < 5.0 and ∆ξ = 0.1, using XSTAR (Kallman & McCray 1982).
We use the main-on spectrum as the ionizing continuum. The Sul are calculated with HULLAC (Klapisch
et al. 1977). The log To and log ξo of formation for the Lyα or Heα lines are defined as the average
values of log T and log ξ weighted by ǫ. The XSTAR/HULLAC model temperatures of formation span
32, 500 K < To < 6.7 × 106 K, for N VI through Fe XXVI, respectively. We calculate the recombination rate
at a single ξo and assume fi+1 = 0.25. We use fi+1 = 0.25 because the log ξ-averaged, emissivity-weighted
value of fi+1 is approximately equal to that value. While this is just an approximation, it is more realistic
than simply setting fi+1 = 1. The EM analysis in Figure 7 has the advantages of being simple and of making
-- 9 --
the result independent of the plasma models; however, one sacrifices the accuracy of the EM distribution
with ξ.
In principle, a more accurate analysis than the above is obtained with DEM = d(EM)/d(log ξ), by
assuming that the DEM is constant in the ξ range at which each ion is formed. Again, we enforce continuity
of the DEM as a function of ξ. We use the aforementioned Solar abundances. Figure 8 shows that the N
abundance has a large excess, C is moderately depleted, and O is near the Solar value. A recalculation of
the DEM with the Wilms, Allen, & McCray (2000) Solar abundances establishes that the N excess is just
as large, while both the C and O depletion appear significant. This stems from the fact that the Wilms et
al. abundances for both C and O are larger than Allende Prieto et al.'s by ∼ 62% and ∼ 74%, respectively.
The abundance pattern, and especially the N abundance, is clearly a signature of CNO cycle products,
as shown previously by the XMM-Newton RGS spectral analysis, where the DEM was parameterized as a
power-law and the Wilms et al. abundances were used (Jimenez-Garate et al. 2002). The uncertainties in the
Solar abundances do not affect this conclusion, since no significant O depletion is needed. The CNO process
consists of the rapid conversion of C to N and the 103 times slower conversion of O to N (Clayton 1983,
and references therein). Evolutionary models of 2 -- 12 M⊙ stars produce an increase in the N abundance by
factors of a few, with C depletion in the tens of percent and O depletion of a few percent or less, after the
H-burning phase is over (Schaller, Schaerer, Meynet, & Maeder 1992). Clearly the most noticeable signature
will be the N abundance. It can be seen in the DEM that the abundance ratios among Ne, Mg, Si, S, and Fe
are close to the Solar values. Figure 8 shows that successive pairs of the H-like and He-like ions of a given
element indicate that the DEM decreases with ξ, from log ξ ∼ 1.5 to log ξ ∼ 2.4. There, the DEM turns over
and flattens out. The consequences of the DEM shape being rather flat are explored in the next section.
This DEM is sensitive to the modeled thermal stability of the plasma, as will be discussed in §7.
4.1. Volume, Scale Height, and Density Diagnostics
Consider the filling volume V of each line emission region in the optically thin limit. Note that the
DEM is rather flat, i.e. it varies by less than an order of magnitude (once abundances are accounted for).
Since ξ ∝ n−1
e , a flat DEM requires that V ∝ ξ2. Therefore, the region emitting Fe XXVI fills ∼ 102
times more volume than the Si XIII-emitting region, and ∼ 104 more volume than the N VI-emitting region.
The large disparity in the volume filled by different ions can be interpreted in two ways. The first way is
consistent with a model in which Fe XXVI is associated with hot corona with ∼ 104 times the scale height
of the compact, flattened disk atmosphere which produces N VI. A scale height ratio of the same order was
calculated from the hydrostatic model atmosphere (Jimenez-Garate, Raymond, & Liedahl 2002). A second
way is a scenario in which the coldest, densest part of the plasma is concentrated in small clumps, while the
hottest, more diffuse part of the plasma is distributed in a large volume.
Combining plasma diagnostics, photoionization balance, and emission measure, we constrain the size
and scale height of the line-emitting corona, set robust density constraints, and we estimate optical depths.
For this purpose, we select Fe XXVI, Mg XI, and O VIII. For each of these ions, we plot the density ne as a
function of distance r from the X-ray source. The emitting plasma is most likely to reside inside the Roche
lobe of the neutron star. The ne(r) limits drawn for Fe XXVI in Figure 9 are:
1. An upper limit on the density set by thermal and ionization balance. If the density is too high, the
plasma would be too cold to emit the line, because a given line can only be emitted at a fixed range of
ξ = L/(ner2). Thus, ne < L/(ξminr2), where ξmin is defined such that 95% of the line emissivity is at
-- 10 --
ξ > ξmin. We take L to be the main-on luminosity, which is reprocessed into line emission during the
low state.
2. A lower limit on the density given by the line flux through the emission measure EM and the maximum
volume Vmax, such that ne > pEM/Vmax = p3 EM/(4 πr3).
Figure 9 includes the density locus for a spherical shell of Thomson depth τ ∼ 0.05, which is consistent
with the above limits. Such is the depth required to Compton scatter the main-on continuum and produce
the continuum observed in the low state. This region of the corona extends to ∼ 1011 cm. The coronal
size derived from a shallow component of an eclipse ingress is r & 6 × 1010 cm: the ingress X-ray light
curves exhibit a steep and fast component and a slow and shallow component, also implying a core of X-ray
emission concentrated at r . 109 cm (Choi et al. 1994). High resolution spectroscopy is allowing us to probe
the extended region the corona. For the Fe XXVI corona with τ ∼ 0.05 to be enclosed by the Roche Lobe,
the scale height must be h/r ∼ 0.1, or more conservatively, 0.01 . h/r . 0.3.
We set stringent density constraints for the Mg XI emission region, taking advantage of the Mg XI R ratio
being at its critical value (see §3.3 and Fig. 10). We calculate the same ne limits as for Fe XXVI. In addition,
we calculate the critical ne (in the low UV flux limit), and the critical r found from UV photoexcitation (in
the low density limit). These are based on the R ratio, as discussed in §3.3. This allows us to compare the
derived density with the modeled density, for the case where the plasma is inside the Roche lobe. To obtain
the model density at the Her X-1 main-on luminosity, we interpolate between two disk atmosphere models
at Ledd and 0.1 Ledd, and we find that the density predicted by the atmosphere model (ne = 3 × 1013 cm−3 )
is consistent with that derived directly from the R ratio. This provides support to the disk atmosphere
model, because photoionization balance alone would allow one or two order of magnitude variations on ne,
depending on the optical depth of the plasma. The agreement in ne also implies that density rather than
photoexcitation effects are driving the R ratio in Mg XI.
The O VIII diagnostics in Figure 11 allow us to assign an optical depth to a given geometry. This figure
shows the two ne limits discussed previously. In addition, we plot the density of the plasma for two simple
geometries: a spherical shell and a flat pill-box or atmosphere. We show two solutions with optical depths
that are consistent with the density limits. The case of the spherical shell requires larger edge depths (of
τOVIII ∼ 3) than the case corresponding to a flat atmosphere, for which τOVIII ∼ 0.01 in the face-on direction,
as would be the case for a warped disk. The values that we will obtain from the optical depth diagnostics
in §5 will favor the scenario of a flat atmosphere.
4.2. Abundances from the (N VI + N VII)/(O VII + O VIII ) line ratio
Using the photoionized XSTAR and HULLAC models discussed in the beginning of §4, we quantify the
relationship between the P ≡ (N VI+N VII)/(O VII+O VIII) line photon flux ratio and the N/O elemental
abundance ratio. Using the P ratio as an abundance diagnostic has the virtue of being independent of
the EM and DEM analyses, relying instead on the XSTAR plasma models and the recombination rates
obtained from the HULLAC atomic code. The XSTAR models are calculated over a grid of log10 ξ values
that cover the range where the lines of interest are produced. We assume that the line emission is dominated
by the radiative recombination process, as is shown to be the case for Her X-1 in §5. Figure 12 shows
the calculated E ≡ [ǫ(N VI)+ǫ(N VII)] / [ǫ(O VII)+ǫ(O VIII)] line emissivity ratio as a function of ξ,
not accounting for elemental abundances (ǫ was defined earlier in §4). Therefore, we have P = E (N/O),
where (N/O) is the fractional abundance ratio between nitrogen and oxygen. The E and P ratios do not
-- 11 --
depend on density. At practically any ξ at which the lines of interest are produced, 0.7 < E < 2.4. The
latter values represent the worst case error for one assuming P ∼ (N/O). The measured line ratio ranges
from P = 0.87 ± 0.11 (Jimenez-Garate et al. 2002) to P = 1.54 ± 0.42 (this work). Combining these
numbers, we obtain (N/O)/(N/O)⊙ = 5.9 ± 0.6, using the Allende Prieto et al. (2002) and Grevesse &
Sauval (1998) Solar abundances as reference, and (N/O)/(N/O)⊙ = 9.2 ± 1.0 using the Wilms et al. (2000)
Solar abundances. The error bars on these abundance ratios only include statistical errors. Compare this to
the 4 . (N/O)/(N/O)⊙ . 8 that can be estimated by enforcing continuity on the DEM in Figure 8.
5. Quantifying optical depth
X-ray spectroscopic analysis allows us to measure or set limits on the mean number of scatters of
line photons, or to the line optical depth. The photoelectric optical depth can also be constrained by the
spectrum, despite the absence of absorption edges. These measurements can have important implications for
the geometry of the plasma. They also allow us to validate the previously determined elemental abundance
ratios.
5.1. The mean number of scatterings for Lyα and Lyβ photons
The mean number of scatterings is constrained by the observed Lyα/Lyβ ratios. The Lyα/Lyβ ratio is
a temperature diagnostic in collisional plasmas. In contrast, this ratio depends very weakly on temperature
for a recombining plasma, as we have verified from the HULLAC recombination rates. A Lyβ photon from
a H-like ion can get resonantly absorbed by another ion of the same species (see Fig. 13). There is a
probability γ = 0.12 that the absorbed Lyβ photon is converted to photons with lower energy (including
an Hα). Otherwise, the absorbed Lyβ is re-emitted. The probability of a Lyβ photon to survive N line
scatterings is:
P = (1 − γ)N = 0.88N ,
(7)
and therefore the Lyβ photons cannot scatter more than a few times before getting destroyed. As the
number of scatterings increases, the Lyα/Lyβ intensity ratio increases with it, because Lyα photons do not
get destroyed in line scatterings (γ = 0). None of the observed Lyα/Lyβ line ratios shown in Table 4 show
evidence for multiple line scatterings, except for O VIII. Since we deal with a large ensemble of photons, the
number related to the line ratio is the mean number of scatterings hN i, instead of N above.
The observed Lyα/Lyβ ratio allows us to constrain the mean number of scatterings of Lyβ photons
directly, and of Lyα photons by inference. If γ > 0, the mean number of scatterings hN i for a slab geometry
satisfies (Hummer & Kunasz 1980) :
hN i =
1 − γ
γ
EL
EG
(8)
where EL is the line energy lost to resonant absorption and EG is the total line energy generated in the gas
column. There are two relevant cases for the upper limit of hN i. In the case γ = 0, photons do not get
destroyed (as for Lyα), then hN i does not have an upper bound, and Equation 8 does not apply. In the case
γ > 0 (as for Lyβ), hN i is bounded above (by hN i < 7.33). as EL/EG → 1 in Equation 8.
We first use Equation 8 for O VIII. We detected 7 counts of the O VIII Lyβ line in the 15.995 A
to 16.015 A range. Given the continuum and background level of 3.0 counts, this represents a detection
significance of 95%, as derived from a Poisson distribution. The caveat is that this is only valid if the
-- 12 --
background is uniform in λ, and it may be subject to systematics. Since this is not a line search (i.e. the
line energy is fixed), this may be a positive detection, but it needs confirmation. Taking Lyβ at face value,
we derive EL/EG < 0.9 for O VIII Lyβ with 90% confidence, and obtain hN i < 6.6. The detected line is
centroided at 16.005 A (O VIII Lyβ), and it is one MEG-resolution FWHM away from Fe XVIII at 16.023 A.
To infer the mean number of scatterings of Lyα1 from those of Lyβ, we need to relate hN i to τ , the
mean line optical depth. Hummer & Kunasz (1980) performed such a calculation for a slab geometry with
a uniform source of line photons, γ = 0, and Voigt parameters of 4.7 × 10−2 to 4.7 × 10−4. They obtained
hN i/τ ∼ 0.55 -- 0.6 for τ = 1, of hN i/τ ∼ 0.60 -- 0.65 for τ = 10, and hN i/τ ∼ 0.73 -- 0.85 for τ = 100. The
Voigt parameter is the ratio between the natural broadening and the velocity broadening of the line. Define
cτ = hN i/τ . To generalize this result for γ > 0, note from Equation 2.15 in Hummer & Kunasz (1980) that
(1 − γ) can be factored out of the expression for hN i, such that hN i = cτ τ (1 − γ) is applicable in that case,
with cτ and τ as before. Note cτ = cτ (τ ) is a function of τ .
The mean line optical depth τ is proportional to the oscillator strength τ ∝ fosc. Therefore, we obtain
hN i ∝ cτ fosc(1 − γ), from which we estimate hN i ∼ 30 for O VIII Lyα1 photons. A similar number can be
obtained through the more simplistic Equation 7. We ignored the continuum optical depth. The effect of a
nonzero continuum depth is to decrease hN i below the derived values. Up to this point, this result depends
on the detection of the Lyβ line, but there is another way to deduce hN i for O VIII Lyα1 photons.
We set an upper limit on hN i for O VIII Lyα1 without relying on O VIII Lyβ, by using the Ne X lines.
This is valid since both ions are produced at a largely overlapping ionization parameter range. We find that
hN i . 2 for Ne X Lyβ photons (at 90% confidence). Using the cτ once again, this implies hN i . 9 for
Ne X Lyα1. The EM derived from Ne X and O VIII are the within 12% of each other, consistent with being
equal given the statistical uncertainties. Scaling with the Wilms, Allen, & McCray (2000) or Allende Prieto,
Lambert, & Asplund (2002) Solar abundances, we deduce hN i . 69 for O VIII Lyα1. Scaling with CNO
abundances, hN i is smaller than or equal to the latter limit. Both are consistent with the hN i ∼ 30 value
for O VIII Lyα1 which was derived above.
In contrast, the Lyα/Lyβ ratios of Mg XII and Si XIV on Table 4 may show evidence for enhancement of
Lyβ due to photoexcitation effects at moderate column densities. Photoexcitation of ions by X-ray continuum
photons, followed by radiative decay, can dominate the emission mechanism. This process, also referred to as
resonant scattering, can produce a characteristic Lyα/Lyβ ratio (Kinkhabwala et al. 2002). At low column
densities, both lines are unsaturated, resonant scattering dominates over recombination emission, and the
Lyα/Lyβ ratio simply depends on the oscillator strengths. At moderate column density, the Lyα/Lyβ ratio
is reduced below the low column value, when the Lyα line reaches saturation but the Lyβ line does not. The
evidence for photoexcitation of Mg XII and Si XIV at the moderate column density regime is a ∼ 2σ result,
so we do not discuss this scenario further. At high column densities, recombination emission dominates over
resonant scattering, because resonant scattering has completely saturated while the emission measure keeps
growing with the column, and the Lyα/Lyβ ratio is determined by the ratio of the recombination efficiency
of the lines, which happens to be similar to the oscillator strength ratio. In this section we have expanded
upon the Kinkhabwala et al. (2002) picture, showing that at high line optical depths, the Lyα/Lyβ ratio
can increase above its pure recombination value due to the destruction of the Lyβ line.
-- 13 --
5.2. The optical depth of the r line in He-like ions
The value of the G = (i + f )/r ratio depends on whether the plasma is collisionally ionized or pho-
toionized, but it can also can be used to quantify the mean optical depth of the r line. The G ratio depends
on optical depth for three reasons: 1) the r line has much larger oscillator strength than either the i or f
lines; 2) the recombination rates of the i, f lines are larger than those of the r line, because the i, f lines
have larger statistical weight than the r line; and 3) the G ratio depends very weakly on temperature in a
photoionized plasma. This diagnostic has been quantified by Kinkhabwala et al. (2002) and Wojdowski et
al. (2003). In the limit of small line optical depth, we have G ∼ 0.3, because the r line is enhanced relative
to the f and i lines by continuum X-rays that photo-excite the ground state (Wojdowski et al. 2003). In
the small line depth regime, the emission is dominated by resonant scattering over recombination emission.
In this regime, the line spectrum is dependent on the spectral energy distribution. In the limit of large line
optical depth, G reaches an asymptotic value of 4.5 (Porquet & Dubau 2000), which depends very weakly
on the the T values derived from photoionization equilibrium, and recombination emission dominates over
resonant scattering.
The observed G ratios shown in Table 3 are consistent with the limit of large line optical depth. By using
the lower bounds on G, we can set lower limits on the mean depth of the r lines as defined by Wojdowski
et al. (2003), of τr & 100 for G > 4 (the maximum value applicable to Ne IX and O VII), τr & 20 for G > 3
(applicable to Mg XI), and τr & 10 for G > 2 (applicable to Si XIII), assuming a continuum power-law index
of α = 1.0. The observed main-on index is α = 0.85, but G depends weakly on α.
5.3. An X-ray atmosphere illuminated at a small grazing angle has anisotropic line depths
If the plasma geometry is anisotropic, the line optical depth will depend on direction. A first special
direction vector corresponds to the path of the X-ray continuum photons from the X-ray pulsar, since this
radiation is photoionizing the plasma. The second special direction is that which minimizes the optical
depth, since photons will tend to escape from the plasma that way. Based on this, we define two relevant
depths: an "illumination" optical depth τi, and an "escape" optical depth τe, which are well-defined and not
equal if the plasma distribution is illuminated along a direction where the optical depth is not minimized.
The illumination depth τi determines how much of the continuum radiation is absorbed, and therefore how
much is reprocessed into resonant and recombination emission lines. On the other hand, once a line photon
is produced, it is more likely to escape in the direction where the optical depth is minimized than through
other directions, and this corresponds to the escape depth τe.
The Heα G ratio is sensitive to τi, because G depends on the ratio between the resonant scattering and
recombination line fluxes, as mentioned in the previous section. Further scattering of the r line after it is
produced does not destroy it unless the continuum depth is significant. On the other hand, the Lyα/Lyβ
ratio is sensitive to both τi and the escape depth τe. Again, resonant line absorption of Lyα or Lyβ produces
a dependency on τi, because both photons resonant scatter strongly.
In addition, the Lyβ photon can
be destroyed and converted to Lyα if it scatters too many times inside the plasma, and this is why the
dependence on the escape depth τe arises.
The flat, geometrically thin shape of the model disk atmosphere is anisotropic, and the atmosphere is
illuminated by the neutron star at a very shallow grazing angle that was determined from self-consistent
calculations (Jimenez-Garate, Raymond, & Liedahl 2002). This geometry is shown schematically in Figure
14.
In the atmosphere model, the calculated grazing angle is such that illumination depth is 25 times
-- 14 --
larger than the escape depth at r = 1011 cm, and therefore the observed Heα r line depth is expected to
be much larger than the Lyβ depth. Sample vertical column densities in the disk atmosphere model are
NOVIII = 5 × 1017 cm−2 , NOVII = 2 × 1017 cm−2 , and NSiXIII = 1.6 × 1016 cm−2 . These columns imply
escape line depths of τe ∼ 46 (150 km s−1/v) for O VIII Lyα1, τe ∼ 53 (150 km s−1/v) for O VII r, and of
τe ∼ 1.4 (150 km s−1/v) for Si XIII r, where v a fiducial velocity width that includes thermal and turbulent
components. Setting v to the thermal value at kT = 25 eV for O VIII yields τe = 178 for Lyα1, and setting
kT = 3.5 eV for O VII yields τe = 544 for He r. Note that τi ∼ 25τe for the disk model geometry, i.e. for
Si XIII we get τi = 36 (150 km s−1/v). These are the expected line depths from the disk atmosphere model,
modulo the assumptions made about the gas kinematics.
Compare the above values with the optical depth measurements from Lyα/Lyβ that yielded hN i ∼ 30
for O VIII Lyα1, which is equivalent to τ ∼ 46 given the calculated hN i/τ ratios and Voigt parameters from
§5.1. This optical depth matches the model at the chosen 150 km s−1 velocity width. If the O VIII Lyβ line
detection is not real, from Ne X it follows that for O VIII Lyα1 still τ . 100. This would simply require a
smaller velocity width than above, so there is parameter space available to achieve consistency of the data
with the disk atmosphere model in this case. For the case of O VII, an upper limit τr > 100 was obtained
from the G ratio, which is easily accommodated by τi ∼ 1300 (150 km s−1/v) for the illumination depth.
This large τi leaves plenty of parameter space to be consistent with the other lower limits on τr set for
Ne IX, Mg XI, and Si XIII. For example, Si XIII was measured to have τr & 10, which is consistent with the
atmosphere model τi = 36. For reference, the model escape depth for the continuum is τe(OVIII) = 0.05, an
illumination depth τi(OVIII) = 1.2.
Our results above depend on the estimated kinematics, and therefore are only approximate consistency
checks. Since the velocity field in the disk is produced by both Keplerian and turbulent motions, τe will in
fact be highly anisotropic, and it will depend on the photon momentum vector relative to the disk velocity
field. To self-consistently derive a turbulent velocity from the Lyα/Lyβ ratio, one requires a calculation of
the line transfer which accounts for this velocity field.
5.4. RRC/RR ratios and electron temperature
The electron temperature measured from the Ne IX RRC is kT = 7 ± 3 eV, or T = 81000 ± 35000 K.
This was measured by fitting the semi-Maxwellian profile of the Ne IX RRC at λ = 10.37 A, the brightest
RRC in the spectrum, detected at the 5σ level (see Fig. 3). The XSTAR plasma code and the HULLAC
recombination rates yield emissivity-weighted average temperatures kTo which are shown on Tables 4 and 5.
The To for Ne IX is in good agreement with the measured T . The significance of detection for other RRCs
obtained with the Cash statistic ranges from the 2σ to the 5σ level (see Table 1). To fit the RRCs, we rebin
the spectra to bin sizes ∆λ of 0.05 A to 0.09 A (coarser than the bin size used for fitting lines). While most
RRCs do not have sufficient statistics for an accurate T measurement, T is left as a free parameter and the
best-fit values are consistent with To for most cases.
In Table 5, we show the observed RRC to RR (radiative recombination lines) flux ratios, compared to
the HULLAC calculations, in the recombination-dominated case, for both a single temperature model and
for the disk atmosphere model. The RRC/RR ratios for O VII, O VIII, and Ne X are significantly weaker
than expected for the single-T model, but agreement is generally better for the disk atmosphere model, which
accounts for a broad distribution of temperatures 2 < kT < 860 eV. It is therefore likely that the RRCs are
composed of multiple temperatures, with the low-T component being observed more readily than the high-T
-- 15 --
component. The high-T RRC component, which usually dominates the recombination flux, is too broad to
be separable from the continuum. In this favored scenario, the weak RRCs are simply an indication of the
multi-T nature of the plasma.
5.5. Verification of CNO abundances: the line optical depths are not large enough to
convert O VIII Lyα to N VII Lyα
We test for the presence of the X-ray Bowen fluorescence mechanism by which a O VIII Lyα line becomes
a N VII Lyα line, as calculated by Sako (2003). We determine whether this X-ray Bowen fluorescence effect
can account self-consistently for the apparent excess intensity observed in N VII Lyα, N VI i and r, and the
N VII RRC with respect to other lines, such as the O VIII Lyα and O VII i and r. We find that:
1. The X-ray Bowen fluorescence mechanism is unlikely to enhance N VI Heα emission in the same
proportion as N VII Lyα. This would require line opacities to overlap in energy space under nearly
identical conditions as those found for N VII Lyζ and O VIII Lyα2, whose energies are 1 Doppler
width away at T = 50 eV. We checked for coincidences between the energies of the brightest lines of
abundant elements in the 22.46 to 26 A band and the resonance line energies of N VI. We did not
find line overlaps closer than 46 Doppler widths at T = 50 eV, and the number of Doppler widths is
likely to be higher since models show a peak N VI Heα flux at T = 2.8 eV. As shown in the HETGS
spectrum in Fig. 4, as well as from the XMM-Newton RGS spectrum (Jimenez-Garate et al. 2002),
both N VI Heα and N VII Lyα indicate a N over-abundance.
2. The measured mean number of scatterings of O VIII Lyα and Lyβ photons is two orders of magnitude
lower than required for the X-ray Bowen fluorescence effect to take place. The measured line ratio
between O VIII Lyα and Lyβ depends on hN i, due to the destruction of Lyβ or its conversion to Lyα.
We derive hN i ∼ 30 for O VIII Lyα1. This was verified via the same ratio in Ne X, which scales to
hN i . 69 for O VIII Lyα1 with Solar or CNO abundances (see §5.1). Compare this to the hN i ∼ 5×103
scatterings needed for the N VII enhancement (Sako 2003).
3. The detection of the N VII RRC with XMM-Newton and Chandra provides additional evidence for
a N/O overabundance. At the escape line depths implied by the measured Lyα/Lyβ ratios, the
RRC fluxes are unaffected by optical depth effects. The N VII RRC detected with XMM-Newton RGS
(Jimenez-Garate et al. 2002) at 10σ has flux (6.5±2.8)×10−5 ph cm−2 s−1 , with 90% confidence errors,
and HETGS detected it at 3.5σ with flux (18 ± 5) × 10−5 ph cm−2 s−1 , with 68% confidence errors
(Table 1). These fluxes are larger than the . 1 × 10−5 ph cm−2 s−1 expected from Solar abundances
and the fluxes of O VII RRC and O VIII RRC.
4. An analysis of the UV line emission from Her X-1 using the same sort of illuminated disk model
concluded that N was over-abundant in HZ Her, with a N/C abundance ratio of ∼ 2 (Raymond 1993),
which implies a N overabundance relative to Solar of ∼ 6. This shows that both UV and X-ray disk
emission models yield consistent results.
5. The velocity shear in the disk atmosphere and corona increases the line escape probability, which
would decrease hN i below the static values calculated by Sako (2003) for a given column density. This
suppresses X-ray Bowen Fluorescence.
-- 16 --
We conclude that the large nitrogen over-abundance interpretation of the Her X-1 spectrum by Jimenez-
Garate et al. (2002) is valid and cannot be self-consistently explained by the X-ray Bowen fluorescence effect.
The large nitrogen over-abundance implies that significant CNO processing occurred in a massive star in the
system. New transfer calculations are required to improve the optical depth and dynamical constraints on
the accretion flow. Such calculations need to include thermal and ionization balance, stratification, and gas
orbital kinematics in order to calculate line transfer more accurately than existing models.
6. Comparison with an accretion disk atmosphere and corona model
In Figs. 2 -- 4 we over-plot the synthetic spectrum of an illuminated accretion disk atmosphere and corona
model from Jimenez-Garate, Raymond, & Liedahl (2002). We compare the data to this disk atmosphere
model. The model assumes that the neutron star continuum energizes the vertically stratified structure of a
hydrostatic accretion disk in thermal and ionization balance. We used two adjustable parameters to fit the
spectrum: a normalization factor of N = 0.11, and the disk radius range 1010.9 < r < 1011 cm (the outer
annulus in the model). Both parameters affect the normalization and do not influence the spectral shape.
We did not include any Doppler velocity broadening effects. The ionizing continuum in the model consists
of a bremsstrahlung with T = 8 keV and a luminosity Lx = 1038.3 (D/6.6 kpc)2 N erg s−1 . Given the
approximate linear dependence of line flux on continuum flux (Jimenez-Garate, Raymond, & Liedahl 2002),
the effective ionizing luminosity in the model fit was Lx = 2.3 × 1037 erg s−1 . This is nearly the unabsorbed
luminosity during main-on state of Her X-1 in the 0.1 -- 200 keV band, L = 3.8 × 1037 ergs−1 (dal Fiume et
al. 1998). The model normalization can be interpreted as simply a geometrical parameter which indicates
that the observed emission is produced by an outer section in the accretion disk atmosphere and corona with
area A ∼ 3 × 1021 cm2. That A is a small fraction of the total disk area is in agreement with observing an
outer disk rim, as suggested by the small velocity broadening of the emission lines (§3.2). The neutron star
continuum observed during the main-on powers the recombination emission observed during the low state.
The precession of the accretion disk allows us to isolate the reprocessed emission from the direct emission
from the neutron star.
The relative strengths of the spectral lines, which are fixed in the model, are a direct measure of the
structure of the accretion disk atmosphere and corona. In general, they constrain the temperature, ionization,
and density distributions.
In the context of our model, the relative line strengths are indicators of the
vertical ionization structure of the disk atmosphere and corona, since the ionization level and temperature
are stratified with vertical depth.
In Table 1, we compare the relative strengths of lines with different
ionization parameters (except for Fe Kα fluorescence at 1.94 A, which is fit separately, because it is not yet
calculated in the disk model). We note that the relative strengths of the Lyα lines of S XVI, Si XIV, Mg XII,
Ne X, and O VIII lines were fit with ∼ 50% or better accuracy, with virtually zero degrees of freedom, because
N and the radius range, the two free parameters in the model, only affect the overall flux normalization.
Since O VIII and S XVI are produced at ionization parameters that are one order of magnitude apart, it is
a notable success for the model to be able to reproduce their flux ratios. In the disk model context, this
indicates that the emission measure distribution of the upper half of the atmosphere agrees well with the data.
The incidence angle on the disk atmosphere is self-consistently determined by the model (Jimenez-Garate,
Raymond, & Liedahl 2002).
The Fe XXVI Lyα line probes the hottest and most extended region of the disk corona observed with
HETGS (see §4.1 and Fig. 2). The XSTAR photoionization equilibrium plasma model (Kallman & McCray
1982), coupled with the recombination rates obtained with the HULLAC atomic code (Klapisch et al. 1977),
-- 17 --
show that the electron temperature of the plasma producing the Fe XXVI line is Te ∼ 6.7 × 106 K, with
ionization parameter log ξ ∼ 3.8. Similar parameters are obtained with the disk atmosphere model. The
observed Fe XXVI Lyα line flux is 3.9 times larger than the disk model's. Judging from the disk model, the
Fe XXVI Hα and Hβ recombination lines at 9.68 A and 9.53 A (see Fig. 3 and Table 1) are 10 times weaker
than expected. The flux ratio of Hα and Lyα lines is not expected to depend strongly on temperature in
the pure photoionization case. The Fe XXVI Lyα line was also detected with XMM-Newton (Zane et al.
2004). An emission line of Fe XXV at 1.85 A appears in the model, as well as a Fe XXIV complex with lines
at 11.18 A, 11.03 A, and 10.62 A (these lines appear in Fig. 3). None of these relatively weak lines are
detected. In short, the upper coronal structure appears to be more ionized and to have a larger emission
measure in Her X-1 than in the disk atmosphere and corona model.
We have shown that several key characteristics of the X-ray recombination spectrum can be explained
by an outer section of the accretion disk which is being illuminated by the neutron star. This disk at-
mosphere model describes much of the Her X-1 X-ray line spectrum from first principles. However, there
are discrepancies between the model and the data. One notable discrepancy is N VII and N VI, which are
under-predicted by a factor of ∼ 10. This is consistent with the §4.2 results. Also, the He-like ion line fluxes
are systematically underpredicted by the model, as we discuss below.
7. Thermal stability of the photoionized plasma and the H-like/He-like ion line ratios
A thermal instability has been predicted to be present in plasmas which are photoionized by an X-ray
continuum. The "type I" instability is produced by an imbalance between bremsstrahlung cooling and Comp-
ton heating, just below the Compton temperature (Krolik, McKee, & Tarter 1981). The "type II" instability
is produced by the imbalance between recombination cooling and photoionization heating (Kallman 1984).
The photoionized disk atmosphere models predict a type II instability regime at 6 × 104 < T < 5 × 105 K
(Jimenez-Garate, Raymond, & Liedahl 2002). X-ray recombination emission is suppressed in the latter
regime. The recombination line fluxes are sensitive to the ion fraction fi+1(ξ). The instability reduces the
interval [log ξo − ∆(log ξ),
log ξo + ∆(log ξ)] at which fi+1 & 0.1. For this reason, the disk atmosphere model
predicts very weak Mg XI line fluxes relative to Mg XII, contrary to observations (as shown in Fig. 3). The
DEM analysis indicates that the Mg XI flux is above the expectation (Fig. 8). The ions with ξo near the
instability regime will have their line power reduced and their derived DEM increased.
The He-like ion lines of Si XIII, Mg XI, Ne IX, O VII, and N VI are underpredicted by the atmosphere
model. Also, the DEMs derived from the same He-like ions are systematically larger than the DEMs from
the H-like ions. This may imply that the plasma is more thermally stable than the models predict, or it
may require a larger EM of T ∼ 105 K plasma. The He-like ion lines are especially sensitive to the thermal
instability because of a key physical difference: the H-like ions form by the recombination of fully stripped
atoms, so their line emissivities are less sensitive to ξ than those of He-like ions. The fi+1(ξ) for the He-like
lines are single-peaked functions, while the fi+1(ξ) for the H-like lines are asymptotic curves with fi+1 → 1
as ξ → ∞, and fi+1 → 0 as ξ → 0. This causes the line emissivities for the He-like ions to have narrow
peaks and small ∆(log ξ), while in contrast, the line emissivities of H-like ions have broad asymmetric peaks
with large ∆(log ξ), with tails that extend to large ξ. This is why, in spite of a possible overlap of the ξ of
formation between the He-like and H-like ions, the H-like ion line fluxes will still dominate over the He-like
ion lines at high ξ. Therefore we expect the He-like ion lines to be more sensitive indicators of the ionization
distribution and the instability regime than the H-like ion lines. The changing trends of the EM in Figure 7
compared to the DEM in Figure 8, suggest that the thermal instability regime is larger in the models than
-- 18 --
in the plasma, but this is difficult to ascertain without a systematic study of the plasma models.
The ne measurement from Mg XI Heα (§4.1) also tests stability of the photoionized plasma, since the
disk atmosphere model predicts a regime of forbidden density. The ne regions allowed by both the spectral
diagnostics and disk atmosphere models are shown in Figure 10. The density of formation for Mg XI is
ne = 3 × 1013 cm−3 according to the disk atmosphere model. The forbidden densities are just below that,
at 2 × 1012 < ne < 3 × 1013 cm−3 . The measured ne = (2 ± 1) × 1013 cm−3 is located in the high density
end of the stable regime. The thermal instability calculation in the disk model is in agreement with the
observations in this case.
The X-ray spectroscopic data is now sufficiently detailed that it is possible to start testing the regimes of
thermal instability in the photoionized plasma. On the one hand, a DEM analysis based on XSTAR plasma
models and the HULLAC recombination rates suggests that the the plasma is more thermally stable than
was originally predicted. On the other hand, the density constraints from the R ratio of Si XIII indicate that
the density is just at the edge of the stability region, as derived from the disk atmosphere model. Further,
the same disk atmosphere model predicts the Mg XI Heα flux to be suppressed due to the instability, but
no such suppression is observed.
8. X-ray Spectral Comparison with other LMXBs
We compare the spectral properties of the Her X-1 low state with two low-mass X-ray binaries: 4U 1822-
37 and 2S 0921-63. Both 4U 1822-37 and 2S 0921-63 are classified as ADC sources. These LMXBs exhibit
X-ray emission lines with unresolved velocity broadening. The limits on the velocity widths (FWHM) of
the recombination lines measured with Chandra HETGS are ∆v < 1500 km s−1 for 2S 0921-63 (Kallman,
Angelini, Boroson, & Cottam 2003), ∆v < 300 km s−1 for 4U 1822-37 (Cottam et al. 2001), and ∆v <
260 km s−1 for Her X-1 (from O VII in Table 1). Her X-1 and 4U 1822-37 exhibit RRCs that indicate
that photoionization is the dominant ionizing mechanism in the plasma. In 2S 0921-63, the bright O VII
intercombination line (giving the value of G ∼ 4 for that diagnostic line ratio) is indicative of a photoionized
plasma (Kallman, Angelini, Boroson, & Cottam 2003). The non-detection of RRCs in 2S 0921-63 is likely
due to the low signal-to-noise ratio of the emission features. The weak or undetected Fe L-shell lines in
these LMXB spectra are yet another signature of photoionization. It is therefore clear that the lines in these
LMXBs are produced by recombination following photoionization.
When looked at more carefully, a comparison of the spectra of 4U 1822-37, 2S 0921-63, and Her X-1
reveals a nuanced picture. First, the continuum level varies significantly from source to source: 4U 1822-37
has ∼ 10 times larger continuum flux in the 0.5 -- 7 keV band than 2S 0921-63 or Her X-1. The continuum
in 4U 1822-37 is harder than in the other two sources. Second, the soft X-ray line spectra in 4U 1822-37
and Her X-1 appear to be quite similar, i.e. the fluxes of the Si XIV, Mg XII, Mg XI, Ne X, Ne IX, O VIII,
and O VII lines are roughly within tens of percent of each other, albeit the exact line ratios and density
diagnostics differ. On the other hand, 2S 0921-63 shows weaker line fluxes than 4U 1822-37 and Her X-1,
perhaps because the eclipse is blocking the source of the line emission. Third, the line ratios between Fe
emission lines from both high and low ionization species are different. Her X-1 exhibits a much more intense
Fe Kα line from Fe I -- Fe XIII at 1.94 A than 2S 0921-63, but a nearly equal flux of the Fe XXVI line at 1.78 A.
2S 0921-63 exhibits both Fe XXV and Fe XXVI lines with nearly equal fluxes, and a Fe Kα fluorescence line
with 1/4 of the flux of the Fe XXV/Fe XXVI lines (Kallman, Angelini, Boroson, & Cottam 2003).
It is
conceivable that the fluorescing material was occulted by the eclipse of 2S 0921-63, while the hotter, more
-- 19 --
extended emission was not. 4U 1822-37 also showed Fe Kα and Fe XXVI lines, with Fe Kα being brighter
(Cottam et al. 2001), but it did not show the large flux contrast between those lines observed in Her X-1.
Fourth, the N VII Lyα and N VI Heα lines are most intense in Her X-1 (see Fig. 4). So far, other LMXBs
and ADC have not shown such strong N lines, albeit the measurement is rendered difficult for most cases
due to the large absorbing hydrogen columns NH that are common in these systems. In §4 and §5.5, we
quantified this abundance anomaly, which was also observed with the XMM-Newton RGS (Jimenez-Garate
et al. 2002).
We now compare X-ray flux variability with orbital phase. While our observations of Her X-1 are
restricted to mid-orbit at 0.33 < φ < 0.67, the 2S 0921-63 observations were restricted to mid-eclipse, and
the 4U 1822-37 observations spanned full orbits. The emission during the 2S 0921-63 eclipse originates in
either the outer disk rim or in an extended circumsource region (such as a corona) with an upper size limit
of 7 × 1010 cm (Kallman, Angelini, Boroson, & Cottam 2003). The emission line region in 4U 1822-37
was identified to be the illuminated disk bulge, because the line fluxes were observed to peak at φ ∼ 0.25
(Cottam et al. 2001). Flux variations were not observed in Her X-1, but with the limited phase coverage we
can ascertain that it does not show the same sort of variation than 4U 1822-37. Judging from the spectral
variations as a function of orbital phase, the azimuthal location of the line emission regions in Her X-1,
2S 0921-63, and 4U 1822-37 appears to be distinct amongst these sources.
Her X-1, 4U 1822-37 and 2S 0921-63 are sources that are observed at high-inclination angles (85◦.
i < 90◦), which causes the accretion disk to block the direct X-ray emission from the accreting object.
ADCs have a higher optical-to-X-ray flux ratio than other LMXBs, which is a well-known indicator that the
accreting object is not observed directly (White, Nagase, & Parmar 1995, and references therein). This, in
turn, causes the EWs of the X-ray emission lines in ADCs to be much larger than in other LMXBs, because
only ∼ 10−2 -- 10−3 of the available continuum energy is reprocessed into recombination lines. This scenario
is consistent with the model presented in §6, where only a fraction of the accretion disk area can account
for the line emission. Incidentally, the only other ADC for which a high-resolution spectrum was obtained
is AC 211 (in the M15 cluster), but it did not show emission lines (White & Angelini 2001), probably due
to contamination from a bright and nearby LMXB burster.
In sum, high-inclination LMXBs exhibit recombination-dominated line spectra and unresolved line
broadening. The illuminated disk bulge seen in 4U 1822-37, however, is not observed in Her X-1 with
the current data, and only the extended emission during eclipse was observed in 2S 0921-63. A definitive
comparison of the accretion flow geometry of these sources will require new spectral data with good orbital
phase coverage for Her X-1 and 2S 0921-63.
9. Conclusions
1. We detect more than two dozen spectral features in the X-ray spectrum of the Hercules X-1 low state
with the Chandra HETGS. Most of the observed features correspond to the recombination signatures
expected from a centrally illuminated accretion disk atmosphere and corona located in the outer ac-
cretion disk, given the following evidence:
• Using photoexcitation diagnostics with the He-like ion lines, coupled with the observed ultraviolet
continuum fluxes by Boroson et al. (2001), we ruled out the possibility that the recombination
emission lines originate on the illuminated companion star, HZ Her.
• The spectral model for a centrally illuminated accretion disk atmosphere and corona fit most of
-- 20 --
the observed fluxes and ratios of the X-ray recombination lines. The model requires the low state
emission to be energized by the continuum observed in the high state. The caveat is that the
He-like lines are underpredicted by the model, and the observed line broadening is smaller than
expected for a plasma corotating with the accretion disk. The small Doppler widths may be due
to vignetting effects that allow only the far edge of the disk to be visible, as would occur for an
edge-on disk with a flared or warped geometry. In support of this, the model fits indicate that
only the outer rim of the disk (8 × 1010 < r < 1011 cm) is observed, in contrast with LMXBs
such as EXO 0748-67, where much of the inner disk emits observable X-ray lines (Jimenez-Garate,
Schulz, & Marshall 2003).
• The measured R line ratio of Mg XI allows us to precisely determine a density of ne = (2 ± 1) ×
1013 cm−3 , which agrees with the disk atmosphere model predictions (ne = 3 × 1013 cm−3 ). This
corresponds to the low UV field limit, which is the most likely scenario for Mg XI. Using UV
observations (Boroson et al. 2001), we deduce that the disk may be far enough from HZ Her for
the Mg XI triplet to be unaffected by photoexcitation. The presence of UV emission and high-ne
necessitates the re-calculation of R ratios in (ne, FUV) space.
• The X-ray spectrum probes radically different length scales in the atmosphere and corona. For
example, an emission measure analysis shows that the volume where Fe XXVI is emitted must
be ∼ 104 times larger than that for N VI. Compare this to the cool model atmosphere emitting
N VII at T ∼ 32500 K, with 107 cm thickness, and 3 × 1013 cm−3 density, which coexists with a
1010 cm thick hot corona emitting Fe XXVI, with 1011 -- 1012 cm−3 density and T ∼ 106 K (with
intermediate layers of ionization in between).
• The RRC/RR ratios and the measured T = 81000 ± 35000 K from Ne IX RRC are generally in
agreement with the disk atmosphere model. Single-T models do not reproduce well the observed
RRC/RR flux ratios of H-like ions, while the disk model produces improved agreement. This
indicates the presence of multiple temperature components, since any high-T component of the
RRCs is lost in the continuum.
• The line depths derived from the spectrum via the Lyα/Lyβ and G ratios are consistent with the
ionic column densities and flattened geometry of the model disk atmosphere, where the first ratio
is most sensitive to escape depth (vertical depth) and the second is determined by the grazing
incidence depth. To obtain consistency, we use a turbulent velocity width of ∆v ∼ 150 km s−1 .
We rely on the relationship between τ and the mean number of scatters hN i calculated for a range
of Voigt parameters by Hummer & Kunasz (1980).
2. We find a prominent Fe Kα1, α2, Kβ complex indicative of a neutral plasma. We cannot ascertain
whether the Fe K fluorescence is produced in the disk or on HZ Her, or both. The Fe Kα line flux
from HETGS does not vary with orbital phase at 0.3 . φ . 0.7, but Zane et al. (2004) did observe a
clear modulation with φ with XMM-Newton.
3. The lines in the X-ray spectrum have unresolved velocity broadening (∆v < 260 km s−1 ), similar to
what has been observed for the lines in ADC sources. This is evidence that the accretion disk is edge-on
with respect to our line of sight. Previously observed spectral changes with the XMM-Newton RGS
showed a transition between broad lines in the main-on state and narrow lines in the short-on and low
states, which suggested emission from the inner disk during main-on and from the outer disk during
short-on and low state (Jimenez-Garate et al. 2002). The well-known 35 d flux modulation has long
been attributed to the precession of an edge-on disk.
-- 21 --
4. By measuring the mean number of line scatterings, we rule out the presence of the X-ray Bowen
fluorescence effect calculated by Sako (2003). We constrain the mean number of scatterings of the
O VIII Lyα1 line to hN i . 69. The upper limit is based on the detection of O VIII and Ne X Lyβ lines,
which must be destroyed after a large number of scatterings. In contrast, the X-ray Bowen fluorescence
effect would require hN i ∼ 5 × 103. Therefore, the N VII/O VIII Lyα line ratio is unaffected by optical
depth effects.
5. The signature of CNO abundances observed with XMM-Newton (Jimenez-Garate et al. 2002) is con-
firmed. Strong N VI Heα, N VII Lyα, and N VII RRC lines are present in the HETGS and in the XMM-
Newton RGS spectra. The N abundance is verified to be higher than Solar. The simplest EM analysis,
the DEM analysis, and the (N VII + N VI)/(O VII + O VIII) line ratio yield 4 . (N/O)/(N/O)⊙ . 9.
The uncertainty is dominated by systematics in the analysis rather than by the effective area calibra-
tion. The depletion of carbon derived from the RGS data is also evidenced in our DEM analysis. This
is supported by measurements of the N/C ratio made by comparing the observed UV spectrum to that
from an illuminated accretion disk model (Raymond 1993). The relative O abundance is sensitive to
the Solar abundance values, being under-abundant with the Wilms, Allen, & McCray (2000) values,
and consistent with Solar with the Allende Prieto, Lambert, & Asplund (2002) values.
6. Given the problems with fitting the He-like ion line fluxes with either the disk model or the DEM, we
deem that either a) the thermal instability regime is over-predicted; or 2) the disk atmosphere model
requires more emission measure at T ∼ 104-105 K. The He-like ion lines are more sensitive probes of the
thermal equilibrium of the plasma than the H-like ion lines. It is possible that the thermal instability
has been over-estimated in the models, therefore cutting the He-like ion fluxes disproportionately.
However, the density diagnostics seem to be consistent with the calculated stable density values.
7. There are line transfer effects which remain to be accounted for in the models, such as the competition
between line emission by photoexcitation and line destruction, and the effects of repeated photoioniza-
tions and recombinations produced by lines and RRC. However, the Her X-1 spectrum indicates that
these are second-order effects, compared to the pure recombination emission.
8. The atmospheric and coronal components unveiled by the X-ray line spectra are more extended (1010 −
1011 cm), cooler, less luminous, and have smaller vertical Thomson depth (τT ∼ 0.01) than the compact
(109 cm) ADC component with τT & 1, which resides close to the compact object. Indeed, both compact
and extended ADC components are evident in the eclipse ingress light curves observed in Hercules X-1
(Choi et al. 1994). Therefore the plasma observed in high resolution spectra is not part of what is
commonly referred to as an ADC (i.e. a fully ionized spherical distribution of plasma that is Thomson
scattering the neutron star X-rays), but instead may be a continuation of it at the Roche lobe scale.
The extended ADC component observed in eclipse coincides with our size measurements for the X-ray
line region.
Allende Prieto, C., Lambert, D. L., & Asplund, M. 2002, ApJ, 573, L137
Bautista, M. A. & Kallman, T. R. 2000, ApJ, 544, 581
REFERENCES
Boroson, B., Vrtilek, S. D., McCray, R., Kallman, T., & Nagase, F. 1996, ApJ, 473, 1079
-- 22 --
Boroson, B., Blair, W. P., Davidsen, A. F., Vrtilek, S. D., Raymond, J., Long, K. S., & McCray, R. 1997,
ApJ, 491, 903
Boroson, B., Kallman, T., Vrtilek, S. D., Raymond, J., Still, M., Bautista, M., & Quaintrell, H. 2000, ApJ,
529, 414
Boroson, B., Vrtilek, S. D., Kallman, T. R., Still, M., Quaintrell, H., McCray, R., Greene, J., & Raymond,
J. 2001, Bulletin of the American Astronomical Society, 33, 1446
Canizares, C. R. et al. 2000, ApJ, 539, L41
Cann, N. M., & Thakkar, A. J. 1992, Phys. Rev. A, 46, 9, 5397
Chiang, J. 2001, ApJ, 549, 537
Choi, C. S., Dotani, T., Nagase, F., Makino, F., Deeter, J. E., & Min, K. W. 1994, ApJ, 427, 400
Clayton, D. D. 1983, Principles of Stellar Evolution and Nucleosynthesis (Chicago: University of Chicago
Press)
Cottam, J., Sako, M., Kahn, S. M., Paerels, F., & Liedahl, D. A. 2001, ApJL, 557, L101
dal Fiume, D. et al. 1998, A&A, 329, L41
Deeter, J. E., Boynton, P. E., Miyamoto, S., Kitamoto, S., Nagase, F., & Kawai, N. 1991, ApJ, 383, 324
Gerend, D. & Boynton, P. E. 1976, ApJ, 209, 562
Giacconi, R., Gursky, H., Kellogg, E., Levinson, R., Schreier, E., & Tananbaum, H. 1973, ApJ, 184, 227
Grevesse, N. & Sauval, A. J. 1998, Space Science Reviews, 85, 161
Houck, J. C. & Denicola, L. A. 2000, in ASP Conf. Ser. 216, Astronomical Data Analysis Software and
Systems IX, ed. N. Manset, C. Veillet, & D. Crabtree (San Francisco: ASP), 591
Hummer, D. G. & Kunasz, P. B. 1980, ApJ, 236, 609
Jimenez-Garate, M. A., Schulz, N. S., & Marshall, H. L. 2003, ApJ, 590, 432
Jimenez-Garate, M. A., Raymond, J. C., & Liedahl, D. A. 2002, ApJ, 581, 1297
Jimenez-Garate, M. A., Hailey, C. J., Herder, J. W. d., Zane, S., & Ramsay, G. 2002, ApJ, 578, 391
Kahn, S. M., Leutenegger, M. A., Cottam, J., Rauw, G., Vreux, J.-M., den Boggende, A. J. F., Mewe, R.,
& Gudel, M. 2001, A&A, 365, L312
Kallman, T. R., Angelini, L., Boroson, B., & Cottam, J. 2003, ApJ, 583, 861
Kallman, T. R. & McCray, R. 1982, ApJS, 50, 263
Kallman, T. R. 1984, ApJ, 280, 269
Kaastra, J. S., Mewe, R., Liedahl, D. A., Singh, K. P., White, N. E., & Drake, S. A. 1996, A&A, 314, 547
Kaastra, J. S. & Mewe, R. 1993, A&AS, 97, 443
-- 23 --
Kinkhabwala, A. et al. 2002, ApJ, 575, 732
Klapisch, M., Schwab, J. L., Fraenkel, J. S., & Oreg, J. 1977, Opt. Soc. Am., 61, 148
Krolik, J. H., McKee, C. F., & Tarter, C. B. 1981, ApJ, 249, 422
Liedahl, D. A. & Paerels, F. 1996, ApJ, 468, L33
Liedahl, D. A. 1999, in X-ray Spectroscopy in Astrophysics, EADN School proceedings, ed. J. A. van Paradijs,
& J. A. M. Bleeker (Amsterdam: Springer), 189
Mewe, R. & Schrijver, J. 1978, A&A, 65, 99
Palmeri, P., Mendoza, C., Kallman, T. R., Bautista, M. A., & Mel´endez, M. 2003, A&A, 410, 359
Petterson, J. A. 1975, ApJ, 201, L61
Petterson, J. A., Rothschild, R. E., & Gruber, D. E. 1991, ApJ, 378, 696
Podsiadlowski, P., Rappaport, S., & Pfahl, E. D. 2002, ApJ, 565, 1107
Porquet, D. & Dubau, J. 2000, A&AS, 143, 495
Ramsay G., Zane S., Jimenez-Garate M., den Herder J., & Hailey C., 2001, MNRAS, submitted
Raymond, J. C. 1993, ApJ, 412, 267
Reynolds, A. P., Quaintrell, H., Still, M. D., Roche, P., Chakrabarty, D., & Levine, S. E. 1997, MNRAS,
288, 43
Sako, M. 2003, ApJ, 594, 1108
Sako, M., Liedahl, D. A., Kahn, S. M., & Paerels, F. 1999, ApJ, 525, 921
Scargle, J. D. in Statistical Challenges in Modern Astronomy (SCMA III), ed. E. D. Feigelson, G. J. Babu
(New York: Springer), 293
Schaller, G., Schaerer, D., Meynet, G., & Maeder, A. 1992, A&AS, 96, 269
Scott, D. M. & Leahy, D. A. 1999, ApJ, 510, 974
Scott, D. M., Leahy, D. A., & Wilson, R. B. 2000, ApJ, 539, 392
Tananbaum, H., Gursky, H., Kellogg, E. M., Levinson, R., Schreier, E., & Giacconi, R. 1972, ApJ, 174, L143
White, N. E. & Angelini, L. 2001, ApJ, 561, L101
White, N. E., Nagase, F., Parmar, A. N. 1995, "The properties of X-ray binaries", pp. 1-57, eds. Lewin, H.
G., Jan Van Paradijs, Van Der Heuvel, E. P., "X-ray Binaries", 1995 (Cambridge University Press).
Wilms, J., Allen, A., & McCray, R. 2000, ApJ, 542, 914
Wilson, R. B., Scott, D. M., & Finger, M. H. 1997, AIP Conf. Proc. 410, Proceedings of the Fourth Compton
Symposium, ed. C. D. Dermer, M. S. Strickman, & J. D. Kurfess (New York: AIP Press), 739
Wojdowski, P. S., Liedahl, D. A., Sako, M., Kahn, S. M., & Paerels, F. 2003, ApJ, 582, 959
-- 24 --
Zane, S., Ramsay, G., Jimenez-Garate, M. A., Willem den Herder, J., & Hailey, C. J. 2004, MNRAS, 350,
506
This preprint was prepared with the AAS LATEX macros v5.2.
-- 25 --
Fig. 1. -- The MEG and HEG count spectra of Hercules X-1 during its low-state shows numerous radiative
recombination lines and continua, a strong Fe Kα fluorescence line, and a continuum which may be produced
by Compton scattering. The large range of temperatures at which these emission lines originate is a direct
probe of the ionization structure of the upper layers of the accretion disk atmosphere and corona.
-- 26 --
Fig. 2. -- HEG spectrum, showing both Fe I-Fe XIII Kα and Kβ fluorescence lines, as well as the Fe XXVI
Lyα radiative recombination line. The disk corona and atmosphere model is over-plotted in red, except for
the fluorescence lines, which were fit independently. The green label indicates that the line was produced in
the model but not detected.
-- 27 --
Fig. 3. -- Added HEG and MEG spectra. The disk corona and atmosphere model is over-plotted in red.
The green label indicates that the line was produced in the model but not detected.
-- 28 --
Fig. 4. -- MEG spectra which show that the nitrogen lines are quite bright relative to the oxygen lines, as
observed with XMM-Newton (Jimenez-Garate et al. 2002). The disk corona and atmosphere model is shown
in red.
-- 29 --
Fig. 5. -- He-like triplet lines for O VII, Ne IX, Mg XI, and Si XIII. The line ratios evolve smoothly from O VII
and Ne IX with R ∼ 0, to Si XIII with R = 3 ± 1. The R ratio of Mg XI has a critical value R = 0.51 ± 0.12,
which can provide a precise measure of the density. The Gaussian line fits (not from the disk model) are
over-plotted in red.
-- 30 --
a)
b)
Fig. 6. -- Light curves of: a) the total HETGS dispersed count rate with λ > 1.5 A, and b) the Fe I-Fe XIII
Kα line HETGS count rate. The orbital period is 1.7 days.
Fig. 7. -- Estimated emission measure (EM) at a single ionization parameter of formation ξo, derived from
the HETGS line fluxes (diamonds) and from the C VI flux measured with RGS (Jimenez-Garate et al. 2002,
filled circle). Both N VII and N VI indicate an excess nitrogen abundance. This rough estimate of the EM
has the virtue of being insensitive to thermal instabilities. We assume a charge state fraction fi+1 = 0.25
for the recombining ions.
-- 31 --
Fig. 8. -- Differential emission measure (DEM) vs.
ionization parameter. The line fluxes measured with
HETGS (diamonds) are complemented with the fluxes from XMM-Newton RGS (Jimenez-Garate et al. 2002,
filled circles). This is more accurate than Fig. 7 because the DEM is calculated in a grid spanning the full
ξ range. However, it is also sensitive to the thermal and ionization balance solution from XSTAR (Kallman
& McCray 1982), as well as the DEM slope. The over-abundance of N and depletion of C is observed. The
predicted thermal instability regime occurs at 1.9 < log10 ξ < 2.4. The excess DEM for the He-like ions in
this regime suggests that the plasma is more stable than predicted.
Fig. 9. -- Density limits derived for the Fe XXVI emission region, versus its maximum distance (r) to the
neutron star. The Thomson depth is plotted for a spherical shell with rmin = r/3 and rmax = r.
-- 32 --
Fig. 10. -- Limits on the density of the Mg XI emission region, versus its maximum distance to the neutron
star. The R ratio strongly limits the plasma to be close to the density and photoexcitation boundary. The
irradiated disk atmosphere model prediction is shown by an X, and the thermally stable density regimes in
that model are shaded in yellow.
Fig. 11. -- Limits on the density of the O VIII emission region, versus its maximum distance to the neutron
star. The velocity broadening suggests that the emission region should be close to the Roche Lobe of the
neutron star. The O VIII edge optical depths for two geometries are shown: for a filled spherical shell with
rmin = r/3 and rmax = r ("shell"), and for a face-on flattened disk geometry ("flat").
✖
-- 33 --
Fig. 12. -- Line emissivity ratio (E) built from the N VI, N VII, O VII, and O VIII line emissivities obtained
from the XSTAR and HULLAC models, as a function of ionization parameter log10 ξ. The observed line
ratio will differ from this emissivity ratio by a factor of AN /AO, the ratio of elemental abundances. At the
top of the figure, we show the log10 ξ range at which the line emissivities are > 20% of the peak emissivity,
with the log10 ξo of formation (peak emissivity) marked.
3p P2
3/2,1/2
Hα
12%
collisional
excitation
Ly β
88%
2
2s S1/2
metastable
2p P2
3/2,1/2
Ly α
2
1s S1/2
Fig. 13. -- Abridged energy level diagram for a H-like ion, showing the decay channels of the 3p 2P3/2,1/2
levels. These channels show how a Lyβ photon gets scattered or converted to other photons. Once a Lyβ
photon is absorbed, there is an 88% probability for an electron at either of the 3p 2P3/2,1/2 levels to re-emit
a Lyβ photon, resulting in a resonant scattering event. Alternatively, there is a 12% probability for the
same initial states to decay by emission of an Hα photon. Once an Hα is emitted, the metastable 2s 2S1/2
level decays to the ground state via two photons (not shown), unless the electron density is above a critical
threshold ncrit. For ne > ncrit, the 2s 2S1/2 level gets excited to 2p 2P3/2, and a Lyα photon is emitted. The
photons shown in the figure are all unresolved doublets.
-- 34 --
Illumination depth
τ
i
neutron star continuum
θ <<1
τ
i >>
τe
Escape depth
τe
recombination & scattered photons
disk atmosphere
Fig. 14. -- In an accretion disk atmosphere which is illuminated with θ ≪ 1, two optical depths are relevant.
In the Her X-1 spectrum, recombination emission dominates over resonance scattering, implying that τi & 100
for the He-like ion r lines. The mean number of scatterings and any line destruction are sensitive to τe. The
X-ray Bowen fluorescence effect is excluded because hN i . 69 and thus τe . 99 for O VIII Lyα1. The
measured τ are beginning to show that the emission region is elongated.
-- 35 --
Table 1. X-ray Emission Features and Disk Atmosphere Model Predictions
Line(s)
λ [A]a
(∆v)σ [km/s]
Observed Fluxb Model Fluxb (λ)
N VI i
N VII Lyα
O VII f
O VII i
O VII r
O VIII Lyα
N VII RRC
Fe XVII
O VII RRC
O VIII Lyβ
O VIII RRC
Ne IX f
Ne IX i
Ne IX r
Ne X Lyα
Ne IX RRC
Ne X Lyβ
Fe XXVI Hα
Mg XI f
Mg XI i
Mg XI r
Ne X RRC
Mg XII Lyα
Mg XII Lyβ
Si XIII f
Si XIII i
Si XIII r
Si XIV Lyα
Si XIV Lyβ
S XVI Lyα
Fe II -- XIII Kα2
Fe II -- XIII Kα1
Fe XXVI Lyα
Fe II -- XIII Kβ
29.08
24.78
22.10
21.800 ± 0.006
21.60
18.970 ± 0.006
18.59
17.05, 17.10
16.78
16.01
14.23
13.70
13.553 ± 0.003
13.45
12.135 ± 0.003
10.37
10.24
9.53, 9.67
9.31
9.229 ± 0.003
9.17
9.10
8.417 ± 0.003
7.11
6.74
6.69
6.647 ± 0.003
6.185 ± 0.003
5.22
4.73
1.940 ± 0.003c
1.936 ± 0.003c
1.78
1.76
· · ·
· · ·
· · ·
< 110
· · ·
< 270
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
< 160
· · ·
< 240
· · ·
· · ·
· · ·
· · ·
< 380
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
< 340
· · ·
· · ·
< 680
< 680
· · ·
· · ·
64 ± 22
25 ± 4.5
1.2 ± 1.2 (< 3.9)
40.5 ± 5.5
3.1 ± 1.6
13 ± 2
18 ± 5
< 3.8
2.5 ± 1.0
0.85 ± 0.30
5.8 ± 1.2
0.17 ± 0.17 (< 0.96)
7.3 ± 0.75
0.87 ± 0.33
3.2 ± 0.4
1.82 ± 0.36
0.59 ± 0.16
< 0.24
0.84 ± 0.17
1.65 ± 0.21
0.59 ± 0.15
0.73 ± 0.19
1.12 ± 0.28
0.42 ± 0.11
1.08 ± 0.15
0.36 ± 0.12
0.54 ± 0.13
1.44 ± 0.23
0.61 ± 0.26
1.19 ± 0.31
10.61 ± 1.84
22.00 ± 1.94
7.37 ± 1.30
5.10 ± 1.20
1.9
2.0
0.36
9.2
2.5
9.0
0.63
0.56
4.2
2.0
2.4
1.4
0.71
0.66
1.8
0.80
0.67
0.79
0.32
0.14
0.10
0.32
1.1
0.30
.38
.18
.11
1.8
0.35
0.88
· · ·
· · ·
1.9
· · ·
Note. -- The model corresponds to an accretion disk atmosphere and corona. The statistical
errors and upper limits are given to 66 % confidence, except for the upper limits, which are 90 %
confidence. Symbols: λ = line wavelength; (∆v)σ = standard deviation of the Gaussian emission
line. We include systematics in the measurements, except for the line fluxes. We compare the data
to the disk atmosphere and corona model.
aWeak features are assigned nominal wavelengths. Otherwise, the measured wavelengths are
shown with error bars. The nominal wavelenghts are rounded up to 0.01 mA.
bIn units of 10−5 photons cm−2s−1.
cThis is the relative shift of the Fe Kα pair with wavelength and flux ratios tied.
-- 36 --
Table 2: Continuum fit parameters
Component (Model) Parameter (Unit)
NH (1018 cm−2)
Absorption (1)
Norm. at 1 keV (phot keV−1 cm−2 s−1)
Power-law (1)
Photon Index
Power-law (1)
Norm. (1039 erg s−1 / (10 kpc)2)
Blackbody (1)
Temperature (keV)
Blackbody (1)
χ2/DOF
(1)
NH (1018 cm−2)
Absorption (2)
Norm. at 1 keV (phot keV−1 cm−2 s−1)
Power-law (2)
Photon Index
Power-law (2)
χ2/DOF
(2)
Value
13 (fixed)
(9.6 ± 0.4) × 10−4
0.27 ± 0.03
(3.5 ± 0.2) × 10−5
0.181 ± 0.015
1292/1844 = 0.70
13 (fixed)
(1.34 ± 0.01) × 10−3
0.53 ± 0.01
1703/1844 = 0.92
Note. -- DOF = degrees of freedom.
Table 3: Helium-like line diagnostics and photoexcitation of the 1s2s 3S1 level
R = f /i
G = (f + i)/r
Ion
Si XIII
Mg XI
Ne IX
O VII
Line
Ratio
3.0 ± 1.0
0.51 ± 0.12
< 0.13
< 0.10
Line
Ratio
2.7 ± 0.7
4.2 ± 1.2
8.6 ± 3.4
13 ± 7
λf→i
(A)(1)
Flux Fλf→i
(10−13 erg cm−2
864
1033
1270
1637
s−1 A−1 )
∼ 1.1(4)
∼ 3.1(4)
1.9 ± 0.1(3)
1.5 ± 0.2(3)
wf
2 3S1 → 1 1S0
Rate (s−1)(2,5)
wf→i(r=1011cm)
2 3S1 → 2 3P0,1,2
Rate (s−1)
3.56 × 105
7.24 × 104
1.09 × 104
1.04 × 103
7 × 104
4 × 105
5 × 105
1 × 106
d
Radius
(cm)
> 8 × 1010
2 × 1011
< 6 × 1011
< 3 × 1012
ne
Density
( cm−3 )
< 5 × 1012
(2 ± 1) × 1013
> 1 × 1013
> 8 × 1011
Note. -- The statistical errors are calculated to 66% confidence and the limits to 90%. Symbols: λf→i = wavelengths of
1s2s 3S1 → 1s2p 3P0,1,2 transitions; wf = radiative decay rates for 1s2p 3S1 → 1s2 1S0 ; wf→i = calculated photoexcitation
rate for 1s2p 3S1 → 1s2p 3P0,1,2. The energy level notation is abridged in the table header.
References. -- (1) Porquet et al. 2001; (2) Drake 1971; (3) from the Hubble GHRS at φ = 0.56 -- 0.60 by Boroson et al. 1996;
(4) Boroson et al. 2001, and private communication 2003, FUSE data near φ = 0.5; (5) From HULLAC atomic code, Klapisch
et al. (1977).
-- 37 --
Table 4. Lyα to Lyβ Line Ratios
Ion
kTo
[eV]
Lyα/Lyβ Lyα/Lyβ
observed
theory
O VIII
Ne X
Mg XII
Si XIV
disk
25
disk
90
disk
154
disk
237
4.7
4.9
3.3
4.6
4.8
4.8
5.5
4.8
15 ± 6
5.4 ± 1.6
2.7 ± 1.0
2.4 ± 1.1
Note. -- The theoretical values were
calculated with HULLAC in the optically
thin limit at the temperature of peak
emission, and for the disk atmosphere
(with temperatures 2 < kT < 860 eV),
which includes possible line blends with
other species. To = average temperature
weighted by the emissivity of the RR line
(fi+1 Sul).
-- 38 --
Table 5. RRC to Radiative Recombination (RR) Line Ratios
Ion
kTo RRC/RR
[eV]
theory
RRC/RR
observed
N VI
N VII
O VII
O VIII
Ne IX
Ne X
2.8
disk
14
disk
3.5
disk
25
disk
6.5
disk
90
disk
0.24
0.26
0.83
0.32
0.32
0.35
0.87
0.27
0.25
0.29
1.02
0.18
· · ·
0.72 ± 0.25, 0.28 ± 0.14(1)
0.06 ± 0.02
0.45 ± 0.12
0.22 ± 0.05
0.23 ± 0.07
Note. -- The theoretical values were calculated with
HULLAC in the optically thin limit. For RR, we use ei-
ther Lyα or Heα, corresponding to each H-like or He-like
ion, respectively. To = average temperature weighted by
the emissivity of the RR line. The disk model tempera-
ture range is 2 < kT < 860 eV.
References. -- (1) Jimenez-Garate et al. (2001).
|
astro-ph/0201323 | 1 | 0201 | 2002-01-19T01:11:16 | New UBVRI colour distributions in E-type galaxies I.The data | [
"astro-ph"
] | New colour distributions have been derived from wide field UBVRI frames for 36 northern bright elliptical galaxies and a few lenticulars. The classical linear representations of colours against log r were derived, with some improvements in the accuracy of the zero point colours and of the gradients. The radial range of signicant measurements was enlarged both towards the galaxian center and towards the outskirts of each object. Thus, the "central colours", integrated within a radius of 3", and the "outermost colours" averaged near the mu_V = 24 surface brightness, could also be obtained. Some typical deviations of colour profiles from linearity are described. Colour-colour relations of interest are presented. Very tight correlations are found between the U-V colour and the Mg2 line-index, measured either at the galaxian center or at the effective radius. | astro-ph | astro-ph |
A&A manuscript no.
(will be inserted by hand later)
Your thesaurus codes are:
ASTRONOMY
AND
ASTROPHYSICS
New UBVRI colour distributions in E-type galaxies
I.The data ⋆
T.P. Idiart1,, R. Michard2,, and J.A. de Freitas Pacheco3
1 Instituto de Astronomia, Geofisica e Ciencias Atmosf´ericas, Depto. de Astronomia, Universidade de Sao Paulo, Av. Miguel
Stefano, 4200-CEP 04301-904, S. Paulo, SP-Brazil; e-mail: [email protected]
2 Observatoire de Paris, LERMA, 77 av. Denfert-Rochereau, F-75015, Paris, France; e-mail: [email protected]
3 Observatoire de la Cote d'Azur, Dept. Augustin Fresnel B.P. 4229, F-06304 Nice Cedex 4, France; e-mail: [email protected]
Received 17 September 2001 / Accepted 7 November 2001
Abstract. New colour distributions have been derived
from wide field UBVRI frames for 36 northern bright el-
liptical galaxies and a few lenticulars. The classical lin-
ear representations of colours against log r were derived,
with some improvements in the accuracy of the zero point
colours and of the gradients. The radial range of signifi-
cant measurements was enlarged both towards the galaxy
center and towards the outskirts of each object. Thus, the
"central colours", integrated within a radius of 3 ′′, and the
"outermost colours" averaged near the µV = 24 surface
brightness could also be obtained. Some typical deviations
of colour profiles from linearity are described. Colour-
colour relations of interest are presented. Very tight cor-
relations are found between the U−V colour and the M g2
line-index, measured either at the galaxian center or at
the effective radius.
Key words: Galaxies: elliptical and lenticulars, CD --
Galaxies: ISM
1. Introduction
The "classical" data on the large scale colour distribu-
tions of E-type galaxies relies on observations by Bender
and Mollenhof (1987), Vigroux et al. (1988), Franx et al.
(1989), Peletier et al. (1990), Goudfrooij et al. (1994), to
quote only the papers discussing the 1-D profiles of colour
against radius, as distinguished from studies of dust pat-
terns. Most of these data were reconsidered by Michard
(2000) (RM00), in an attempt to collect a significant sam-
ple of objects with a complete optical colour set, i.e. U−B,
B−V, B−R and V−I in a coherent photometric system.
This was adequate to confirm previous indications about
the cause of colour gradients: these appear to be due es-
sentially to population gradients within galaxies, with the
Send offprint requests to: R. Michard
⋆ Based in part on observations collected at the Observatoire
de Haute-Provence
dust playing no important role, except in galaxies with
central intense dust patterns. Such objects are rather rare
among the Es.
Similar to most spectral indices of stellar populations,
the colours suffer from the well known age-metallicity de-
generacy, and, except U−B or U−V, are not very sensitive
to the two parameters. They are affected by dust, at least
locally, or perhaps systematically in the central regions ac-
cording to inferences based on a survey by Michard (1999)
(RM99). On the other hand, they may be measured at
lower surface brightnesses or larger radii than the line in-
dices. They could therefore bring useful information to the
study of fossil stellar populations, and further constraints
upon models of the evolution of E galaxies. The present
work aims to provide an enlarged sample of objects with
complete colour data, extending farther in radius than in
previous studies, and hopefully of improved accuracy.
In Paper I, we present the usual information about
the observations and data reduction, and part of the re-
sults in tabular form. A larger set of results will be made
available in electronic form. The frames, partly reduced,
will be made available from the HYPERCAT database,
Observatoire de Lyon.
In Paper II, under the assumption that the observed
colour gradients reflect abundance variations along the
radius, metallicity gradients will be computed from the
present data, using new colour-metallicity calibrations de-
rived from multi-population models for E-galaxies. These
metallicity gradients allow an estimation of central and
mean metallicities. Statistics of galaxies included in our
sample indicate that mean metallicities are about solar,
in agreement with the study by Trager et al. (2000) based
on spectral indices.
Often used notations
-- r isophotal radius; r = (ab)1/2 for an ellipse of semi-
axis a and b.
-- ∆UB
colour
in U−B; ∆UB
d(U − B)/d(log r) and similar for other colours.
gradient
=
2
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
-- diE, boE, unE: subclassification of E galaxies as disky,
boxy and undetermined.
2. Observations
The observations were performed with the 120cm Newto-
nian telescope of the Observatoire de Haute-Provence, in
three runs: April 1-11 2000, May 29-June 5 2000 and Jan-
uary 18-29 2001, noted below as run 1, 2 and 3. Tables 1
to 4 gives lists of the observed galaxies with some param-
eters relevant on the observing conditions. A CCD target
Tek1024 is mounted in the camera, giving a field of view
of 11.6x11.6 ′for a pixel size of 24 microns or 0.68 ′′. The
relatively large field is a favorable feature of this system
for the observation of colour distributions in nearby galax-
ies. Less favourable are the rather poor seeing at the OHP,
with the FWHM of star images usually in the 2-3′′range,
with values at 4 or more during periods of northern wind
(mistral), and also the sky illumination by ever increasing
urban lights.
The camera is unfortunately affected by the so-called
"red-halo" effect.
3. Data analysis
3.1. Outline of the operations
The analysis of the frames entails the following steps:
1. The usual corrections for offset, the "flat-fielding", and
the interpolation of bad columns. As explained below
we tried to improve the flat-fields by ad hoc "super-
flats".
2. The registration of the 5 frames in each passband to a
common geometry, based on a set of measured coordi-
nates for 6 to 12 stars. This is intended to simplify the
derivation of colour maps if needed.
3. The preparation of each frame for measurement, in-
volving a final attempt to measure and correct residual
large scale background trends, corrections for parasitic
objects, a treatment against cosmic rays peaks, and
the calibration against the available results of aperture
photometry (see Sect. 3.2.3).
4. Large errors in colour measurements may result from
small differences between the widths of the PSFs of the
two frames involved (see for instance Michard 1999,
and the previous literature quoted therein). When the
FWHMs of measured PSFs in the colour set for a given
object differed by more than 10%, it was our practice
to modify the frame PSFs and try to make them equal
to that of the best frame of the group (see Sect. 3.2.4).
5. The isophotal analysis of the V frame was performed
according to Carter (1978), as implemented in the Nice
technique described in Michard and Marchal (1994)
(MM94).
6. The correction procedures for the red halo effect in
V−I, and eventually for the effects of different PSF
Table 1. A list of the observations. The dates refer to the
begining of the night. Two values Wo and Wi are given
for the frame FWHM (in ′′): the first is the original one;
the second was attained after the treatment tending to
equalize the FWHM of all 5 frames in a colour set, this at
the value of the best one. Notes: 2974 mistral, interfering
bright star; 3115 mistral; 3193 interfering bright star; 3605
mistral, 3607 also observed; 3608 NGC3607 interfering.
F
NGC Date
2768
id
id
id
id
2974
id
id
id
id
3115
id
id
id
id
3193
id
id
id
id
3377
id
id
id
id
3377
id
id
id
id
3379
id
id
id
id
3605
id
id
id
id
3608
id
id
id
id
3610
id
id
id
id
File
01/04/00 U t613
id
B t614
V t615
id
R t616
id
id
i
t617
05/04/00 U t828
id
B t830
V t829
id
R t831
id
id
i
t832
20/01/01 U p437
B p438
id
V p439
id
R p440
id
id
i
p441
01/04/00 U t619
B t620
id
V t621
id
R t622
id
id
i
t623
20/01/01 U p443
B p444
id
V p445
id
R p446
id
id
i
p447
27/01/01 U p674
B p675
id
V p676
id
id
R p677
id
p678
i
20/01/01 U p449
B p450
id
V p451
id
id
R p452
id
p453
i
05/04/00 U t842
B t843
id
id
V t844
R t845
id
id
i
t846
06/04/00 U t937
B t938
id
id
V t939
R t940
id
id
i
t941
20/01/01 U p456
id
B p457
V p458
id
R p459
id
id
i
p460
Exp
3000
600
300
240
240
3000
600
300
240
240
2400
600
300
200
160
3000
600
300
240
240
2300
600
300
200
160
2500
600
250
160
130
2500
600
250
160
130
3000
600
300
240
240
3000
600
300
240
240
2500
600
250
160
130
Sky
21.50
21.71
20.59
19.93
18.86
20.75
21.47
20.54
20.11
18.47
20.45
21.35
20.44
19.93
18.74
21.09
21.60
20.78
20.03
19.27
21.10
21.72
20.72
20.16
18.85
20.73
21.45
20.47
29.96
18.68
21.28
21.86
20.89
20.32
19.07
21.16
21.73
20.76
20.35
19.35
21.04
21.77
20.79
20.29
19.25
21.39
21.99
21.07
20.49
19.21
Wo Wi
2.32
2.22
2.14
2.29
2.36
5.10
4.63
4.04
4.58
4.23
5.28
4.12
3.81
3.86
3.78
3.11
3.23
3.26
2.90
2.75
3.64
3.02
2.73
2.80
2.76
3.62
2.96
2.70
2.78
2.71
3.96
3.29
2.91
3.11
2.99
4.62
4.13
3.93
3.78
3.88
3.44
2.90
2.82
2.79
2.69
2.99
3.11
2.71
2.51
2.75
2.14
2.11
-
2.14
2.15
4.10
4.05
-
-
-
3.96
3.74
-
-
-
-
-
-
-
-
2.81
2.69
-
-
-
2.75
2.67
-
-
-
2.90
2.84
-
-
-
4.05
-
-
-
-
2.71
2.71
2.67
2.74
-
2.49
2.51
2.44
-
2.43
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
3
Table 2. A list of the observations (continued). See con-
ventions above. Notes: 4125 clouds; 4278 NGC4283 also
observed; 4387 mistral; 4406 (30/5) clouds; 4406 (31/5)
mistral.
Table 3. A list of the observations (continued). See con-
ventions above. Notes: 4551 NGC4550 also observed; 4552
clouds.
3613
id
id
id
id
3640
id
id
id
id
3872
id
id
id
id
4125
id
id
id
id
4261
id
id
id
id
4278
id
id
id
id
4365
id
id
id
id
4374
id
id
id
id
4387
id
id
id
id
4406
id
id
id
id
4406
id
id
id
id
20/01/01 U p462
B p463
id
id
V p464
R p465
id
id
i
p466
07/04/00 U t040
id
B t041
V t042
id
R t043
id
id
i
t044
06/04/00 U t953
id
B t954
V t955
id
R t956
id
id
i
t957
21/01/01 U p521
B p522
id
V p523
id
R p524
id
id
i
p525
21/01/01 U p527
B p528
id
V p529
id
R p530
id
id
i
p531
25/01/01 U p600
B p601
id
V p602
id
R p603
id
id
i
p604
25/01/01 U p606
B p607
id
V p608
id
R p609
id
id
i
p610
01/04/00 U t629
B t630
id
V t631
id
id
R t632
id
t633
i
05/04/00 U t856
B t857
id
V t858
id
id
R t859
id
t860
i
30/05/00 U m606
B m607
id
id
V m608
R m609
id
id
i
m610
31/05/00 U m616
B m617
id
id
V m618
R m619
id
id
i
m620
2500
600
250
160
130
3000
600
300
240
240
3000
600
300
240
240
2500
720
420
280
250
2500
720
420
280
250
2500
600
250
160
130
2500
600
250
160
130
3000
600
300
240
240
3000
600
300
240
240
2400
600
300
150
120
2400
500
250
180
150
21.25
21.92
20.90
20.26
19.03
20.68
21.27
20.40
20.04
19.06
21.03
21.68
20.70
20.17
19.18
20.07
21.08
19.85
18.47
17.40
20.05
21.10
20.25
20.07
18.78
20.96
22.03
21.10
20.33
19.05
20.72
21.75
20.74
20.01
18.64
21.20
21.84
20.85
20.36
19.32
21.23
21.79
20.77
20.35
19.35
20.06
20.69
19.21
17.76
17.28
20.42
20.50
19.24
19.08
18.16
3.58
3.16
2.42
2.16
2.20
3.00
2.29
2.14
2.22
2.26
2.89
2.49
2.38
2.49
2.67
2.13
2.14
1.97
2.48
2.11
2.84
2.61
2.41
2.61
2.53
3.20
2.89
3.54
3.60
3.23
3.52
3.39
2.99
3.17
2.82
3.08
3.20
3.05
2.94
2.85
4.40
4.35
4.08
4.23
4.19
2.65
2.69
2.65
2.39
2.09
4.66
4.27
4.32
4.14
3.85
2.43
2.25
2.11
-
-
2.14
2.14
-
2.18
2.09
2.36
2.43
-
2.37
2.43
1.93
1.94
-
1.98
2.00
2.39
2.40
-
2.50
2.52
2.93
-
2.96
2.98
2.86
2.93
2.95
-
2.94
-
-
-
-
-
-
-
-
-
-
-
2.14
2.19
2.12
2.09
-
2.14
2.19
2.12
2.09
2.09
4472
id
id
id
id
4473
id
id
id
id
4478
id
id
id
id
4486
id
id
id
id
4494
id
id
id
id
4551
id
id
id
id
4552
id
id
id
id
4564
id
id
id
id
4621
id
id
id
id
4636
id
id
id
id
4649
id
id
id
id
07/04/00 U t053
B t054
id
V t055
id
id
R t056
id
t057
i
29/05/00 U m571
B m572
id
id
V m573
R m574
id
id
i
m575
25/01/01 U p600
B p601
id
id
V p602
R p603
id
id
i
p604
27/01/01 U p680
B p681
id
id
V p682
R p683
id
id
i
p684
05/04/00 U t871
id
B t872
V t873
id
R t874
id
id
i
t875
01/06/00 U m647
id
B m648
V m649
id
R m650
id
id
i
m651
02/06/00 U m693
B m694
id
V m695
id
R m696
id
id
i
m697
27/01/01 U p686
B p687
id
V p688
id
R p689
id
id
i
p690
06/04/00 U t966
B t967
id
V t968
id
R t971
id
id
i
t972
27/01/01 U p693
B p694
id
V p695
id
id
R p696
id
p697
i
03/06/00 U m723
B m724
id
V m725
id
id
R m726
m727
i
id
3000
600
300
120
120
2700
600
300
180
150
2500
600
250
160
130
2500
600
250
160
130
3000
600
300
240
240
2400
500
250
180
150
2400
500
250
180
150
2500
600
250
160
130
3000
600
300
120
120
2263
662
300
210
180
2400
500
250
180
150
20.98
21.53
20.60
20.16
19.09
20.73
21.61
20.70
20.22
18.99
21.00
21.59
20.62
19.92
18.62
20.86
21.67
20.70
20.15
18.92
21.18
21.66
20.56
20.04
18.92
20.69
21.44
20.59
20.33
19.43
19.67
21.16
20.26
19.90
18.96
20.86
21.59
20.70
20.00
18.62
20.75
21.50
20.39
19.68
19.24
20.68
20.78
20.25
19.47
18.22
20.15
21.20
20.18
19.82
18.80
2.43
2.49
2.24
2.37
2.32
2.88
3.01
2.99
2.99
2.88
3.56
3.64
3.12
3.11
2.52
4.35
3.98
3.92
4.03
3.62
3.62
3.31
3.47
3.47
3.26
2.48
2.58
2.40
2.16
2.28
2.92
2.79
2.50
2.82
2.99
3.92
3.44
3.56
3.85
3.53
2.85
2.81
3.05
2.82
2.88
3.86
3.61
3.35
3.47
3.05
2.57
2.62
2.52
2.52
2.34
2.29
2.30
-
-
-
-
-
-
-
-
2.58
2.65
2.48
2.54
2.86
3.67
3.59
3.71
3.72
-
3.30
-
-
-
-
2.16
2.15
2.19
-
2.13
2.49
2.48
-
2.51
2.50
3.41
-
-
3.47
-
-
-
2.77
-
-
3.11
3.06
2.99
2.99
-
-
-
-
-
-
4
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
Table 4. A list of the observations (continued). See con-
ventions above. Notes: 5831 mistral and clouds.
Table 5. Measurements of residual fluctuations in back-
ground before and after the final "rectification". Unit: %
of sky background.
5322
id
id
id
id
5576
id
id
id
id
5813
id
id
id
id
5831
id
id
id
id
5846
id
id
id
id
5866
id
id
id
id
5982
id
id
id
id
06/04/00 U t992
id
B t993
07/04/00 V t994
R t080
id
id
i
t081
07/04/00 U t083
B t084
id
V t085
id
R t086
id
id
i
t087
29/05/00 U m579
B m580
id
V m581
id
id
R m582
id
m583
i
31/05/00 U m622
B m623
id
V m624
id
id
R m625
id
m626
i
01/06/00 U m654
B m655
id
id
V m656
R m657
id
id
i
m658
02/06/00 U m700
B m701
id
id
V m702
R m703
id
id
i
m704
03/06/00 U m730
id
B m731
V m732
id
R m733
id
id
i
m734
3000
600
300
240
240
3000
600
300
240
240
2400
500
250
180
150
2400
500
250
180
150
2400
500
250
180
150
2400
500
250
180
150
2400
500
250
180
150
20.95
21.72
20.69
20.62
19.49
20.73
21.22
20.38
19.87
18.79
20.95
21.66
20.71
20.28
19.13
20.43
20.66
18.84
18.06
16.73
20.66
21.28
20.38
20.08
19.08
21.16
21.76
20.80
20.36
19.37
20.89
21.15
20.46
20.18
19.13
2.91
2.64
2.93
1.80
1.87
2.31
2.52
2.67
2.48
2.39
3.47
3.28
3.19
3.09
3.16
3.89
4.07
3.33
3.23
3.19
2.55
2.74
2.07
2.28
2.33
2.43
2.26
1.84
1.91
1.82
2.20
2.30
2.07
2.11
2.33
2.52
2.64
2.53
2.51
2.51
-
2.31
2.30
2.26
-
3.18
3.18
-
-
-
3.28
3.28
-
-
-
2.11
2.16
-
2.11
2.11
1.93
1.88
-
-
-
2.00
2.06
-
-
2.06
far wings in other colours, were performed (see Sect.
3.2.5). The correction necessitates the crossed corre-
lation of the V frame by the I PSF and conversely.
This operation cancels out the errors in the colour dis-
tribution induced by the red halo, but degrades the
resolution. A correction to the calibrations performed
before the convolutions is needed.
7. Finally, colour measurements were performed along
the previously found isophotal contours, and the av-
erage isophotal colours tabulated. Our routine at this
stage involves corrections to the adopted values of the
sky backgrounds, in order to eliminate the obvious ef-
fects of inaccuracies in these (see Sect. 3.2.6).
8. The V surface brightness and colours have been col-
lected in ad hoc files, and corrected for galactic absorp-
tion (or reddening) and the K effect, according to the
precepts and constants given in the Third Reference
Colour
Flats+superflats 1st run
id. 2nd run
id. 3rd run
Final treatment 1st run
id. 2nd run
id. 3rd run
U
1.84
1.90
0.69
0.36
0.40
0.28
B
1.19
0.77
0.79
0.34
0.27
0.28
V
0.72
0.45
0.56
0.30
0.26
0.24
R
0.68
0.54
0.60
0.29
0.31
0.20
i
0.95
0.79
0.64
0.28
0.30
0.32
Catalogue of Bright Galaxies (RC3, de Vaucouleurs et
al., 1991). The usual linear representation of colours
against log r have been calculated in selected ranges,
avoiding on the one hand the central regions affected
by imperfect seeing corrections and/or by important
dust patterns; and on the other the outer range pre-
sumably affected by poor background corrections and
residual noise.
3.2. Detail of operations.
Some important details of the above summarized proce-
dures will now be discussed.
3.2.1. The background
The sky background of flat-fielded frames showed disap-
pointingly large scale trends, specially in the U band. To
improve upon this situation, it was tried to derive cor-
rections by mapping the background of such frames, at
least those which were not "filled" by a large galaxy. Such
maps were found to be correlated, although less so than
expected, and their average used as a "superflat". The
quality of the background was then generally improved: if
not, the superflats were not used. Note that, in the ob-
serving runs of May 2000 and January 2001, a number
of frames of "empty" fields were obtained (sometimes in
moonlight hours) to contribute to the derivation of super-
flats.
A final improvement was obtained by measuring, dur-
ing the treatment of each frame, a number of background
patches, and subtracting a linearly interpolated map of
these, instead of a constant. The background residual large
scale fluctuations were often measured at the three steps
of the procedure, that is, after the application of the flats,
of the superflats, and after the final treatment. Table 5
summarizes the results. It may be noted that the combi-
nation of flats and superflats left large errors in our first
and second runs, specially in the U colour. The final back-
ground linear "rectification" allowed quite significant im-
provements, as seen by comparing the upper and lower
halves of the table.
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
5
If we consider an E galaxy observed under the typical
conditions of the present series (see above for a tabulation
of sky background values), the final residual fluctuations
quoted here represent local errors of less than 0.1 mag near
the isophote µB = 25. We will return later to the question
of errors resulting from background uncertainties.
3.2.2. Parasitic objects
In galaxy photometry it is necessary to remove parasitic
objects, stars and galaxies, that overlap the measured ob-
ject. In the present work we used concurrently the follow-
ing techniques:
. replacement of pixels in a circle enclosing the "para-
site" by a circle symmetric about the center of the galaxy.
. replacement of the circle by another one chosen in a
nearby area.
. marking of the pixels to be discarded in such a way
that they are later left aside in the measuring programmes.
3.2.3. Calibrations
The frames were calibrated by comparisons with the re-
sults of aperture photometry. Our first choice was to use
the UBVRI data of Poulain (1988) (PP88), and Poulain
and Nieto (1994) (PN94) that are available for 26 objects
of our survey, and are in Cousins's system, notably for R
and I. In a few cases the data collected in the HYPER-
CAT catalogues were used. These contain both primary
data (those with an independent photometric calibration),
and secondary data (actually calibrated with part of the
primary data). Only primary data were used, selected ac-
cording to our previous experience or prejudices. A few
completely missing calibrations in R and I were replaced
by values calculated from the tight correlations of V−R
and V−I with B−V, derived from Poulain's data for E
galaxies.
Although the I filter in the camera is of Gunn's type,
our photometry is transfered to Cousins's system through
the calibration. It is assumed that the difference of pass-
bands has no significant effect on colour gradients.
3.2.4. Equalization of PSF FWHM's
Many studies of colour distribution in galaxies are affected
by errors resulting from the difference in the PSFs of the
two frames involved in a colour measurement. Franx et
al. (1989), Peletier et al. (1990), Goudfrooij et al. (1994)
calculated the radial range where the errors due to "dif-
ferential seeing" are larger than some accepted threshold,
and discarded the corresponding colours. Our policy, for
instance in RM99, was to correct for this effect by adjust-
ing the two frames to have PSFs with a common FWHM.
In the present study we tried to equalize the PSFs of
the 5 frames in a given colour set. This is feasible if the
5 frames are taken in rapid succession, so that the PSFs
have similar widths. The frame with the narrowest PSF
is selected, and we find by trial and error a narrow gaus-
sian (or sum of two gaussians) which can restore another
frame to the same quality, or rather the same FWHM, by
deconvolution. The parameters at hand are the σ of the
gaussian and the number of iterations in the deconvolu-
tion. In Table 1 to 4 we give the original FWHM Wo of
each frame, and the improved Wi after the procedure de-
scribed here. Obviously this cannot lead to perfect results,
and sometimes we find in our data the signature of "differ-
ential seeing", in the form of large colour variations, peaks
or dips, within the seeing disk: these defects were edited
out unless there was some good reason to suspect a gen-
uine central colour anomaly, such as large dust patterns,
or the jet of NGC4486.
The reader may notice that two observations of
NGC4406 are listed in Table 2, one of May 30, the other of
May 31 2000. The first one was taken through fog and with
average seeing, while for the other the "mistral" brought
a clearer sky and very poor seeing. A special treatment
was then applied: the central peak of the sharp images
was "grafted" on the corresponding regions of the unsharp
but deeper images. This explains why the Wi is so much
narrower than the original Wo for the frames of May 31.
3.2.5. "Red halo" and PSFs far wings.
Before the start of this survey, the CCD camera on the
telescope used was known to be affected by the "red
halo", an unfortunate property of thinned CCDs. The au-
reoles surrounding stellar images are obviously brighter
and more extended in the I band than in B or V. Not only
the red halo, but more generally the outermost wings of
PSFs, were measured during our observing runs in 2000-
1. The techniques and results are described in Michard
(2001) (RM01). The choice of appropriate star fields al-
lowed us to extend the measurements up to a radius of
nearly 3 ′, and down to a level of about 0.5.10−6 of the
central peak. Due to the red halo effect, PSF wings in I
may be a factor of 3 brighter in an extended radius range
than the V ones. Much smaller but still significant differ-
ences may also occur between the PSFs of various spectral
bands, the V PSF wings always being fainter. The V PSF
wings however, and all the others at the same time, were
reinforced between our observing runs of spring 2000 and
winter 2001, probably an effect of 10 months ageing of mir-
rors coatings. The final output of the measurements are
average "synthetic" PSFs in the format 512x512 pixels, or
5.8x5.8 ′, for each run and pass-band.
A set of numerical experiments on model galaxies is
also presented in RM01, to illustrate the effects of these
far wings on the observed surface brightness and colour
distributions. The most striking effect occurs for the gra-
dient ∆V I , which appears strongly positive, while it is
negative according to the classical results of Bender and
Mollenhof (1987), or Goudfrooij et al. (1994). More subtle
6
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
Table 6. Comparisons of average color gradients for different subsamples of E galaxies, before or after tentative
corrections for the effects of PSF far wings according to RM01. No corrections are applied if the average observed
gradients are close to the adopted reference. N, number of objects. ∆U −B, etc. mean gradients.
Subsample
RM00
2000 Observ.
2000 Correc.
2001 Observ.
2001 Correc.
N ∆U −B
-0.152
29
-0.138
23
-0.116
23
14
-0.174
-0.140
14
σ
0.048
0.037
0.038
0.045
0.036
∆B−V
-0.061
-0.064
-
-0.080
-
∆V −R
σ
0.025
-0.018
0.018 +0.018
-0.016
-
0.022
-0.017
-
-
∆V −I
σ
0.030
-0.053
0.015 +0.093
0.013
-0.048
0.013 +0.040
-0.062
-
σ
0.022
0.047
0.026
0.037
0.025
17 18 19 20 21 22 23 24
uncorrected
corrected
)
V
-
B
(
1.2
1.1
1.0
0.9
0.8
V
Fig. 1. Example of the "correction" of a colour profile
through changes in the sky background constants. Abscis-
sae: V surface brightness in mag. Ordinates: B−V. Uncor-
rected: open circles. Corrected: filled circles. The changes
of the background amount here to -0.25% in B and 0.10%
in V, that is more than average (see Table 7)
effects are found for other colours, with relatively small
but definite changes in gradients.
To correct for the consequences of the red halo, or other
similar effects upon the colour distribution in the index
C1−C2, frame C1 is convolved with the PSF of frame C2
and conversely. After this operation, the resulting images
have been submitted to the same set of convolutions, one
in the atmosphere plus instrument, the other in the com-
puter: they lead therefore to correct colour distributions,
but with a significant loss of resolution. As the convolu-
tions attenuate the central regions of the galaxy, and much
more so for the V frame convolved by the I PSF, the mean
colours are biased: a correction to the calibrations per-
formed before the convolutions is needed. This has been
done by a comparison of simulated aperture photometry
to the observed one, an operation also used to estimate
the errors in calibration (see below).
Since the extended PSFs are found with limited ac-
curacy, it is necessary to discuss the validity of the cor-
responding corrections obtained through crossed convolu-
tions, the more so because of the obvious changes of the
PSF far wings between run 1 and run 3. The mean values
of the colour gradients for subsamples of E galaxies have
been used for these checks, with the results of Table 6.
For a subsample of 12 or more E galaxies the mean colour
gradients and their dispersions cannot differ much, so that
their values may be used as checks of the need for a cor-
rection and its eventual success. 1 The reference for these
comparisons is the subsample in Michard (2000) (RM00),
mostly a rediscussion of the "classical" data by Peletier et
al. (1990), Goudfrooij et al. (1994) and others.
Looking at the Table 6, it is clear that the red halo
introduces enormous errors in the V−I gradients, but that
the corrections are remarkably successful in restoring the
agreement of the results with the accepted reference, both
as regards the mean values and the dispersion. The same
may be said about the V−R gradients. The wings of the V
PSF were strongly reinforced between our run 3 and run
1 or 2, but much less so for the I and R PSF wings. As a
result the red halo effect is less in V−I for the frames of
run 3 and disappears in V−R.
The situation is less clear for the U−B distributions.
Our mean uncorrected gradients are in good agreement
with the "classical" data, essentially from Peletier et al.
(1990), as rediscussed in RM00. On the other hand, the
1 This assertion is questionable as pointed out by the referee.
We therefore tested it by comparing the distributions of the
gradients in subsamples of 15 to 25 objects, sorted out by NGC
numbers, from the surveys of Peletier et al. 1990, Goudfrooij
et al. 1994 and our discussion in RM00. This procures 8 tests,
according to the number of colours in each of the sources. It
turns out that the subsamples are statistically coincident in 4
cases. They differ by slightly more than the calculated errors
in the others.
m
m
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
7
U PSF wings are consistently above the V ones in all our
runs, so that the true slopes of the U−B variations may
be a bit smaller than the observed ones. This error might
well be present in the classical observations. Incidentally,
the data of Peletier et al. were obtained in U−R, and it is
impossible to be sure that the far PSF wings of the used
telescope were the same in both pass-bands!
Similar remarks might be made about our B−V data.
Since the mean measured gradient is the same for our
data of the year 2000 and the adopted reference (and as
a good B PSF is not available) we take as correct this set
of results. For our data of 2001, there is evidence that the
wings of the B PSF were slightly above those of the V one,
so that the B−V gradients might also be biased upwards.
It appears that a significant source of error in the mea-
surement of the small colour gradients in E galaxies has
hitherto been overlooked. It might be that small systematic
errors, of the order of some 15-20%, are still present in the
U−B or U−V gradients published here. Although such er-
rors would not have significant astrophysical implications,
control observations are planned.
3.2.6. Measurements of isophotal colours
Our procedure uses Carter's isophotal representation of
the V frame. The successive contours at 0.1 mag intervals
are fitted to each of the two frames to be compared, the
mean surface brigtnesses calculated and the corresponding
magnitudes and colours derived. A sliding mean smooth-
ing is applied to the data for the outer contours and a
graph of the colour against log r displayed. This might
hopefully be linear, or nearly so, in the studied range.
If it is not, it is our practice to introduce corrections to
the provisional values of the sky background for one or
both of the frames, in order to get rid of the " breaks" in
the colour-radius relation typically associated with a poor
choice of the background constant. The reader is referred
to the graphs published in Goudfrooij et al. (1994) for ex-
amples of such features. In Fig.1 we show a color profile
with a rather important defect due to poor backgrounds,
and its adopted correction.
The introduction of such "aesthetic" corrections to the
raw data might be criticized, since it assumes a regular
behaviour of the colours at large log r. This is however a
reasonable hypothesis: the introduced corrections remain
small, as shown by the statistics of Table 7. It should be
noted that the mean sky background values derived for our
large field frames are more precise than in previous works
based on small field frames, where the sky was not reached
at all. The problem lies in the presence of residual large-
scale background fluctuations (see above): their effects are
similar to those resulting from the poor evaluation of a
constant background, and can be approximately corrected
by the introduction of an ad hoc constant, or rather a set
of constants, for the 5 frames in the colour set.
Table 7. Corrections to provisional sky background val-
ues applied to cancel out "breaks" in the run of colours
aginst log r. The table gives the mean absolute values of
the applied corrections. Unit: % of sky background.
Colour
1st and 2nd run
3rd run
U
0.22
0.14
V
0.12
0.02
R
0.18
0.08
I
0.20
0.14
B
0.14
0.07
2
10
50
100
NGC 4473
)
-
B
U
(
)
V
B
-
(
)
-
R
V
(
)
I
-
V
(
0.6
0.5
0.4
0.3
1.0
0.9
0.8
0.7
0.8
0.7
0.6
0.5
1.4
1.3
1.2
1.1
2
10
r (arcsec)
50
100
Fig. 2. Example of a set of "regular" colour profiles for
NGC4473. In this case, the colours are nearly linear in
log r through the observed range. In this particular case
a linear fit was made in the range 5 -- 80 ′′to provide the
tabulated data.
The linear fit was finally performed on a range of log r
selected so as to avoid the central regions affected by
known dust patterns or possible residual seeing-induced
errors, and the outermost regions visibly affected by devi-
ations from the expected straight line.
3.2.7. Errors
The noise is not a significant source of error in this type of
work, because averages can be performed upon thousands
of pixels in the galaxy regions of low S/N ratio. Residual
noise effects at large r can be easily recognized in sample
8
2
)
-
B
U
(
)
V
B
-
(
)
-
R
V
(
)
I
-
V
(
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
10
50
100
2
10
50
100
NGC 3377
NGC 4125
0.7
0.6
0.5
0.4
1.1
1.0
0.9
0.8
0.7
0.8
0.7
0.6
0.5
1.3
1.2
1.1
1.0
)
-
B
U
(
)
V
B
-
(
)
-
R
V
(
)
I
-
V
(
2
10
r (arcsec)
50
100
2
10
r (arcsec)
50
100
0.5
0.4
0.3
0.2
1.1
1.0
0.9
0.8
0.7
0.7
0.6
0.5
0.4
1.3
1.2
1.1
1.0
0.9
Fig. 3. Example of a set of colour profiles for NGC4125, a
galaxy with a central dust pattern of importance index 3.
In this case, the colours show a hump for r < 10 ′′, small
in U−B but much larger in other colours. In this case, the
linear fit was restricted to the range 10 -- 80 ′′
Fig. 4. Example of a set of colour profiles for NGC3377,
a galaxy with a central red hump in U−B or B−V but
not in V−I. This suggests a metallicity effect. The fit was
obtained in the range 8 -- 80 ′′.
plots of the data (see Fig. 1, 2 , 3 , 4). The main sources
of errors lie:
-- in the calibration, giving a probable error σC , affect-
ing equally all parts of a given object, but dependent
upon the quality of the available aperture photometry
in each colour.
-- in the "differential seeing", i.e. the small differences
in PSF between the frames in a colour set, also after
the adjustments described above. This error σ0 occurs
only near the center of the objects, say in a diameter
of twice the seeing FWHM (it would be a much larger
range without the performed adjustments).There is no
obvious reason for it to be colour dependent.
-- in the sky background residual large-scale fluctuations,
or equivalently, errors in the sky background estimates.
This error σS is strongly dependent upon colour and
the studied radius, or rather the corresponding surface
brightness.
These three components of the total error will now be
considered, together with other relevant topics.
1. Errors in colours from calibration inaccuracies.
When a number of measurements are available for a
given object, for instance 5 apertures in PP88 and
PN94, a probable error of the resulting calibration is
readily derived from the dispersion of these measured
values about the results of simulated aperture pho-
tometry, calculated from our data, i.e. V magnitudes,
isophotal parameters and colours. For the preferred
calibrations with Poulain's data, the computed error
is often less than 0.01 in B−V or V−R but may rise
to 0.02 in U−B or V−I. Still larger calibration errors,
up to 0.04, have been estimated for objects with very
scanty or uncertain aperture photometry. These prob-
able errors apply to the zero point of the colour regres-
sion as shown in Table 9, and to all colour data from
the same object.
2. Residual errors from PSF equalization.
The residual errors after this step in our data treat-
ment can be objectively ascertained by comparing the
"central" B−R colours in our survey with the equiv-
alent data from RM99, derived from high-resolution
CFHT frames. For the present survey the "central
colours" are the integrated colours within a radius of
3 ′′. From RM99, Table 6, we find the colours at the
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
9
isophote of 1.5 ′′, which are likely similar. The statistics
of the difference B−R(new) -- B−R(RM99) are for 31
objects in common: mean=0.01; σ = 0.038. Assuming
then that the errors are equal in the two surveys, the
probable error associated with poor PSF adjustements
is σ0 = 0.027 in B−R. This source of error has no rea-
son to vary significantly from one colour to another.
It is independent of the error of calibration previously
discussed.
The comparison between the two surveys is made pos-
sible because the same set of calibrations has been
used, although the field of the CFHT frames was often
not sufficient to use all calibrations apertures. A mi-
nor part of the above differences may come from this
source. In RM99, the B−R of NGC2768 is quoted too
red by 0.08 and that of NGC3610 too red by 0.10.
3. Errors from sky background inaccuracies
Given ǫA and ǫC, the relative errors in the sky back-
ground evaluation for frames A and C, A and C bee-
ing a pair among UBVRI, the magnitude error in the
colour A−C may be expressed as δA−C = −2.5 log(1 +
ǫA100.4∆µA) + 2.5 log(1 + ǫC 100.4∆µC ). Here ∆µA is
the magnitude contrast between the object and the
sky in colour A. Using average values for the colours
of E-galaxies and for the sky brightnesses, ∆µA for
any colour may be expressed in terms of ∆µV . Then,
in the range of small δA−C , the expression reduces to
δA−C = 1.0857(KAǫA − KCǫC )100.4∆µV where KA =
100.4∆µA/100.4∆µV .
KV = 1 by definition; we find KU = 3.1 and KI =
1.55, while KB and KR are slightly below 1.
The ǫA,... are unknown, but it is feasible to get statis-
tics of the linear combinations KAǫA − KCǫC . Indeed,
as explained above, we have adopted ad hoc correc-
tions to provisional sky background values, in order to
regularise the colour-log r relations. It is reasonable to
assume that the errors left after these corrections are
proportional to the adopted corrections themselves. We
take δA−C = ηKAC = η(KAǫA − KC ǫC), where η is a
small constant and the KAC may be derived from the
statistics of the adopted corrections given in Table 7.
In practice somewhat different statistics have been cal-
culated, to take into account the fact that our correc-
tions for two colours are not necessarily uncorrelated.
The constant η, different for our observing runs of 2000
and 2001, was chosen so as to get a system of errors
compatible with the appearance of the data and also
with the errors found for the slopes of the colour-log r
relation. The finally adopted errors σS from sky back-
ground inaccuracies are given in Table 8. Note that this
source of error is negligible for µV = 22 or smaller. Pre-
dicted errors are reduced for our run 3 as compared to
the two others.
4. Total errors
The above estimated errors are independent and
should be added quadratically. In the central region
Table 8. Probable errors σS associated with sky back-
ground inaccuracies. Units: magnitude. Estimated errors
are the same in V−R as in B−V.
Colour
1st and 2nd run
id
id
3rd run
id
id
µV
23
24
24.5
23
24
24.5
U−B U−V B−V V−I
0.012
0.023
0.030
0.058
0.048
0.093
0.017
0.010
0.026
0.042
0.067
0.041
0.026
0.064
0.102
0.019
0.048
0.077
0.008
0.020
0.032
0.006
0.015
0.024
0.00
0.05
0.10
0.15
)
-
B
U
(
0.25
0.20
0.15
0.10
0.05
0.00
D (B-V)
Fig. 5. Correlation between the colour gradients, with the
signs changed. Abscissae: ∆BV . Ordinates: ∆UB. Com-
pare with the similar diagram in RM00, or with the origi-
nal correlation diagram between ∆BR and ∆UR in Peletier
et al. (1990). The dispersion is clearly reduced here, which
can only be attributed to an improved accuracy.
0)1/2.
with r < 6 ′′, the total error will be σT = (σ2
In the mean region with 18 < µV < 22 the total error
equals the calibration error. Finally, in the outer range
of µV > 23 one can use σT = (σ2
C + σ2
C + σ2
S)1/2.
5. Errors from the corrections for red halo and other far
wings effects
To our knowledge, the crossed convolutions used to
correct for the red halo and similar effects do not give
rise to random errors, but rather to systematic errors
due to inaccuracies in the adopted PSFs. Such prob-
lems may be detected from the study of the distri-
butions of the slopes of the colour-log r relations dis-
cussed below.
6. Errors in the slopes of the colour-log r relations
These have been calculated by two complementary
methods. On the one hand, we can look for the corre-
D
D
D
10
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
0.00
0.05
0.10
0.15
0.15
0.6
0.7
0.8
0.9
1.0
1.1
1.2
)
I
-
V
(
re/8
re/2
V= 24
)
-
B
U
(
NGC 3610
0.10
0.05
0.00
0.7
0.6
0.5
0.4
0.3
0.2
D (B-V)
(B-V)
Fig. 6. Correlation between the colour gradients, with the
signs changed.. Abscissae: ∆BV . Ordinates: ∆V I . Here the
comparison with similar diagrams in RM00 does not show
much increase in the accuracy of the new data.
lations between the slopes derived here and those from
the literature, notably the data collected and discussed
in RM00. Assuming then that the errors are the same
in both sources, we get an estimate of our slope errors,
hopefully an upper limit. On the other hand we can
consider the internal correlations between the slopes
of the various colour-log r relations, specifically ∆UB
and others with ∆BV . Figure 5 and 6 shows the cor-
relations of the U−B and V−I colour gradients with
that in B−V. The coefficients of correlation are re-
spectively 0.73 and 0.40. A weighted mean Gm4 of the
slopes in the 4 colours has also been used as reference
instead of ∆BV with analogous results. From the dis-
persions of such correlations the slope errors can be
estimated, if the error for the reference dB − V /d log r
or Gm4 is "guessed". The two techniques give results
in very good agreement, the internal correlations indi-
cating somewhat smaller errors.
The probable errors of the slope estimates are then
0.03 in U−B or U−V, 0.01 in B−V, 0.015 in V−R
or B−R, 0.02 in V−I.
Fig. 7. Colour-colour diagram of U−B against B−V: the
colours are calculated from the linear representations of
Table 9 at the effective radius re/8 (dots), at re/2 (circles)
and at the outermost range near µV = 24 (stars) from
Table 11. Although the various symbols refer to widely
different regions of the galaxies they define a common re-
lation.
1.1
1.2
1.3
1.4
1.5
1.6
1.7
re/8
re/2
V= 24
)
I
-
V
(
(U-V)
1.5
1.4
1.3
1.2
1.1
1.0
0.9
4. Observational results
The available data from the present study are:
1. Table 9, giving for each object the linear representation
of the colour-log r relations, i.e. the selected inner and
outer radii of the fit, the zero point colour with its
probable error σC and the slope. The probable errors
Fig. 8. Colour-colour diagram of V−I against U−V: the
colours are calculated from the linear representations of
Table 9 at the effective radius re/8 (dots), at re/2 (circles)
and at the outermost range near µV = 24 (stars) from
Table 11.
D
D
D
m
m
m
m
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
11
for the slopes are given above. This table allows easy
calculation of the colours at any radius, notably the
effective radius re and others as considered below.
2. In electronic form only, the tables giving for each
galaxy, as a function of the radius r, the V magnitude
µV and the colours U−B, B−V, V−R, V−I. These ta-
bles are presently available with the U−B indices as
observed, and consistent with Table 9. They will be
later made available after correction for suspected ef-
fects of PSF far wings (see discussion in 3.2.5).
3. Series of graphs from the above tables, showing the
colours as a function of log r or of µV . Examples of
these graphs are shown here to illustrate a number of
properties of the radius-colour relations.
4. Table 10 gives the "central" colours according to sev-
eral definitions: we have considered the colours inte-
grated inside the area of radius r = 3 ′′, and the colours
calculated for r = 1.5 ′′from the linear representations
of Table 9. They should be nearly equal if these repre-
sentations remain valid at small r, which is not always
the case: see below for a description of typical devia-
tions.
5. Table 11 collects colours measured at the outermost
range of the available data expressed in V magni-
tude. This gross limit varies between µV = 23 (for
NGC4472) and 24.5. It is controlled by the size of the
object and the "cleanliness" of the nearby field.
The SA0 galaxies NGC3115, 3607, 4550 and 5866 have
been observed with the E-type sample. The corresponding
results are given in the Tables, but they have been discarded
from the discussion.
4.1. Statistical comparison with previous work
-- Assuming that colours result only from population
variations, "perfect" correlations between zero point
colours in different pass-bands are expected. We have
compared colour-colour correlations for previous sur-
veys and the present one. For B−I against B−V we find
from Goudfrooij et al. (1994) a coefficient of correla-
tion ρ = 0.61 (41 objects) compared to ρ = 0.94 from
Table 9 (37 objects). For U−R against B−R the data
of Peletier et al. 1990 (38 objects) give ρ = 0.58, while
the present data leads to ρ = 0.94. The improvement
is less if we compare our results with the discussion in
RM00 where the calibration of the available photome-
try was reconsidered.
-- The correlation between colour gradients should also
be as good as allowed by errors of measurements.
Again we compare the correlation coefficients between
gradients from the literature and our results for the
same colours. From Goudfrooij et al. (1994), consider-
ing the gradient ∆BI against ∆BV , we find ρ = 0.78
which is quite good. The result derived from Table 9
is still better, i.e. ρ = 0.85. The improvement is more
pronounced when the ∆UR against ∆BR correlation
is calculated from Peletier et al. (1990): one obtains
ρ = 0.36 only, instead of ρ = 0.88 with our data. 2
-- Our colour measurements could be obtained at much
lower surface brightness (or larger radii) than in pre-
vious work. The present data extend to 23.2 < µV <
25.2, with a median value near µV = 24.5 in all colours.
The tables by Goudfrooij et al. (1994), available from
CDS Strasbourg, are mostly limited to µV = 22.75
in B−V and µV = 22.25 in V−I (median values),
due to the small fields of the frames, notably in the
I band (and probably an ad hoc cut off). Their pub-
lished graphs generally extend to the same radius in
both colours. The tables from Peletier et al. (1990),
again at the CDS, extend to µV = 23.2 in B−R and
µV = 22.6 in U−R (mean values) due to a systenatic
cut-off at 10% of the sky. The printed tables may be
still more severely truncated.
In summary, our colour data generally extends 1.5 to 2
magnitudes deeper than in previous works, so that "ex-
ternal colours" refering to the level µV = 24.5 when-
ever possible, are presented in Table 11 with realistic
error estimates.
4.2. Description of isophotal colour profiles.
Most of the profiles relating colour to log r, or equivalently
to the surface brightness µV , are regular, meaning that
they deviate very little from a straight line in the range of
abscissae relevant to the present data, roughly r > 2 ′′and
µV < 24.5. In this case, the "central colours" integrated
within r < 3 ′′, and the colours calculated from the linear
representation at the average radius of r = 1.5, differ very
little. Central colour according to these two definitions are
given in Table 10: compare columns (5) and (6) for U−V,
and (3) with (7) for V−R. Figure 2 gives an example of
a set of regular colour profiles for NGC4473. Of course
the linear colour -- log r relation is only approximate and
breaks down at small r for all galaxies observed at high
resolution. Carollo et al. (1997) obtained V−i maps of a
number of E-galaxies from HST frames, disclosing minute
colour structures in some cases. In RM99, the CFHT reso-
lution proved sufficient for the detection and classification
of B−R central " red peaks". These are smoothed out, at
least partly, with the OHP seeing.
Non regular profiles have been observed in the follow-
ing cases:
1. When an important dust pattern occurs near the cen-
ter of an object, it produces a central red hump in the
colour profile. This is the case for galaxies with the
value of 3 for the dust pattern importance index (DPII)
introduced in RM99, such as NGC2768, 4125 and 4374,
2 The coefficients of correlation given here are of course dif-
ferent from the previously calculated ones in Sect. 3.2.7.6 ,
which refered to other colours with smaller gradients.
12
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
but also for 5813 and 5831. The dust ring of NGC3607,
type SA0, produces noticeable bumps in its colour pro-
files. Fig. 3 gives an example of the colour profiles of
such a centrally dusty galaxy, i.e. NGC4125. In such
cases, the integrated colours within r = 3 ′′are redder
than the extrapolated colours at r = 1.5: this "extra
reddening" is smaller in U−B than in other colours.
The consideration of the central reddening in V−R, a
colour nearly insensitive to age-metallicity variations
but sensitive to dust, gives a possibility to correct other
colours for dust effects. This has been done, in some
applications, for galaxies of DPII 3.
2. A few objects show a central red hump in some of
the colour profiles, specially U−B and also eventually
B−V. This is the case of NGC3377, 3379, 3610(?) and
4494. Fig.4 illustrates the case of NGC3377. For such
objects the "extra reddening" defined above is near
zero in V−R and V−I. It is therefore permitted to
attribute it to a metallicity increase rather than to
dust.
3. On the contrary, the central colours U−B, B−V may
be bluer than the extrapolation of the linear portion of
the profile. This occurs for objects with larger than av-
erage colour gradients. The best such case is NGC4636;
others are 4283, 4478 and the SA0 4550. An "extra
blueing" in U−V is then apparent from the compar-
ison of indices in columns (5)-(6) of Table 10. This
is also marginally the case for such giant galaxies as
NGC4406, 4472, 4649: according to RM99 the colour
profile of such objects tend to flatten out near the cen-
ter.
4. NGC4486 is remarkable in showing a central "blue
deep", the U−B colour beeing 0.51 at r = 0 as com-
pared to 0.72 near r = 6.5 ′′. This central feature is
probably somehow related to the famous non-thermal
jet of this galaxy. The jet is of course very conspicuous
in U−B , with a peak colour near -0.1. Needless to say,
both the central "blue deep" and the jet are affected
by seeing (and our attempts to correct its effects).
4.3. Correlations of interest.
4.3.1. Correlations between colour gradients.
The correlations between the various colour gradients have
already been noted as useful tools to evaluate the probable
errors in gradients. These correlations are displayed in Fig.
5 and 6.
The coefficient of correlation between ∆UB and ∆BV
is 0.75; that between ∆V I and ∆BV is only 0.40. For ∆V R
it falls to 0.20, because the errors are of the same order
of magnitude as the V−R gradients. Imposing regression
lines running through the origin, the relative slopes are
< ∆UB/∆BV = 2.2 >, < ∆V I /∆BV = 0.8 > and <
∆V R/∆BV = 0.25 >. These relative slopes are in good
agreement with the results in RM00, and its conclusion,
1.3
1.4
1.5
1.6
1.7
1.8
)
g
a
m
(
C
2
g
M
(U-V)C
0.45
0.40
0.35
0.30
0.25
0.20
0.15
Fig. 9. Correlation between the near center U−V colour
in abscissae, and the M g2 index from Faber et al. 1989,
in ordinates
1.1
1.2
1.3
1.4
1.5
1.6
0.4
e
r
2
g
M
0.3
0.2
0.1
(U-V)re
Fig. 10. Correlation between the U−V colour and the
M g2 index at the effective radius. This index, and the
M gb one merged in the data, are taken from Kobayashi
and Arimoto, 1999.
i.e. the negligible influence of dust upon colour gradient,
is confirmed.
The distribution of colour gradients for E-galaxies
may be of interest. The following parameters are found:
< ∆UB >= −0.152 with σ = 0.044; < ∆BV >= −0.070
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
13
with σ = 0.021; < ∆V R >= −0.018 with σ = 0.012;
< ∆V I >= −0.054 with σ = 0.023. The dispersions are
not much larger than the errors estimated above. The
distributions are asymmetric: there are 4 objects with
∆BV larger than +1.8σ above the mean, but none at less
than the same deviation. These galaxies, with ∆BV clearly
steeper than average, by about twice the estimated prob-
able error of measurement, are NGC4283, 4478, 4564 and
4636, seemingly a random collection.
Remark: An attempt to sort the E-galaxies by flatten-
ing as measured in MM94, and to look for some relation to
the gradients, lead to negative results. Similarly no signifi-
cant difference was found between diE and other galaxies.
The SA0 NGC4550 has quite exceptional gradients in
all colours, and an admixture of dust and relatively young
stars could be invoked to explain its properties. This ob-
ject also has very remarkable kinematics, as first described
by Rubin et al. (1992); a model has been proposed by Rix
et al. (1992). The few other S0s in the present sample are
similar to Es with regard to their colour gradients.
4.3.2. Colour-colour diagrams
Many colour-colour diagrams can be built from the present
data. The indices used may be calculated from the linear
representations in Table 9 at the effective radius re, at a
near center position re/8, and at an intermediate position
re/2. The system of re here used is an average of esti-
mates in MM94, Prugniel and H´eraudeau, 1998, and the
RC3. It is satisfactory that the various indices (B − V)re ,
(B − V)re/2, (B − V)re/8, define a common diagram with
the corresponding U−B. This may suggest that a common
physical variable controls the variations inside an object,
and the object to object changes, of the two colours. Such
graphs readily show larger than average calibrations er-
rors: for instance, the U−B colour of NGC3610 is clearly
too red for its B−V.
We have also traced colour-colour diagrams for the
"central" colours, (Table 10), i.e. integrated in the radius
r < 3 ′′. They are similar to those traced with the in-
terpolated colours, but with larger dispersions: this is not
surprising since the central colours suffer from larger er-
rors (see 3.2.7).
Finally, one can trace colour-colour diagrams with the
"outermost colours" collected in Table 11. They are in fair
agreement with the diagrams derived from Table 9, and
extend these towards the blue. We show in Fig. 7 a com-
posite colour-colour diagram U−B against B−V, using the
colours ar re/8, re/2 and the outermost range. Similarly
Fig. 8 displays the diagram of V−I against U−V.
4.3.3. The U−V colour as a metallicity index for
ellipticals ?
Burstein et al. (1988) showed a correlation between the
central M g2 index and a global B−V colour measured in
a large aperture (see also Bender et al. , 1993). This type
of correlation is reconsidered here using the U−V colour
which is much more sensitive to metallicity than B−V, and
taking advantage of recent estimates of the M g2 index far
from center.
Two correlations between the M g2 index and U−V are
considered in Fig. 9 and 10. The first shows the relation
between the two quantities near the galaxy center: the
M g2C index is taken from the tabulation by Faber et al.
1989. The (U − V)C index is the integrated colour in a
circle of radius r = 3 ′′. The value for three galaxies with
important central dust patterns have been corrected by
reference to the central bump in (V−R), a colour sensitive
to dust but less so to metallicity changes. The coefficient of
correlation reaches 0.825. Taking (U − V)C as x and M g2C
as y we find the regression y = 0.221 ±.027x−0.061 ±.003.
The second, i.e. Fig. 10, displays the correlation be-
tween the U−V colour and the M g2 index at the ef-
fective radius re. M g2re has been taken from Kobayashi
and Arimoto (1999) (KA99). To increase the number of
data points the M gb gradients were introduced, using
the linear relation M gb = 15 M g2 derived from the dis-
tribution of the values of both indices in the tables of
KA99. The coefficient of correlation still reaches 0.72.
Again, with the colour in x and the M g2re in y we find
y = 0.283 ± .069x − 0.133 ± .007. The difference to the
above correlation for central indices is barely significant.
The quality of these correlations proves that both in-
dices are essentially controlled by the same physical vari-
ables, and leaves little room for the effects of diffuse dust
upon the colours of E-galaxies.
5. Conclusions
New colour-radius relations have been derived for 36 E-
type galaxies of the northern Local Supercluster, using
UBVRI frames obtained with the 120cm telescope of the
Observatoire de Haute-Provence. Four SA0, i.e. NGC3115,
3607, 4551 and 5866 were also observed.
We aimed to take advantage of the large field of the
camera to observe the galaxies at larger radii than hitherto
feasible, and thus improve the accuracy of colour gradi-
ents. The availability of the series of aperture photometry
in PP88 and PN94 for most of the sample was also consid-
ered an asset towards a more coherent system of colours.
It appears indeed that the colour calibrations are improved
here compared to previous work, if this can be judged from
the quality of correlations between zero point colours in
various surveys (see Sect. 4.1)
Two steps in the reduction procedure were thought
significant in improving the quality of colour profiles: the
first was the adjustment of the FWHM of the PSFs in a
given colour set of 5 frames to the best of the five. This
allowed us to get significant colours much closer to the
galaxy center than otherwise feasible. The second was a
careful "mapping" of the background of each frame, in or-
14
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
Table 9. Linear representation of colour against log r. Successive columns: NGC No.; Type; ri inner radius of calcu-
lation; ro outer radius of calculation; r0 radius of colour evaluation (in log and "); U−B at r0 and estimated standard
error; ∆UB radial gradient; B−V at r0 and estimated standard error; ∆BV radial gradient; V−R at r0 and estimated
standard error; ∆V R radial gradient; V−I at r0 and estimated standard error; ∆V I radial gradient; du Dust visibility
index (Michard 1999) Notes: 2974: V−R and V−I not measurable; 3193: V−I not measurable.
NGC Type
2768
2974
3115
3193
3377
3377
3379
3605
3607
3608
3610
3613
3640
3872
4125
4261
4278
4283
4365
4374
4387
4406
4472
4473
4478
4486
4494
4550
4551
4552
4564
4621
4636
4649
5322
5576
5813
5831
5846
5866
5982
diE
diE
SA0
unE
diE
diE
unE
boE
SA0
boE
diE
diE
boE
diEp
diE
boE
diE
unE
boE
unE
boE
boE
unE
diE
boE
unE
unE
SA0
boE
unE
diE
diE
unE
unE
boE
boEp
unE
diE
unE
SA0
boE
ri
10
8
8
6
8
8
8
4
15
5
4
6
5
5
10
5
10
5
5
8
5
5
5
5
6
10
5
5
5
5
5
5
7
5
5
5
5
4
5
15
8
ro
80
80
100
40
80
80
100
30
100
50
50
60
60
60
80
80
60
30
120
80
30
80
150
80
50
120
80
40
40
100
50
80
100
150
80
60
60
50
100
110
100
r0
1.467
1.404
1.464
1.169
1.399
1.403
1.460
1.043
1.569
1.201
1.159
1.287
1.266
1.202
1.445
1.320
1.361
1.080
1.430
1.394
1.102
1.312
1.510
1.310
1.286
1.559
1.340
1.199
1.184
1.366
1.233
1.327
1.416
1.515
1.313
1.216
1.229
1.175
1.407
1.612
1.447
U−B
0.521±.03
0.429±.01
0.501±.01
0.461±.02
0.298±.02
0.289±.02
0.538±.02
0.403±.03
0.451±.02
0.415±.03
0.448±.04
0.470±.03
0.455±.03
0.506±.03
0.497±.03
0.588±.03
0.408±.02
0.410±.02
0.552±.02
0.497±.03
0.431±.02
0.463±.02
0.590±.02
0.423±.03
0.300±.02
0.566±.02
0.452±.03
0.260±.02
0.475±.02
0.489±.01
0.369±.02
0.522±.02
0.495±.02
0.592±.01
0.445±.03
0.382±.02
0.477±.01
0.423±.03
0.594±.02
0.367±.02
0.431±.02
∆U B
-0.093
-0.237
-0.169
-0.163
-0.144
-0.170
-0.111
-0.175
-0.111
-0.189
-0.132
-0.174
-0.120
-0.127
-0.165
-0.170
-0.149
-0.256
-0.120
-0.107
-0.106
-0.121
-0.133
-0.169
-0.234
-0.192
-0.114
-0.252
-0.113
-0.171
-0.247
-0.182
-0.169
-0.121
-0.146
-0.147
-0.076
-0.124
-0.129
-0.190
-0.117
B−V
0.892±.02
0.905±.01
0.912±.01
0. 924±.01
0.854±.01
0.852±.01
0.918±.01
0.821±.02
0.882±.01
0.949±.01
0.785±.01
0.875±.02
0.892±.02
0.928±.02
0.890±.02
0.948±.01
0.845±.01
0.870±.01
0.939±.01
0.915±.01
0.883±.01
0.961±.01
0.963±.01
0.881±.01
0.806±.01
0.921±.01
0.862±.01
0.826±.01
0.880±.01
0.943±.01
0.884±.01
0.919±.01
0.906±.02
0.982±.01
0.843±.02
0.821±.01
0.945±.01
0.876±.02
0.955±.02
0.819±.01
0.868±.02
∆BV
-0.054
-0.080
-0.069
-0.086
-0.064
-0.075
-0.043
-0.062
-0.067
-0.064
-0.075
-0.074
-0.043
-0.072
-0.083
-0.089
-0.088
-0.117
-0.058
-0.044
-0.052
-0.078
-0.040
-0.063
-0.103
-0.063
-0.050
-0.196
-0.040
-0.080
-0.111
-0.080
-0.115
-0.053
-0.067
-0.076
-0.045
-0.078
-0.051
-0.100
-0.062
V−R
0.536±.02
-
0.585±.01
0.567±.02
0.510±.01
0.505±.01
0.585±.01
0.517±.02
0.537±.02
0.542±.01
0.503±.01
0.533±.04
0.544±.02
0.571±.01
0.568±.03
0.585±.01
0.554±.03
0.533±.02
0.588±.01
0.565±.02
0.565±.01
0.561±.01
0.592±.01
0.598±.01
0.525±.01
0.601±.01
0.518±.02
0.511±.01
0.525±.02
0.546±.01
0.545±.01
0.584±.01
0.562±.02
0.571±.01
0.530±.02
0.521±.01
0.589±.01
0.515±.02
0.586±.02
0.537±.01
0.522±.02
∆V R
-0.012
-
-0.010
0.009
-0.014
-0.028
-0.013
-0.014
-0.017
-0.020
0.003
-0.018
-0.012
-0.005
-0.011
-0.014
-0.031
-0.040
-0.011
-0.020
-0.028
-0.018
-0.005
-0.021
-0.023
-0.019
-0.016
-0.108
-0.014
-0.035
-0.024
-0.004
-0.034
-0.036
-0.001
-0.003
-0.022
-0.040
-0.016
-0.048
-0.043
V−I
1.136±.02
-
1.222±.01
-
1.067±.01
1.074±.01
1.206±.01
1.076±.02
1.149±.02
1.175±.01
1.144±.02
1.100±.04
1.162±.03
1.176±.02
1.183±.03
1.246±.02
1.145±.03
1.147±.02
1.226±.02
1.189±.02
1.170±.02
1.204±.01
1.257±.01
1.225±.01
1.129±.01
1.235±.02
1.137±.02
1.119±.02
1.151±.02
1.191±.02
1.124±.02
1.216±.02
1.185±.03
1.231±.02
1.091±.03
1.126±.01
1.244±.02
1.168±.02
1.245±.02
1.095±.02
1.145±.02
∆V I
-0.081
-
-0.047
-
-0.080
-0.105
-0.037
-0.038
-0.046
-0.038
-0.030
-0.050
-0.045
-0.078
-0.046
-0.048
-0.093
-0.078
-0.057
-0.025
-0.036
-0.054
-0.010
-0.065
-0.051
-0.068
-0.019
-0.185
-0.027
-0.074
-0.083
-0.033
-0.105
-0.048
-0.028
-0.040
-0.069
-0.075
-0.035
-0.042
-0.080
du
3
3
0
0
1-
1-
1-
0
3
1-
0
1-
0
0
3
1-
?
0
0
3
0
0
0
0
0
0
?
0
0
0
1-
0
?
0
0
0
1
?
0
3+
?
der to lessen the background fluctuations remaining after
the usual flat-fielding procedures. Both these precautions
proved successful, and, as a result, the radial range of sat-
isfactory colour measurements was greatly enlarged. Near
the galaxy center, it proved feasible to obtain "central
colours", i.e. colours integrated in the circle r = 3′′, in fair
agreement with high resolution data (see Sect. 3.2.7.2 and
Table 10).
On the other hand, colours could be obtained at much
lower surface brightness (or larger radii) than in previous
work. Our colour data extend to 23.2 < µV < 25.2, with a
median value near µV = 24.5 in all colours. According to
the comparisons in Sect. 4.1, this is 1.5 to 2 magnitudes
deeper than in previous work. "External colours", refering
to the level µV = 24.5 whenever possible, are published
for the first time (see Table 11), and may be useful to give
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
15
Table 10. Columns (1) to (5): integrated colours within
a radius r = 3 ′′, i.e. U B3, BV3 , V R3 , V I3, U V3; (6):
U V1.5 U−V colour at r = 1.5′′ calculated from Table 9 .
(7) V R1.5 V−R colour at r = 1.5′′
NGC U B3 BV3
1.04
2768
2974
1.03
1.03
3115
0.96
3193
0.98
3377
3377
0.98
1.03
3379
0.90
3605
1.04
3607
1.01
3608
3610
0.88
0.98
3613
0.94
3640
0.98
3872
1.06
4125
4261
1.03
0.99
4278
0.96
4283
1.04
4365
4374
1.03
0.98
4387
1.02
4406
1.02
4472
0.95
4473
4478
0.89
0.98
4486
0.93
4494
0.95
4550
4551
0.99
1.07
4552
0.95
4564
1.00
4621
1.00
4636
4649
1.04
0.91
5322
0.85
5576
1.01
5813
5831
0.95
1.01
5846
-
5866
5982
0.94
0.72
0.58
0.68
0.60
0.56
0.60
0.72
0.50
0.63
0.61
0.60
0.63
0.58
0.63
0.73
0.73
0.55
0.56
0.71
0.63
0.51
0.59
0.70
0.60
0.45
0.55
0.65
0.39
0.59
0.67
0.63
0.72
0.67
0.71
0.62
0.53
0.64
0.59
0.69
-
0.59
V R3
0.61
-
0.60
0.65
0.54
0.55
0.60
0.57
0.64
0.60
0.55
0.57
0.59
0.62
0.65
0.61
0.60
0.60
0.61
0.64
0.57
0.57
0.61
0.64
0.55
0.63
0.57
0.58
0.58
0.61
0.61
0.63
0.60
0.62
0.62
0.58
0.65
0.57
0.62
-
0.54
V I3
1.28
-
1.25
1.32
1.14
1.14
1.23
1.13
1.31
1.24
1.07
1.21
1.21
1.22
1.29
1.29
1.21
1.18
1.28
1.30
1.16
1.26
1.29
1.29
1.17
1.26
1.21
1.22
1.23
1.29
1.19
1.27
1.29
1.29
1.21
1.17
1.33
1.25
1.28
-
1.18
U V3
1.76
1.60
1.72
1.55
1.54
1.57
1.76
1.40
1.67
1.62
1.47
1.61
1.52
1.61
1.79
1.76
1.54
1.53
1.74
1.66
1.49
1.61
1.72
1.55
1.33
1.53
1.57
1.34
1.58
1.74
1.58
1.72
1.66
1.75
1.54
1.38
1.65
1.54
1.71
-
1.53
U V1.5
1.60
1.72
1.72
1.63
1.41
1.44
1.65
1.43
1.69
1.62
1.44
1.63
1.52
1.64
1.64
1.83
1.53
1.62
1.71
1.60
1.46
1.65
1.78
1.57
1.48
1.84
1.50
1.54
1.51
1.73
1.63
1.74
1.79
1.81
1.53
1.43
1.55
1.50
1.77
1.60
1.53
V R1.5
0.55
0.00
0.60
0.58
0.53
0.54
0.60
0.53
0.56
0.56
0.50
0.53
0.56
0.58
0.63
0.60
0.64
0.57
0.64
0.59
0.59
0.58
0.60
0.62
0.55
0.63
0.54
0.62
0.54
0.59
0.57
0.59
0.61
0.62
0.53
0.53
0.61
0.55
0.61
0.61
0.58
Table 11. "External colours" measured in the outer-
most range of the data. Successive columns: µV V surface
brightness of measurement, a range of 0.5 magnitude cen-
tered at the quoted µV beeing used. Columns (2), (3), (4)
and (5): U−B, B−V, V−R, V−I respectively with esti-
mated errors. These are the same in V−R as in B−V.
NGC µV
2768
2974
3115
3193
3377
3377
3379
3605
3607
3608
3610
3613
3640
3872
4125
4278
4283
4365
4374
4387
4406
4472
4473
4478
4486
4494
4550
4551
4552
4564
4621
4636
4649
5322
5576
5813
5831
5846
5866
5982
24.25
id
id
23.5
24.5
24.25
24.0
23.5
24.25
23.75
24.25
24.5
23.5
24.5
23.5
23.5
23.75
24.
23.25
24.5
23.25
23.
24.5
24.
23.5
24.5
24.0
24.5
24.
24.5
24.
23.5
23.5
24.25
23.75
24.
23.
24.
24.5
24.25
U−B
0.455±.06
0.335±.06
0.365±.05
0.370±.04
0.210±.06
0.175±.05
0.480±.04
0.340±.03
0.370±.07
0.300±.05
0.365±.05
0.395±.07
0.360±.04
0.420±.07
0.405±.03
0.350±.04
0.300±.05
0.470±.04
0.420±.03
0.350±.08
0.330±.04
0.490±.03
0.315±.08
0.200±.04
0.415±.03
0.390±.08
0.120±.06
0.325±.07
0.390±.06
0.200±.07
0.390±.06
0.370±.03
0.525±.04
0.315±.07
0.230±.05
0.410±.06
0.365±.06
0.420±.06
0.270±.08
0.365±.07
B−V
0.851±.02
0.835±.03
0.855±.02
0.880±.02
0.815±.02
0.795±.02
0.890±.02
0.795±.02
0.845±.03
0.900±.02
0.730±.02
0.830±.02
0.875±.02
0.890±.03
0.830±.01
0.800±.02
0.805±.02
0.900±.02
0.895±.01
0.860±.03
0.870±.02
0.930±.01
0.840±.03
0.750±.02
0.885±.02
0.830±.03
0.700±.02
0.850±.03
0.880±.02
0.820±.02
0.370±.02
0.780±.02
0.950±.02
0.785±.03
0.770±.02
0.910±.02
0.840±.02
0.940±.02
0.775±.03
0.820±.03
V−R V−I
0.530
0.000
0.580
0.000
0.505
0.480
0.575
0.510
0.530
0.525
0.510
0.500
0.540
0.555
0.550
0.530
0.510
0.585
0.545
0.550
0.565
0.590
0.575
0.490
0.590
0.500
0.470
0.520
0.515
0.530
0.565
0.510
0.540
0.520
0.525
0.575
0.495
0.580
0.505
0.500
1.110±.04
0.000±.00
1.180±.03
0.000±.00
1.000±.03
0.950±.03
1.160±.03
1.060±.02
1.130±.04
1.165±.03
0.985±.04
1.130±.04
1.135±.03
1.110±.05
1.120±.02
1.055±.03
1.080±.03
1.180±.03
1.180±.03
1.140±.05
1.160±.03
1.255±.02
1.130±.05
1.090±.03
1.200±.03
1.110±.05
1.010±.03
1.120±.04
1.120±.03
1.020±.03
1.175±.03
1.140±.03
1.200±.03
1.110±.04
1.085±.03
1.175±.03
1.115±.04
1.225±.03
1.075±.05
1.100±.04
some indications about stellar populations at the outskirts
of E-galaxies.
On the other hand, the "red halo" effect of the camera
was found to give enormous errors in V−I colours and gra-
dients. These were corrected by a rigorous technique, and
results in agreement with "classical" data were obtained.
Considering the V−I gradients, one is not happy however
to introduce in their evaluation, corrections larger than
the quantity to be measured! Besides this specific prob-
lem of the red halo of thinned CCD, the far wings of the
PSFs have been proven in a recent paper (RM01) to have
non negligible effects in the gradients of other colours, and
also to vary with the age of mirror coatings. It is not im-
possible that the U−B or U−V gradients given here are
overestimated by 15-20%, although they are in excellent
statistical agreement with the well-known work of Peletier
et al. (1990).
16
T.P. Idiart, et al.: New UBVRI colour distributions in E-type galaxies
Franx, M., Illingworth, G., Heckman, T., 1989, AJ 98, 538
(FIH)
Goudfrooij, P., Hansen, L., Jorgensen, H.E., et al., 1994, A&AS
104, 179
Michard, R., Marchal, J., 1994, A&AS 105, 481 (MM94)
Michard, R., 1999, A&AS 137, 245 (RM99)
Michard, R., 2000, A&A 360, 85 (RM00)
Michard, R., 2001, A&A submitted (RM01)
Peletier, R.F., Davies, R.L., Illingworth, G.D., et al., 1990, AJ
100, 1091
Poulain, P., 1988, A&AS 72, 215
Poulain, P., Nieto, J.L. 1994, A&AS 103, 573
Prugniel, P., H´eraudeau, P., 1998, A&ASS 128, 299
Rubin, V.C., Graham, J.A., Kenney, J.D.P., 1992, ApJ 394, L9
Rix, H.W., Franx, M., Fisher, D., 1992, ApJ 400, L5
Trager S.C., Faber S.M., Worthey G. et al., 2000, ApJ 119,
1645
Vigroux, L., Souviron, J., Lachieze-Rey, M., et al., 1988, A&AS
73, 1
Wise, M.W., Silva, D.R. 1996, ApJ 461,155
Witt, A., Thronson, H.A.jr, Capuano, J.M., 1992, ApJ 393,
611
Worthey, G., 1994, ApJS 95, 107
Various colour gradients against log r for a given ob-
ject are well correlated, generally better than in previous
work (see statistics in Sect. 4.1), which is interpreted as
due to smaller measuring errors, notably in U−B. These
improvements in accuracy did not bring out any obvious
correlation between gradients and other galaxy properties.
A few galaxies have exceptionally steep colour gradients
(nearly at 2σ) without sharing other properties.
Colour-colour relations can be built from the present
data for several locations in galaxies, such as near cen-
ter, various fractions of the effective radius re, or the
"outermost" measured range around µV = 24. All these
diagrams overlap to form a single stripe with moderate
scatter (except for one rather obvious calibration error?).
These might prove useful to test theories of old stellar
populations and of their host galaxies. Colour-colour di-
agrams based upon integrated colours have already been
used for this purpose (Worthey, 1994).
The U−B or U−V colours correlate very well with the
M g2 index, both near the galaxy center and at the effec-
tive radius re. This seems to rule out any large influence
of diffuse dust in the colours and colour gradients in E-
galaxies. This was considered likely by Witt et al. (1992)
and discussed by Wise and Silva (1996) with inconclusive
results. Previous arguments against such an influence were
presented in RM00: they were based upon the relative av-
erage values of the gradients in various colours, and are
reinforced in the present work, since the mean gradients
are nearly unchanged, and their errors lessened.
It is well known that, for single-burst stellar popula-
tions, colours and line indices depend both on the metallic-
ity and on the age of the system (Worthey 1994; Borges et
al. 1995). However, E-galaxies are constituted by a popu-
lation mix, having age and metallicity distributions which
reflect their star formation histories. Therefore a colour-
metallicity calibration requires the use of models able to
provide those distributions and, consequently, the inte-
grated colours along the galaxy lifetime. Such a calibration
will be presented in Paper II.
Acknowledgements. TPI acknowledges a Fapesp pos-doc fel-
lowship n◦ 97/13083-7.
References
Bender, R., Mollenhof, C., 1987, A&A 177, 71
Bender, R., Burstein, D., Faber, S.M., 1993, ApJ 411, 153
Borges A.C., Idiart, T.P., de Freitas Pacheco J.A. et al., 1995,
AJ 110, 2408
Burstein, D., Davies, R.C, Dressler, A. et al., 1988, in Towards
Understanding Galaxies at Large Redshift, Ed. R.G. Kron
& A. Renzini, Kluwer, Dordrecht.
Carter, D., 1978, MNRAS 182, 797
de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H.G.Jr. et
al., 1991, Third Reference Catalogue of Bright Galaxies,
Springer, New York
Faber, S.M., Wegner, G., Burstein, D. et al., 1989, ApJSS 69,
763
|
astro-ph/0002389 | 1 | 0002 | 2000-02-21T13:34:21 | Formation of intermediate-mass black holes in circumnuclear regions of galaxies | [
"astro-ph"
] | Recent high-resolution X-ray imaging studies have discovered possible candidates of intermediate-mass black holes with masses of $M_\bullet \sim 10^{2-4} \MO$ in circumnuclear regions of many (disk) galaxies. It is known that a large number of massive stars are formed in a circumnuclear giant H {\sc ii} region. Therefore, we propose that continual merger of compact remnants left from these massive stars is responsible for the formation of such an intermediate-mass black hole within a timescale of $\sim 10^9$ years. A necessary condition is that several hundreds of massive stars are formed in a compact region with a radius of a few pc. | astro-ph | astro-ph | PASJ: Publ. Astron. Soc. Japan , -- ?? (2018)
Formation of Intermediate-Mass Black Holes in Circumnuclear
0
0
0
2
b
e
F
1
2
1
v
9
8
3
2
0
0
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Regions of Galaxies
Yoshiaki Taniguchi, Yasuhiro Shioya
Astronomical Institute, Graduate School of Science, Tohoku University,
Aramaki, Aoba, Sendai 980-8578
E-mail (YT): [email protected]
Takeshi G. Tsuru
Physics Department, Graduate School of Science, Kyoto University,
Kitashirakawa, Sakyo, Kyoto 606-8502
and
Satoru Ikeuchi
Physics Department, Graduate School of Science, Nagoya University,
Furo, Chikusa, Nagoya 464-8602
(Received ; accepted )
Abstract
Recent high-resolution X-ray imaging studies have discovered possible candidates of intermediate-mass
black holes with masses of M• ∼ 102−4M•⊙ in circumnuclear regions of many (disk) galaxies. It is known
that a large number of massive stars are formed in a circumnuclear giant H ii region. Therefore, we propose
that continual merger of compact remnants left from these massive stars is responsible for the formation
of such an intermediate-mass black hole within a timescale of ∼ 109 years. A necessary condition is that
several hundreds of massive stars are formed in a compact region with a radius of a few pc.
Key words: black holes -- galaxies: starburst -- galaxies: star clusters -- X-rays: sources
2
1.
Introduction
Y. Taniguchi et al.
[Vol. ,
Recent sensitive, high-resolution X-ray imaging studies have found unusually bright, compact X-ray sources in
circumnuclear regions of many (disk) galaxies [Colbert & Mushotzky 1999 (hereafter CM99), Okada et al. 1998;
Ptak & Griffiths 1999; Matsumoto & Tsuru 1999; Makishima et al. 2000]. Prior to these observations, it has been
known that some nearby galaxies have luminous X-ray sources in their circumnuclear regions (Fabbiano & Trinchieri
1987; Kohmura et al. 1994; Petre et al. 1994; Takano et al. 1994; Colbert et al. 1995; Reynolds et al. 1997; see for a
review Fabbiano 1988). The observational properties of such circumnuclear X-ray sources are summarized as follows
(e.g., CM 99); 1) ROSAT X-ray luminosities, LX(0.2 -- 2.4 keV) ∼ 1037 -- 1040 ergs s−1 with a mean of 3 × 1039 ergs
s−1; note that some of them are also detected in the hard X ray and thus the total X-ray luminosities are higher by
about one order of magnitude than the above values, 2) the mean displacement between the location of the compact
X-ray source and the optical photometric center of the galaxy is ∼ 390 pc, and 3) they are common; the detection
rate of such compact X-ray sources is >∼ 50%.
Possible origins of these sources are; 1) black hole X-ray binaries, 2) low-luminosity AGN, 3) young X-ray luminous
supernovae, or 4) a new X-ray population (e.g., CM99). Since the Eddington luminosity of a 1.4 M•⊙ accreting neutron
star is 1038.3 ergs s−1, the luminous compact X-ray sources cannot be explained by a single accreting neutron star.
If they are black hole X-ray binaries, their masses are estimated to be M• ∼ 102 -- 104M•⊙; i.e., intermediate-mass
black holes (hereafter IMBHs). This urges us to consider how such IMBHs can be formed in circumnuclear regions
of galaxies.
One of ideas may be that they are formed through the collapse of massive clouds (see Rees 1978). In typical
disk galaxies, gas clouds are present as a form of giant molecular gas clouds; hereafter GMCs. A typical GMC
has the following characteristics (e.g., Blitz 1993); 1) the mass, MGMC ≃ (1 -- 2) ×105M•⊙, 2) the mean diameter,
DGMC ≃ 45 pc, 3) the volume, VGMC ≃ 9.6 × 104 pc3, 4) the average density of H2, nH2 ≃ 50 cm−3, and 5) the
mean column density of H2, N H2 ≃ (3 -- 6) ×1021 cm−2. Although the typical kinetic temperature is Tkin ∼ 10 K,
the sound speed, cs, is generally higher than the velocity dispersion derived from the kinetic temperature because
the turbulence plays an important role in GMCs; therefore, cs ∼ 1 km s−1. If a GMC is perturbed gravitationally
by some physical mechanism, it is likely that it begins to experience fragmentation, resulting in the formation of
sub-giant molecular clouds. Such fragmentation could occur preferentially in denser parts of the GMC. Since a
No. ]
Formation of Sub-Supermassive Black Holes
3
typical H2 density and a typical kinetic temperature in such dense parts are nH2 ∼ 104 cm−3 and Tkin ∼ 10 K,
respectively, we obtain a Jeans' mass of mJ ∼ λ3
JρH2 ∼ 20M•⊙ for λJ = cs(π/GρH2)1/2 ∼ 0.36 (cs/0.3 km s−1) pc and
ρH2 = nH2mH2 where mH2 is the mass of a hydrogen molecule (Jeans 1929). Therefore, IMBHs may not be formed
directly from such massive gas clouds.
Instead, in this Letter, we propose that an IMBH is formed by continual merging of compact remnants (i.e.,
neutron stars and black holes) left from massive stars in a giant H ii region. The reason for this is that giant H
ii regions are often observed in circumnuclear regions of disk galaxies (e.g., Kennicutt, Keel, & Blaha 1989 and
references therein; see for the formation mechanism of such circumnuclear giant H ii regions, Elmegreen 1994). We
also discuss some other proposed mechanisms responsible for the formation of IMBHs.
2. Proposed Model
2.1. Dynamical Relaxation of Massive-Star Compact Remnants
It has been considered that one plausible mechanism for the formation of supermassive black holes (SMBHs) in
galactic nuclei is to pile up compact remnants of massive stars born in nuclear regions (e.g., Spitzer & Saslaw 1966;
Spitzer & Stone 1967, Spitzer 1969; Saslaw 1973; Begelman & Rees 1978; Weedman 1983; Norman & Scoville 1988;
Quinlan & Shapiro 1989, 1990; Lee 1995; Taniguchi, Ikeuchi, & Shioya 1999). On the analogy of this mechanism,
we investigate the dynamical evolution of compact remnants left from massive stars born in the core of a giant H ii
region in circumnuclear regions of galaxies.
Typical masses of IMBHs lie in the range between ∼ 102M•⊙ and ∼ 104M•⊙ (CM99). Therefore, we investigate a
possibility that an IMBH with mass of 103M•⊙ can be formed from mergers of compact remnants left in the core of
a giant H ii region. We consider a case that N massive stars with masses higher than 8 M•⊙ are formed within a
sphere with a radius of rcl. Here it is noted that all such massive stars yield a compact remnant after the supernova
explosion while the core of stars with masses less massive than 8 M•⊙ evolves into a white dwarf. It is known that
the mass of a neutron star is ≈ 1.4M•⊙ while that of a stellar-size black hole is from a few M•⊙ to ≈ 10M•⊙; e.g.,
Brown & Bethe (1994) suggested that main-sequence stars with 18M•⊙ < m∗ < 25M•⊙ evolve to low-mass black
holes (m• ≃ 1.5M•⊙) while stars with m∗ > 25M•⊙ evolve to high-mass black holes (m• >∼ 10M•⊙). However, the mass
function for giant H ii regions has not yet been well known. Therefore, in the above estimate, we have assumed for
4
Y. Taniguchi et al.
[Vol. ,
simplicity that all the compact remnants are black holes with a mass of 2 M•⊙ and thus 500 seed compact remnants
are necessary to make an IMBH with a mass of 103M•⊙.
All the massive stars die within a timescale of ∼ 107 years and then a cluster of compact remnants will result in.
The candidates of IMBHs are observed in circumnuclear regions of galaxies (i.e., a typical radial distance is R ≃ a
few 100 pc). At such a distance, the rotation curve of galaxies may be described by the rigid rotation (e.g., Rubin
et al. 1985; cf. Sofue 1997); we assume that the rigid rotation continues up to a radius of R = 1 kpc and reaches
the maximum rotation velocity of vrot = 200 km s−1. If a giant H ii region is formed at R = 100 pc, the cluster of
massive stars (and thus the cluster of compact remnants too) is influenced by the mass contained within R = 100
pc. This mass is estimated to be M ∼ Rv2
rot/G ∼ 9 × 106R3
100M•⊙ where R100 is the radial distance in units of 100
pc; note that vrot = 20R100 km s−1. The radius of the remnant cluster can be estimated as the tidal radius,
rcl ∼ rtidal ∼ R[N m•/(2M )]1/3 ∼ 3.8N 1/3
500 m1/3
•,2 pc
(1)
where N500 is the number of compact remnants in units of 500 and m•,2 is the mass of a compact remnant in units
of 2 M•⊙. In this case, the cluster can be relaxed dynamically with a timescale of
τdyn ∼ N 1/2G−1/2r3/2
cl m−1/2
•
∼ 1.8 × 109 y.
(2)
Since this timescale is shorter than the age of galaxies, this mechanism may be responsible for the formation of
IMBHs in circumnuclear regions of galaxies.
It is noted that the dynamical situation for a cluster of compact remnants considered here is completely different
from both that for a cluster of compact remnants in globular clusters (e.g., Kulkarni, Hut, & McMillan 1993) and
that for a cluster of compact remnants in the nuclear region of galaxies (e.g., Weedman 1983; Lee 1995 and references
therein) because low-mass stars dominate the cluster gravitational potential for the latter two cases. On the other
hand, it is expected that massive stars may be preferentially formed in the core. Since supernovae could blow the gas
resided originally in the core region, compact remnants themselves may dominate the cluster gravitational potential.
Furthermore, lifetimes of OB stars are less than ∼ 107 years, being much shorter than the dynamical relaxation
timescale estimated above. Therefore, the star cluster evolves to a cluster of compact remnants quickly. Some black
holes may be ejected from the cluster by the dynamical effect during the formation of black hole binaries through
the three-body encounters (Spitzer 1987; Kulkarni et al. 1993). However, this effect is negligible if N >∼ 100 (e.g.,
Binney & Tremaine 1987); this is indeed the case discussed here.
No. ]
Formation of Sub-Supermassive Black Holes
5
Then we investigate what kind of massive star formation is necessary. We estimate an average number density
of gas, nH = Mgas/[(4π/3)r3
clmH] ≃ 1 × 104 cm−3 where Mgas = N m∗η−1
SF,0.1 = 5 × 104M•⊙ given m∗ = 10M•⊙ and
ηSF,0.1 is the star formation efficiency in units of 0.1; i.e., 10% of the gas in the cloud is used to form massive stars.
This density is comparable to those in typical dense cores in star forming regions (e.g., nH ∼ 104 cm−3; e.g., Lada,
Strom, & Myers 1993). Note that the gas mass lies in the observed range of MGMC. The mass volume density of
massive stars is estimated as N m∗/[(4π/3)r3
cl] ≃ 22M•⊙ pc−3. This is lower by two orders of magnitude than that
of the compact star cluster R136a, whose half light radius is 1.7 pc, embedded in the central region of the 30 Dor
nebula in Large Magellanic cloud; 5.5 × 104M•⊙ pc−3 (Hunter et al. 1995). Therefore, the star formation properties
as well as the star cluster properties in our model are not unusual.
Finally we mention which type of galaxies tends to have more massive IMBHs. Using equations (1) and (2)
together with the relation M ∼ Rv2
rot/G, we obtain
τdyn ∼
N
2
R
vrot
∼
N
2
τrot
2π
,
(3)
where τrot is the rotation period. If the rigid rotation is achieved up to a radius of R = 1 kpc as observed, the
dynamical timescale is independent of R if R ≤ 1 kpc; i.e., any cluster of compact remnants left from the core of
a giant H ii region within R ≤ 1 kpc could evolve to an IMBH. Since any disk galaxy have been able to form such
IMBHs during its lifetime, i.e., ∼ 1010 years, the total number of IMBHs formed in its lifetime is estimated to be
N ∼ 4π ×(cid:18) τdyn
τrot (cid:19) ∼ 4 × 103(cid:18) τdyn
1010 y(cid:19)(cid:18)
τrot
3.1 × 107 y(cid:19)−1
.
(4)
More importantly, the relation given in equation (3) implies that galaxies with slower rigid-rotation velocities (i.e.,
late-type spirals) tend to have less massive IMBHs while those with higher rigid-rotation velocities (i.e., early-type
spirals) tend to have more massive IMBHs. Since the survey conducted by CM99 is biased to late-type spirals, we
are unable to make this observational test. X-ray surveys will be recommended for a sample of early-type spirals to
test whether or not this prediction is consistent with observation.
2.2.
Intermediate-mass Black Holes Supplied from Satellite Galaxies
We have shown that an IMBH can be made through the merger of compact remnants of massive stars born in
a giant H ii region. Since any galaxies have satellite galaxies (e.g., Zaritsky et al. 1997) and giant H ii regions
could be made in some gas-rich satellites (e.g., 30 Dor in LMC), it seems important to investigate a possibility that
6
Y. Taniguchi et al.
[Vol. ,
IMBHs can be supplied by minor mergers with satellite galaxies having IMBHs. M 32, one of the satellite galaxies
of M 31, may have a supermassive black hole with a mass of ∼ 106M•⊙ (Dressler & Richstone 1988; see for a review
Kormendy et al. 1998). At present, there is no observational evidence for the presence of IMBHs in any satellite
galaxies. However, if some satellite galaxies had IMBHs in their nuclei, it could be possible that minor mergers with
such satellites are responsible for the presence of IMBHs in circumnuclear regions of the host galaxies.
We estimate the frequency of occurrence of compact X-ray sources if all of them are attributed to the IMBHs
supplied by minor mergers with nucleated satellite galaxies (see Taniguchi & Wada 1996). Tremaine (1981) estimated
that every galaxy would experience minor mergers with its satellite galaxies several times. Since a typical galaxy may
have several satellite galaxies, the probability of merger for a satellite galaxy may be estimated to be fmerger ≃ 0.5;
i.e., half of the satellite galaxies have already merged to a host galaxy, while the rest are still orbiting. Another
important value is the number of nucleated satellite galaxies. For example, M31 has two nucleated satellites (M32
and NGC 205), and a field S0 galaxy NGC 3115 has a nucleated dwarf (van den Bergh 1986). Although there has
been no systematic search for nucleated satellite galaxies, it is likely that every galaxy has (or had) a few nucleated
satellites. Therefore, for simplicity we assume nsat = 2. If we assume that the typical lifetime of the X-ray sources
is τactive ≃ 108 years, we obtain an expected frequency,
PIMBH ≃ fmerger nsat τactive τ −1
Hubble ∼ 0.01nsat,2τactive,8,
(5)
where τHubble is the Hubble time, ∼ 1010 years, nsat,2 is the number of nucleated satellites in units of 2, τactive,8 is the
duration of the active phase in units of 108 years. Hence, if minor mergers with nucleated satellites are responsible
for the IMBHs in the circumnuclear regions of their host galaxies, it is statistically expected that hard X-ray sources
are found in about 1 % of field disk galaxies. This is significantly smaller than the observed frequency, >∼ 50%
(CM99).
However, it seems possible that IMBHs can be formed through the dynamical relaxation of cores of giant H ii
regions in gas-rich satellite galaxies and they travel to circumnuclear regions of their host galaxies by minor mergers.
This case gives another expected frequency,
PIMBH ≃ fmerger nsat nIMBH τactive τ −1
Hubble
(6)
where nIMBH is the number of IMBH in each satellite galaxy. Since the maximum number of IMBHs can be estimated
No. ]
Formation of Sub-Supermassive Black Holes
as nIMBH ∼ τHubble/τdyn, we obtain
PIMBH ≃ fmerger nsat τactive τ −1
dyn ∼ 0.1nsat,2τactive,8τ −1
dyn,9
7
(7)
where τdyn,9 is the dynamical timescale in units of 109 years. Note that nsat is not the number of nucleated satellites
but that of gas-rich satellites. However, for simplicity, we adopt nsat = 2 even in this estimate.
In summary, the frequency of occurrence of IMBHs supplied from satellite galaxies lies in a range between 0.01
and 0.1. Therefore, we suggest that minor mergers may not explain the observed higher frequency of IMBHs unless
nsat > 10.
3. Alternative Mechanisms
We have shown that continual merging of compact remnants left from massive stars in a giant H ii region formed
in circumnuclear regions of galaxies may be responsible for the formation of an IMBH intermediate-mass black hole
within a timescale of ∼ 109 years. A necessary condition is that several hundreds of massive stars are formed in a
compact region with a radius of a few pc. In this section, we discuss some alternative mechanisms responsible for
the formation of IMBHs.
3.1. Bondi-type gas accretion onto a seed black hole
Compact remnants left from massive stars formed in the initial star formation in a galaxy have been surviving for
the galaxy age, i.e., τage ∼ 1010 years. They have been experiencing a number of encounters with gas clouds in the
galaxy. Therefore, classical Bondi-type (Bondi 1952) gas accretion is the most probable accretion process for them
(Yoshii 1981). This gas accretion rate is given by MBondi = 2 π mH nH r2
a ve, where mH, nH, ra, and ve are the
mass of a hydrogen atom, the number density of the hydrogen atom, the accretion radius defined as ra = GM•v−2
e
(M• is the mass of the seed compact remnant), and the effective relative velocity between the seed black hole and
the ambient gas, respectively.
Since the Bondi-type gas accretion is most important for the low-velocity encounter, we adopt ve = 1 km s−1 in
order to estimate the maximum accretion rate (Yoshii 1981) although ve is realistically much higher than this value.
Adopting an average gas density of the interstellar medium, nH ∼ 1 cm−3, we obtain the total accreting mass due
8
Y. Taniguchi et al.
to the Bondi-type accretion,
Macc,Bondi = MBondi τage ∼ 3 × 10−2M 2
•,1nH,1v−3
e,1 τage,10 M•⊙
[Vol. ,
(8)
where M•,1 is the mass of the seed compact remnant in units of 1M•⊙, nH,1 is the average gas density in units of 1
cm−3, ve,1 is the orbital velocity with respect to the ambient gas in units of 1 km s−1, and τage,10 is the age of the
galaxy in units of 1010 years. Even if a black hole with a mass of 10 M•⊙ was born 1010 years ago in the galaxy (see
for the formation of black holes with m• ≃ 10M•⊙, Brown & Bethe 1994), its accreting mass amounts only to a few
M•⊙. It is expected that some black holes might encounter dense molecular clouds. However, the crossing time is as
short as τcross ∼ DGMC,45/ve,1 ∼ 4.5 × 107 years where DGMC,45 is the mean diameter of GMCs in units of 45 pc.
It is also noted that the age of a GMC may range between ∼ 105 years (the molecule formation time) and several
times 108 years (e.g., Elmegreen 1985). Therefore, in the case of an encounter between a black hole and a GMC, we
should adopt τage ∼ 108 years in equation (8). Therefore, the mass growth is negligibly small even if nH ∼ 50 cm−3.
However, if ve ≪ 1 km s−1, an IMBH could be formed through this accretion process. Since it seems unlikely that
such a very slow encounter occurs frequently, it is suggested that the Bondi-type gas accretion onto the seed black
hole may not be responsible for the formation of an IMBH within the age of the galaxy.
3.2. Disk-type gas accretion onto a seed black hole
Next we consider a case that a compact remnant is a black hole with a mass of 1 M• and the disk-type gas
accretion has been occurring at the Eddington accretion rate,
MDisk ≃ 2.2 × 10−8ηacc,0.1M•,1 M•⊙ y−1 where ηacc is
the conversion efficiency from the gravitational energy to the radiation in units of 0.1 (e.g., Rees 1984). The seed
black hole with a mass of 1 M•⊙ can increase in mass and its mass is given by
M•(t8) = M•,1e2.2ηacc
,
.
0
1t8 M•⊙
(9)
where t8 is the duration of the gas accretion in units of 108 years. One obtains M• = 102M•⊙ at t8 = 2.09,
M• = 103M•⊙ at t8 = 3.14, and M• = 104M•⊙ at t8 = 4.19. Therefore, it seems that IMBHs could be easily formed
by this mechanism.
However, there is a serious problem in this model. The star formation rate may be higher in earlier phases of
galaxy evolution. Even if we assume that massive stars have been made continuously at a constant rate, the number
of compact remnants made during the last 109 years is estimated to be ∼ 107 for a typical galaxy with mass of
No. ]
Formation of Sub-Supermassive Black Holes
9
1012M•⊙ in which stars were formed with the Salpeter mass function (see Taniguchi et al. 1999). If all these compact
remnants have been experiencing the disk accretion at the Eddington rate, we would observe a lot of IMBHs in the
galaxy if they have not yet merged into one SMBH. If they were already merged into one SMBH, there would be
a very supermassive black hole with mass of M• ∼ 107 × 104M•⊙ ∼ 1011M•⊙ in some galaxies. Therefore, the idea
described here may not be a dominant mechanism responsible for the formation of IMBHs.
We would like to thank an anonymous referee for his/her many useful comments and suggestions which improved
this paper significantly. This work was supported in part by the Ministry of Education, Science, Sports and Culture
in Japan under Grant Nos. 10044052, and 10304013.
10
References
Y. Taniguchi et al.
[Vol. ,
Blitz L. 1993, in Protostars and Planets III, ed E.H. Levy, J.I. Lunine (The University of Arizona Press, Tucson),
p125
Begelman M.C., Rees M.J. 1978, MNRAS, 185, 847
Binney J., Tremaine S. 1987, Galactic Dynamics (Princeton University Press, Princeton), 492
Bondi H. 1952, MNRAS, 112, 195
Brown G.E., Bethe H.A. 1994, ApJ, 423, 659
Colbert E.J.M., Mushotzky R.F. 1999, ApJ, 519, 89
Colbert E.J.M., Petre R., Schlegel E.M., Ryder S.D. 1995, ApJ, 446, 177
Dressler A., Richstone D.O. 1988, ApJ, 324, 701
Elmegreen B.G. 1985, in Protostars and Planets, edited by D. C. Black, and M. S. Matthews (The University of
Arizona Press), 33
Elmegreen B.G. 1994, ApJ, 425, L73
Fabbiano G. 1988, ARA&A, 27, 87
Fabbiano G., Trinchieri G. 1987, ApJ, 315, 46
Hunter D.A., Shaya E.J., Holtzman J.A., Light R.M., O'Neil E.J., Lynds R. 1995, ApJ, 448, 179
Jeans J.H. 1929, Astronomy and Cosmogony, 2nd ed. Cambridge, Eng.: (Cambridge University Press, Cambridge)
Kennicutt R.C. Jr, Keel W.C., Blaha C.A. 1989, AJ, 97, 1022
Kohmura Y., et al. 1994, PASJ, 46, L157
Kormendy J., Bender R., Evans A.S., Richstone D. 1998, AJ, 115, 1823
Kulkarni S.R., Hut P., McMillan S. 1993, Nature, 364, 421
Lada E.A., Strom K.M., Myers P.C. 1993, in Protostars and Planets III, ed E.H. Levy, J.I. Lunine (The University
of Arizona Press, Tucson), p245
Lee H.M. 1995, MNRAS, 272, 605
Makishima K., et al. 2000, ApJ, submitted (astro-ph/0001009)
Matsumoto H., Tsuru T.G. 1999, PASJ, 51, 321
Mushotzky R. F., Done C., Pounds K. A. 1993, ARA&A, 31, 717
No. ]
Formation of Sub-Supermassive Black Holes
11
Norman C., Scoville, N. 1988, ApJ, 332, 124
Okada K., Dotani T., Makishima K., Mitsuda K., Mihara T. 1998, PASJ, 50, 25
Petre R., Okada K., Mihara T., Makishima K., Colbert E.J.M. 1994, PASJ, 46, L115
Ptak A., Griffiths R. 1999, ApJ, 517, L85
Rees M.J. 1978, Observatory, 98, 210
Rees M.J. 1984, ARA&A, 22, 471
Reynolds C.S., Loan A.J., Fabian A.C., Makishima K., Brandt W.N., Mizuno T. 1997, MNRAS, 286, 349
Rubin V.C., Burstein D., Ford W.K. Jr, Thonnard N. 1985, ApJ, 289, 81
Quinlan G.D., Shapiro S.L. 1989, ApJ, 343, 725
Quinlan G.D., Shapiro S.L. 1990, ApJ, 356, 483
Salpeter, E.E. 1955, ApJ, 121, 161
Saslaw W.C. 1973, PASP, 85, 5
Sofue Y. 1997, PASJ, 49 17
Spitzer L. 1969, ApJ, 158, L139
Spitzer L. 1987, Dynamical Evolution of Globular Clusters (Princeton University Press, Princeton), ch3
Spitzer L., Saslaw W.C. 1966, ApJ, 143, 400
Spitzer L., Stone, M.E. 1967, ApJ, 147, 519
Stevens I.R., Strickland D.K., Wills K.A. MNRAS, 308, L23
Takano M., Mitsuda K., Fukazawa Y., Nagase F. 1994, ApJ, 436, L47
Taniguchi Y., Ikeuchi S., Shioya Y. 1999, ApJ, 514, L9
Taniguchi Y., Wada K. 1996, ApJ, 469, 581
Tremaine S. 1981, in The Structure and Evolution of Normal Galaxies, ed S.M. Fall, D. Lynden-Bell (Cambridge
University Press, Cambridge), p67
van den Bergh S. 1986, AJ, 91, 271
Weedman D.W. 1983, ApJ, 266, 479
Yoshii Y. 1981, A&A, 97, 280
Zaritsky D., Smith R., Frenk C., White S.D.M. 1997, ApJ, 478, 39
|
astro-ph/9609022 | 1 | 9609 | 1996-09-03T20:28:48 | The secondary infall model of galactic halo formation and the spectrum of cold dark matter particles on Earth | [
"astro-ph",
"hep-ph"
] | The spectrum of cold dark matter particles on Earth is expected to have peaks in velocity space associated with particles which are falling onto the Galaxy for the first time and with particles which have fallen in and out of the Galaxy only a small number of times in the past. We obtain estimates for the velocity magnitudes and the local densities of the particles in these peaks. To this end we use the secondary infall model of galactic halo formation which we have generalized to take account of the angular momentum of the dark matter particles. The new model is still spherically symmetric and it admits self-similar solutions. In the absence of angular momentum, the model produces flat rotation curves for a large range of values of a parameter $\epsilon$ which is related to the spectrum of primordial density perturbations. We find that the presence of angular momentum produces an effective core radius, i.e. it makes the contribution of the halo to the rotation curve go to zero at zero radius. The model provides a detailed description of the large scale properties of galactic halos including their density profiles, their extent and total mass. We obtain predictions for the kinetic energies of the particles in the velocity peaks and estimates for their local densities as functions of the amount of angular momentum, the age of the universe and $\epsilon$. | astro-ph | astro-ph |
IFT-HET-96-20,OSU-TA-20/96
The secondary infall model of galactic halo formation and the
spectrum of cold dark matter particles on Earth.
P. Sikivie1 , I. I. Tkachev2,3 and Yun Wang4
1Department of Physics, University of Florida, Gainesvil le, FL 32611
2Department of Physics, The Ohio State University, Columbus, OH 43210
3 Institute for Nuclear Research, Russian Academy of Sciences, Moscow 117312, Russia
4NASA/Fermilab Astrophysics Center, FNAL, Batavia, IL 60510
(September 3, 1996)
Abstract
The spectrum of cold dark matter particles on Earth is expected to have
peaks in velocity space associated with particles which are falling onto the
Galaxy for the first time and with particles which have fallen in and out of
the Galaxy only a small number of times in the past. We obtain estimates
for the velocity magnitudes and the local densities of the particles in these
peaks. To this end we use the secondary infall model of galactic halo formation
which we have generalized to take account of the angular momentum of the
dark matter particles. The new model is still spherically symmetric and it
admits self-similar solutions. In the absence of angular momentum, the model
produces flat rotation curves for a large range of values of a parameter ǫ which
is related to the spectrum of primordial density perturbations. We find that
the presence of angular momentum produces an effective core radius, i.e.
it
makes the contribution of the halo to the rotation curve go to zero at zero
radius. The model provides a detailed description of the large scale properties
of galactic halos including their density profiles, their extent and total mass.
We obtain predictions for the kinetic energies of the particles in the velocity
peaks and estimates for their local densities as functions of the amount of
angular momentum, the age of the universe and ǫ.
PACS numbers: 98.80.Cq, 95.35.+d, 14.80.Mz, 14.80.Ly
Typeset using REVTEX
1
I. INTRODUCTION
Experiments are presently under way which attempt to identify the nature of dark matter
[1] by direct detection on Earth. The dark matter candidates which are being searched for in
this manner are WIMPs and axions. WIMPs is an acronym for “weakly interactive massive
particles”. The best motivated candidate of this type is the lightest supersymmetric partner
in supersymmetric extensions of the standard model of particle physics [2]. The mass range
for which WIMPs provide the critical energy density for closing the universe is a few GeV
to a few hundred GeV. The axion is a light pseudo-scalar particle whose existence has been
postulated to explain why, in the context of the standard model of particle physics, the
strong interactions conserve P and CP [3]. The likely mass range for which axions provide
the critical energy density for closing the universe is 10−4 eV < ma < 10−7 eV [4].
Axions and WIMPs are the leading cold dark matter (CDM) candidates. Other forms
of dark matter are neutrinos and dark baryons. From the point of view of galaxy formation,
the defining properties of CDM are:
1. that CDM particles, unlike baryons, are guaranteed to interact with their surroundings
only through gravity, and
2. that CDM particles, unlike neutrinos, have negligibly small primordial velocity disper-
sion.
Studies of large scale structure formation support the view that the dominant fraction of
dark matter is CDM. Moreover, if some fraction of the dark matter is CDM, it necessarily
contributes to galactic halos by falling into the gravitational wells of galaxies and hence
is susceptible to direct detection on Earth. WIMPs are being searched for by looking for
WIMP + nucleus elastic scattering in a laboratory detector [5]. The nuclear recoil can be put
into evidence by low temperature calorimetry, by ionization detection or by the detection of
ballistic phonons. Axions are being searched for by stimulating their conversion to photons
in a laboratory magnetic field [6,7]. The experimental apparatus involves an electromagnetic
cavity placed in the bore of a superconducting solenoid. When the resonant frequency of
the lowest TM mode of the cavity equals the axion mass (hν = ma c2), some galactic halo
axions convert to microwave photons inside the cavity.
If a signal is found in the cavity
detector of dark matter axions, it will be possible to measure the energy spectrum of the
axions with great precision and resolution. (The energy resolution is limited only by the
measurement integration time and the stability of the local oscillator with which the axion
signal is compared.) Hence the question arises: what can be learned about our galaxy and
the universe by analyzing such a signal? The main purpose of this paper is to address this
question by predicting properties of the CDM spectrum on Earth in terms of cosmological
input parameters. Incidentally, all CDM candidates have the same phase space distribution,
and hence the same spectrum on Earth, in the limit where their small primordial velocity
dispersions are neglected. We are motivated in part by the fact that knowledge of the
spectrum may help in the discovery of a signal.
In many past discussions of dark matter detection on Earth, it has been assumed that
the dark matter particles have an isothermal distribution, or an adiabatic deformation of
an isothermal distribution. A strong argument in favor of this assumption is the fact that
2
it predicts the rotation curve of the galaxy to be flat at large radial distances. Indeed, self-
gravitating isothermal spheres always have density distributions ρ(r) which fall off at large
r as 1/r2 . Moreover, they have a “core radius”, i.e. a radius a within which the density ρ(r)
no longer behaves as 1/r2 but goes to a constant as r → 0. The behaviour may, for most
practical purposes, be approximated by the function ρ(r) = ρ(0)[1 + (r/a)2 ]−1 . Thus, the
contribution of an isothermal halo to the galactic rotation velocity goes to zero as r → 0.
This feature of isothermal halos is attractive as well because it is known that, in spiral
galaxies like our own, the rotation velocity at small radii may be entirely accounted for by
the bulge and the disk. In our galaxy, the core radius is such that roughly half of the rotation
velocity squared at the solar radius rs ≃ 8.5 kpc is due to the disk and bulge while the other
half is due to the dark halo. Thermalization of the galactic halo has been argued to be the
outcome of a period of “violent relaxation” [8] following the collapse of the protogalaxy. If
it is strictly true that the velocity distribution of the dark matter particles is isothermal,
then the only information that can be gained from its observation is the corresponding virial
velocity and our own velocity relative to its standard of rest.
However, one may convince oneself that the velocity distribution of dark matter particles
has a non-thermal component. Consider the fact that our closest neighbor on the galactic
scale, the galaxy M31 in Andromeda, at a distance of order 730 kpc from us, is falling towards
our galaxy with a line-of-sight velocity of 120 km/sec. This motion can be understood to be
due to the mutual gravitational attraction between the two galaxies: first they were receding
from each other as part of the general Hubble flow, this relative motion was halted and now
they are falling towards one another. We may use M31 as an indicator of the motion of any
matter in our neighborhood. Moreover, if cold dark matter exists, then there is cold dark
matter at every physical point in space (including everywhere we see nothing and which
appears empty), because by Liouville’s theorem the 3-dim. sheet in 6-dim. phase-space on
which the CDM particles lie can not be ruptured. The thickness of that sheet is the tiny
primordial velocity dispersion of the CDM particles, of order 10−12 for WIMPs and 10−17 for
axions (c = 1). The implication of the above is that, if CDM exists, there are CDM particles
falling onto our galaxy continuously and from all directions. The motion of these particles
gets randomized by gravitational scattering off giant molecular clouds, globular clusters and
other inhomogeneities but complete thermalization of their velocity distribution occurs only
after they have fallen in and out of the galaxy many times. As a result there are peaks in the
velocity distribution of CDM particles at any physical point in the galaxy [9]. One peak is
due to particles falling onto the Galaxy for the first time, one peak is due to particles falling
out of the Galaxy for the first time, one peak is due to particles falling in for the second
time, and so on. In particular, this is true on Earth. The width of the first two peaks, which
we label n=1, is related to and is of order the velocity dispersion of the particles before they
fall in for the first time. The width of the next two peaks (n=2) is expected to be somewhat
larger as a result of scattering of the particles off inhomogeneities in the galaxy. The width
of the next two peaks (n=3) is larger still because these particles have been scattered more.
And so on.
One of the main purposes of this paper is to obtain estimates of the sizes and velocity
magnitudes of the velocity peaks on Earth. By “size”, we mean the contribution of the peak
to the local mass density of the halo. By “velocity magnitude” , we mean the magnitude of
the velocity vector of the particles in the peak as measured in the rest frame of the Galaxy,
3
in a frame which is not corotating with the disk. The tool we use is the secondary
i.e.
infall model [10] of galactic halo formation. This model assumes a single overdensity in an
expanding universe. A halo forms around the overdensity because dark matter keeps falling
onto it. The dark matter is assumed to have gravitational interactions only and to have
zero initial velocity dispersion. The model also assumes spherical symmetry. Finally, in
its original form, it assumes that the dark matter particles have zero angular momentum
with respect to the center and hence that their motion is purely radial. Much progress
[11,12] in the analysis of the model was made as a result of the realization that the time-
evolution is self-similar provided Ω = 1 and provided the initial overdensity has a special
scale-free form; see Eq.(3.11). The parameter ǫ that appears in this ansatz is related to
the slope of the power spectrum of primordial density perturbations at the galactic scale.
Self-similarity means that the halo is time-independent after all distances are rescaled by
an overall time-dependent size R(t) and all masses by a time-dependent mass M (t). The
self-similar solutions can be obtained numerically with great precision and some of their
properties may be derived analytically. When the parameter ǫ is in the range 0 ≤ ǫ ≤ 2/3,
the density profile ρ(r) ∼ 1/r2 and thus the rotation curve is flat.
However, for the purpose of estimating the sizes of velocity peaks, the secondary infall
model without angular momentum is rather inadequate. In particular, it tends to overesti-
mate the size of the peaks due to dark matter particles falling in and out of the galaxy for
the first time. Indeed, angular momentum has the effect of keeping infalling dark matter
away from the galactic center and this effect is largest for particles falling into the galaxy
last. On the other hand, the presence of angular momentum destroys spherical symmetry
and thus makes the actual evolution far more complicated and untractable. However, as
will be explained in detail below, it is possible to include the effect of angular momentum
into the secondary infall model without destroying its spherical symmetry by averaging over
all possible orientations of an actual physical halo [13]. Moreover, the time evolution of
the model with angular momentum thus included is still self-similar provided the angular
momentum distribution is of a particular scale-free form. It was also found [13] that angular
momentum has the effect of making the halo contribution to the galactic rotation curve go
to zero at the galactic center, thus introducing an effective core radius for the halo mass
distribution. We define the effective core radius b to be the radius at which the halo con-
tributes half of the galaxy’s rotation velocity squared. For our Galaxy, b is of order our own
distance to the Galactic center. This by itself suggests that the effect of angular momentum
on the velocity peaks on Earth is not small. The model with angular momentum can be
accurately solved on a computer. Its predictions for the effective core radius b, the local
halo density, and the expected sizes and velocity magnitudes of the first few velocity peaks
are tabulated below for representative values of the input parameters, which are the age
of the Universe, the parameter ǫ and the average amount of angular momentum. We also
give an analytical treatment of the model under simplifying but realistic assumptions. It
yields general formulae which may be used to estimate the expected sizes and the velocity
magnitudes of the velocity peaks for a wide range of the input parameter values.
In Section II we review the arguments of ref [9] why velocity peaks in the cold dark matter
spectrum on Earth are expected, and add some comments of our own. In Section III we give
a detailed description of the self-similar infall model, without and with angular momentum.
In Section IV we describe how some of the model parameters are determined in terms of
4
observed properties of our Galaxy and we give the results of the numerical integration of the
model. Section V contains our analytical treatment of the model. Section VI summarizes
our results.
II. PHASE SPACE STRUCTURE OF COLD DARK MATTER HALOS.
In cold dark matter scenarios, the initial phase space distribution is a very thin sheet
near ~v = H~r, where H is the Hubble rate and ~r is the position relative to an arbitrarily
chosen reference point. The deviations from perfect Hubble flow which are present are
associated with the primordial density perturbations that will produce galaxies and large
scale structures by gravitational instability. Where a galaxy (or some other ob ject) forms
and grows, the phase space sheet is folding itself up. The process is illustrated in Fig. 1
for the simplified case where a single spherically symmetric overdensity is present in an
otherwise homogeneous universe and where all dark matter particles move on radial orbits
through the center of the overdensity. The line in the figure indicates the location of the
dark matter particles in (r, r) phase space at an instant of time. r is the distance to center
of the overdensity and r = dr/dt. As time goes on the line “winds up” in the clockwise
direction, rotating most rapidly at the center.
Fig. 1 shows that the velocity spectrum of cold dark matter particles on Earth, or
anywhere else in the galaxy, has a series of peaks. One peak is due to particles falling onto
the galaxy for the first time, passing by Earth while going towards the galactic center. A
second peak is due to particles which are falling out of the galaxy for the first time, passing
by Earth while going away from the galactic center. A third peak is due to particles falling
onto the galaxy for the second time. A fourth peak is due to particles falling out of the
galaxy for the second time. And so on. A rough estimate of the number N of velocity peaks
on Earth in this idealized case may be obtained as the ratio of the age of the galaxy (∼ 1010
years) to the time (∼ 0.5 × 108 years) it takes a particle to fall to the center of the galaxy
starting from rest at the Earth’s location, with the result N ∼ 200. However, the presence
of angular momentum of the dark matter particles tends to decrease N by restricting the
range of radii over which dark matter orbits vary. (In the extreme limit of circular orbits,
N = 1.) As will be seen below, this expectation [9] is confirmed by our calculations. We
shall also find that the number of peaks depends upon ǫ.
Of course the description of a galactic halo presented in Fig. 1 is much simplified. In the
remainder of this section, we discuss the sensitivity of the conclusion, that there are peaks
in the cold dark matter velocity distribution on Earth, to the simplifying assumptions that
were made. In particular, we inquire into the effect of
1. the gravitational scattering of the dark matter particles by inhomogeneities in the
galaxy
2. the angular momentum that the dark matter particles have with respect to the galactic
center
3. the velocity dispersion that the dark matter particles have before they fall onto the
galaxy.
5
A. Scattering by inhomogeneities in the galaxy.
The effect of the gravitational scattering of the dark matter particles by the inhomo-
geneities in the galaxy, such as stars, globular clusters and large molecular clouds, is to
“fuzz up” the phase space sheets. Consider a phase space sheet which in the absence of
scattering produces on Earth a stream with unique velocity ~v . The collective effect of scat-
tering by a class of ob jects of mass M and density n is to diffuse the velocities in the stream
over a cone of opening angle ∆θ given by [9]
4G2M 2
(∆θ)2 = Z dt Z bmax
b2 v 4 nv 2πbdb
bmin
M⊙ !2
≃ 2 × 10−7 M
bmin ! Z dt
ln bmax
where the time integral is over the past history of the particles in the stream, b is the impact
parameter of a scattering and v is the velocity of the sheet relative to the scattering center.
In the galactic disk, the giant molecular clouds are most likely the main contributors. With
M ∼ 106 M⊙ , n ∼ 3 kpc−3 , bmax ∼ kpc and bmin ∼ 20 pc, they yield ∆θ ≃ 0.05 for dark
matter particles that have spent most of their past in the galactic disk. The contributions
due to globular clusters (M ∼ 5×105M⊙ , n ∼ 0.3 kpc−3 ) and stars (M ∼ M⊙ , n ∼ 0.1 pc−3)
are less important. At any rate, peaks due to dark matter particles that have spent much
of their past in the central parts of the Galaxy are likely to be washed out. On the other
hand, the peaks due to dark matter particles which have fallen in and out of the Galaxy
only a small number of times in the past are not erased by scattering.
(300 km/s)3
1010 year pc−3 ,
n
v 3
(2.1)
B. Angular momentum.
The presence of the rotating galactic disk clearly indicates that the baryons in our galaxy
carry angular momentum. This angular momentum is thought to have been produced by
the gravitational forces of nearby galaxies when ours started to form. One should expect the
dark matter in the galactic halo to have similar amounts of angular momentum and hence
to move on non-radial orbits. If an infalling particle’s angular momentum is large enough,
its distance of closest approach to the galactic center is larger than our own distance (≃ 8.5
kpc) to the galactic center and hence it can not possibly reach us. It is nonetheless true that
particles falling onto the galaxy for the first time reach us at all times, even if the typical
distance of closest approach of such particles to the galactic center is much larger than 8.5
kpc.
Indeed, consider all particles that reach their turnaround radius at a given time.
“Turnaround” refers to the moment in a particle’s history when it reaches zero radial veloc-
ity with respect to the galactic center for the first time, after its initial Hubble flow velocity
has been halted by the gravitational pull of the galaxy and before it starts to fall onto the
galaxy for the first time; see Fig. 1. All particles that reach their turnaround radius at a
given time are on a surface which, from a topological viewpoint, is a sphere enclosing the
galactic center. Let us call this surface the “turnaround sphere”. By consulting a catalog of
6
the galaxies in our neighborhood and plotting their radial velocities as a function of distance,
one concludes that the radius of the present turnaround sphere is of order 1-2 Mpc for our
galaxy. Consider the turnaround sphere at an arbitrary time t. At any point on that surface
the angular momentum vector has a unique value parallel to the surface. Now, it is well
known that a continuous vector field on a 2-sphere can not everywhere differ from zero. It
must have at least two zeros.
Hence there are two places on any turnaround sphere where the angular momentum
vanishes. The particles from these two locations will pass through the galactic center when
they fall onto the galaxy next, producing two velocity peaks there. By continuity (the
phase space sheet can not tear), other particles on the turn-around sphere will produce two
velocity peaks at any point sufficiently close to the center. Fig. 2 shows the time evolution of
a turn-around sphere which is initially rigidly rotating about an axis and which subsequently
is moving under the influence of the gravitational potential of an isothermal sphere. The
figure demonstrates that any turnaround sphere passes (at least) twice by any point inside
of it, once on the way in and once on the way out, assuming only that the point is inside
the sphere both at its first and its second turnaround. By definition, second turnaround is
when the sphere reaches its maximum size for the second time in its history, just after its
first oscillation. Thus we find that when angular momentum is included, there are still two
(possibly more but necessarily an even number) velocity peaks on Earth due to particles
falling through the galaxy for the first time, two peaks or more due to particles falling
through the galaxy for the second time, and so on. As we saw in the preceding subsection,
these peaks are not erased by scattering off stars, globular clusters or giant molecular clouds.
The sizes and velocity magnitudes of these peaks constitute the main topic of this paper.
Finally, let us note the effect angular momentum has on the caustics of a halo. A
“caustic” is a place where the dark matter density is large because the phase space sheet folds
back there. The density actually diverges at the caustic in the limit where the thickness of
the phase space sheet goes to zero. There is an outer caustic surface near the n-th turnaround
radius with n = 2, 3, 4... ; see Fig. 1. It can be shown that near a caustic surface, the dark
matter density behaves as ρ ∼ Θ(x)/√x + constant, where x is the distance to the caustic
surface and Θ(x) is the Heaviside function: Θ(x) = 1 for x > 0 and Θ(x) = 0 otherwise.
Now, when angular momentum is absent, the center of the galaxy is a very special point
because all dark matter particles go through the center at each orbital oscillation. The dark
matter density goes as ρ ∼ 1/r2 at the center if there is no angular momentum and the
rotation curve is flat. Thus, in the absence of angular momentum the galactic center is a
caustic point. When angular momentum is present, that caustic point spreads into a set of
inner caustic rings. Fig.2 shows the appearance of such an inner caustic ring for the case of
axial symmetry. The fact that the caustic appears has nothing to do with axial symmetry
however. Rather, it is a consequence of the fact that when a sphere is turned “inside out”,
as illustrated in Fig.2, a ring singularity must appear on the surface at some point during
the process. Generally, the caustic ring appears near the place where the particles with
the most angular momentum on a given turn-around sphere turn back at their distance of
closest approach to the galactic center. Outer caustics at the Earth’s location are likely
to be very much degraded by scattering of the dark matter particles off inhomogeneities in
the galaxy. However, inner caustics associated with particles which have gone through the
central parts of the galaxy only a small number of times in the past are not much degraded.
7
The dark matter density on Earth could be much enhanced if we happen to be close to an
inner caustic.
C. On the velocity dispersion of infalling cold dark matter.
In this study, when obtaining estimates of the average sizes and of the velocity magnitudes
of the velocity peaks, we neglect the velocity dispersion δvin the cold dark matter has when
it falls onto the galaxy for the first time. Presumably, this is a valid approximation provided
δvin is much smaller than the velocity dispersion δvgal ∼ 10−3 of the galaxy as a whole. We
argue in this section that this condition is probably satisfied although we will not attempt
to provide a reliable estimate for the size of δvin . The width δvn of the velocity peaks due to
particles falling in and out of the galaxy for the n-th time, where n is sufficiently small that
the broadening effect of scattering of the particles by the galaxy’s inhomogeneities can be
neglected, is related to δvin by Liouville’s theorem: δvn = δvin (t∗,n )[ρn /ρ(t∗,n )]1/3 where ρn
is the contribution to the local halo density from particles in the n-th peak, and δvin (t∗,n )
and ρ(t∗,n ) are the velocity dispersion and density those particles had at the time t∗,n of
their first turn-around.
Let us emphasize that the values of the peak widths δvn may some day be measured in a
direct CDM detection experiment and that such data would provide information about our
universe which is not readily accessible by other means. Also, if the widths of some peaks
are small enough the sensitivity of the cavity detector of dark matter axions is improved
by looking for narrow peaks. In the case of the present LLNL experiment [7], which does
look for narrow peaks in addition to looking for a signal whose width is set by the galaxy’s
overall velocity dispersion δvgal ∼ 10−3 , the sensitivity of the search is improved if there is
a velocity peak with δvn less than about 10−8 and with a fraction of the local density larger
than about 1%.
Turning to the question how large δvin may be, let us start by describing the primor-
dial velocity dispersion which is the contribution that is present even if the universe were
completely homogeneous, i.e., it is the value of δvin if our galaxy were the only density
perturbation in the universe. For WIMPs, the primordial velocity dispersiom δvW is due
to the finite temperature TD the WIMPs have when their kinetic energies decouple from
the primordial heat bath. Thus δvW ∼ (2TD /m)1/2 (RD /R0), where m is the WIMP mass,
and RD and R0 are the scale factors at temperature TD and now. For m ∼ 50 GeV and
TD ∼ 1 MeV, one has vW ∼ 10−12 which is very small. For axions, the primordial velocity
dispersion is due to the inhomogeneity of the axion field at temperature T1 ≃ 1 GeV and
time t1 ≃ 2 × 10−7 sec when the axion mass becomes equal to the Hubble expansion rate. If
there is no inflation after the Peccei-Quinn phase transition at which the UP Q(1) symmetry
gets spontaneously broken, then the scale of inhomogeneity of the axion field is of order
the horizon scale at time t1 and the primordial axion velocity dispersion today is therefore
δva ∼ (ma t1 )−1(R1/R0) ∼ 10−17 × (10−5 eV /ma). If there is inflation after the Peccei-Quinn
phase transition, then the axion field gets homogenized over enormous distances and δva is
exponentially small.
The primordial velocity dispersion, δvW or δva , discussed in the preceding paragraph is
8
the thickness of the CDM phase space sheet. It constitutes a lower bound on the velocity
dispersion δvin of infalling CDM. Additional velocity dispersion is expected because the
phase space sheet may wrap itself up on smaller scales than that of the galaxy as a whole,
as illustrated in Fig.3. The phase space sheet wraps itself up wherever an overdensity has
grown by gravitational instability past the linear regime (δρ/ρ < 1) into the non-linear one
(δρ/ρ > 1). In theories of structure formation based upon cold dark matter, the spectrum of
primordial density perturbations is flat, i.e., it has approximately equal power on all length
scales. The matter density perturbations do not grow till the time teq of equality between
the matter and radiation energy densities. After teq , all density perturbations which have
wavelength less than the horizon (this includes all length scales of order a galaxy size) grow
together at the same rate and therefore they all reach the non-linear regime at approximately
the same time. In the standard CDM cosmology, the smaller scale clumps reach the non-
linear regime somewhat earlier because the processed spectrum of density perturbations is
not exactly flat on galaxy scales but is slightly tilted with more power on small scales.
What happens in the non-linear regime is far from obvious. The rate of growth of an
overdensity in the non-linear regime is of order √Gρ, where ρ is its mean density. Indeed
1/√Gρ is of order the free infall time, which is also the time necessary to produce a new
fold in the phase space sheet. At the start of the non-linear regime, as we just argued,
all overdensities have densities of the same order of magnitude and they therefore grow at
comparable rates by locally wrapping up the phase space sheet. However, overdensities of
large physical size will tidally disrupt and therfore inhibit the growth of overdensities of
smaller physical size in their neighborhood. In the immediate vicinity of our galaxy, there
are no visible overdensities other than nearby dwarf galaxies such as the Magellanic clouds.
Dark matter and accompanying baryons are nonetheless falling onto the galaxy now for the
first time. It is the velocity dispersion of this unseen matter that we are interested in. If this
matter is in large clumps, one might expect it to light up stars and thus become visible. On
the other hand, it could be in clumps which have not lit up for some reason. However, any
known ob ject smaller than a galaxy (e.g. stars, globular clusters, large molecular clouds,
dwarf galaxies) has velocity dispersion smaller than δvgal ≃ 10−3 , and dark matter ob jects
should be expected to be less clumped than baryonic ob jects because they can not dissipate
their energy. On this basis, it seems safe to assume that δvin is considerably less than δvgal .
There is a particular kind of clumpiness which is expected to affect axion dark matter if
there is no inflation after the Peccei-Quinn phase transition. This is due to the fact that cold
dark matter axions are inhomogeneous with δρ/ρ ∼ 1 over the horizon scale at temperature
T1 ≃ 1 GeV when they are produced at the start of the QCD phase-transition, combined
with the fact that their velocities are so small that they do not erase these inhomogeneities
by free-streaming before the time teq when matter perturbations can start to grow. These
particular inhomogeneities in the axion dark matter are immediately in the non-linear regime
after time teq and thus form clumps, called “axion mini-clusters” [14–16]. These have [16]
mass Mmc ≃ 10−13M⊙ and size lmc ≃ 1012 cm, and therefore their associated velocity
dispersion vmc = qGMmc /lmc ≃ 10−10 at time teq . This velocity dispersion increases by
about a factor 10 from teq till the onset of galaxy formation because of the hierarchical
clustering of the axion mini-clusters. This yields vmc ≃ 10−9 as the contribution of mini-
clusters to the velocity dispersion δvin of infalling axions if there is no inflation after the
Peccei-Quinn transition.
9
III. SELF-SIMILAR INFALL MODELS.
A. The radial infall model.
The tool we use to obtain estimates of the sizes and the velocity magnitudes of the
highest energy peaks is the secondary infall model of galactic halo formation. In its original
form, this model is based on the following assumptions:
1. the dark matter is non-dissipative
2. it has negligible initial velocity dispersion
3. the gravitational potential of the galaxy is spherically symmetric and is dominated by
the dark matter contribution
4. the dark matter particles move on radial orbits through the galactic center.
Assumption 1 means that the only force acting upon the dark matter particles is the grav-
itational pull of the galaxy. Assumption 2 states that before the galaxy starts to form, at
some initial time ti , all the dark matter particles at the same position ~ri relative to the
galactic center move with the same velocity ~vi . Provided ti is chosen early enough, this
initial velocity is given by the Hubble expansion:
~vi = H (ti)~ri .
(3.1)
H (ti ) is the Hubble rate at time ti . The issue of the validity of assumption 2 is discussed at
length in the previous section. Henceforth, we will take its validity for granted. Assumption
3 is realistic because the gravitational potential of a galaxy as a whole (luminous plus dark
matter) is nearly spherically symmetric even if the distribution of its luminous matter is not
spherically symmetric at all. Assumption 4 is the most doubtful. As was discussed in the
previous section, the rotating disks of spiral galaxies show that their baryons carry angular
momentum and one should expect the dark matter to have similar amounts of angular
momentum and hence to move on non-radial orbits with distances of closest approach to the
galactic center at least of order the radius (∼ 10 kpc) of the disk. Assumption 4 is motivated
mainly by simplicity. Below, in the next subsection, we will generalize the model to rid it
of this assumption. For clarity, we refer to the model with the fourth assumption included
as the radial infall model.
Let us call Mi the mass inside ri at the initial time ti . In a perfectly homogeneous and
flat universe, Mi is equal to
M Ω=1
i
=
4π
3
ρ(ti )r3
i =
i H (ti )2
r3
2G
=
2r3
i
9Gt2
i
,
if we take ti to be in the matter dominated epoch. Instead,
Mi (ri ) =
2r3
i
9Gt2
i
+ δMi (ri ) .
10
(3.2)
(3.3)
where δMi (ri ) is a spherically symmetric overdensity. The dark matter shell which is initially
at radius ri has initially the radial velocity
vi (ri ) = H (ti )ri = 2ri/3ti ,
(3.4)
assuming that ti is small enough so that the very earliest deviations from perfect Hubble
flow may be neglected. The position r(ri , t) of each shell at time t is determined by solving
the equations
d2r
dt2 = −
G M (r, t)
r2
,
(3.5)
M (r, t) = Z ∞
0
with the initial conditions given in Eq. (3.4). Θ(x) is the Heaviside function, defined earlier.
Θ(r − r(ri , t)) ,
dMi
dri
(3.6)
dri
The qualitative evolution of the dark matter distribution in phase-space (r, r) may be
described as follows. Initially, the dark matter particles are located on the line r = H (ti )r .
As time goes on, this line “winds up” in a clockwise fashion rotating most rapidly near
r = r = 0. Fig.1 shows the line on which the dark matter particles are located at a
particular moment in time.
The radius r(ri , t) of a given shell initially increases till it reaches a maximum value r∗(ri )
at a time t∗ (ri ). r∗ (ri ) and t∗ (ri ) are called the turnaround radius and turnaround time of
shell ri . After t∗ (ri ), the radius of shell ri will oscillate with decreasing amplitude. As long
as it does not cross any other shells, the mass interior to shell ri is constant, with value Mi ,
and its motion is the well-known motion of a particle attracted by a central mass Mi in the
limit of zero angular momentum. Shell ri does not cross any other shells till some time after
t∗ (ri ), when it is falling onto the galactic center for the first time. Thus, one readily finds:
2 vuut
π
r∗ (ri ) = ri
r∗ (ri )3
2GMi (ri )
t∗ (ri ) =
(3.7b)
(3.7a)
,
and
Mi (ri )
δMi (ri )
.
After a given shell crosses other shells its motion depends on that of the other shells and
becomes more difficult to determine.
Much progress in the analysis of the model came about as a result of the realization
[11,12] that Eqs.(3.5) and (3.6) have self-similar solutions for appropriate initial conditions.
A solution is self-similar if it remains identical to itself after all distances have been rescaled
by a time-dependent length R(t) and all masses by a time-dependent mass M (t). R(t) is
taken to be the radius at which dark matter particles are turning around at time t; see
Fig. 1. M (t) is taken to be the mass interior to R(t) at time t. So, R(t) = r∗(ri ) and
M (t) = Mi (ri ) with ri such that t∗ (ri ) = t. A self-similar solution has the properties:
11
and
M (r, t) = M (t)M(r/R(t)) ,
r(ri , t) = r∗ (ri )λ(t/t∗(ri )) ,
(3.8)
(3.9)
where M and λ are functions of a single variable. Let us verify that indeed the evolution
is self-similar for appropriate initial conditions. Substituting Eqs (3.8) and (3.9) into Eq.
(3.5) and using Eq. (3.7a), one finds:
d2λ
dτ 2 = −
π 2
8λ2
R(t) ! ,
Mi (ri )M r∗(ri )λ
M (t)
(3.10)
,
δMi (ri )
Mi (ri )
where τ ≡ t/t∗ (ri ). We want the RHS of Eq. (3.10) to depend only upon τ and λ(τ ). This
happens for the initial condition:
Mi (ri ) !ǫ
= M0
where M0 and ǫ are parameters. ǫ should be in the range 0 ≤ ǫ ≤ 1, since ǫ = 0 corresponds
to the extreme case of a ri -independent overdensity whereas ǫ = 1 corresponds to the extreme
case of an excess point mass located at r = 0. The initial density profile (3.11) does not have
any feature that would distinguish an epoch in the evolution of the galactic halo from other
epochs. It is this “scale free” property that makes the initial density profile (3.11) consistent
with self-similarity, as we are about to show. From now on, for the sake of convenience and
following Fillmore and Goldreich, we will use Mi instead of ri to label the shells. Using Eqs.
(3.3) and (3.11), and neglecting terms of order δMi/Mi versus terms of order one, we find
that Eqs. (3.7) become
(3.11)
,
3π
4
t∗ (ri ) =
3ǫ/2
ti (cid:18) Mi
M0 (cid:19)
∗ (Mi )GMi(cid:21)1/3
r∗(Mi ) = (cid:20) 8
π 2 t2
.
Hence
Therefore
,
3π ti (cid:19)2/3ǫ
M (t) = M0 (cid:18) 4t
π 2 M (t)#1/3
R(t) = " 8t2G
and
.
M (t)
Mi
=
M (t)
M (t∗ (Mi ))
t∗(Mi ) !2/3ǫ
= t
= τ 2/3ǫ ,
12
(3.12a)
(3.12b)
(3.13a)
(3.13b)
(3.14a)
r∗ (Mi )
R(t)
=
R(t∗ (Mi ))
R(t)
t !2/3+2/9ǫ
= t∗ (Mi )
= τ −2/3−2/9ǫ .
(3.14b)
Thus, Eq. (3.10) has the desired form:
d2λ
dτ 2 = −
π 2
8
τ 2/3ǫ
λ2 M
τ 2/3+2/9ǫ ! .
λ
(3.15)
Similarly, using Eq. (3.8) and Eqs. (3.14), we rewrite Eq. (3.6) as
Θ ξR(t) − r∗(Mi )λ
dMi
M (ξR(t), t)
= Z ∞
M (t)
M (t)
0
τ 2/3+2/9ǫ ! ,
τ 1+2/3ǫ Θ ξ −
2
λ(τ )
dτ
3ǫ Z ∞
1
which also has the desired form. τ varies from 1 to ∞. The boundary conditions at τ = 1
are:
r∗(Mi ) !!
t
(3.16)
M(ξ ) =
=
λ(1) = 1,
dλ
dτ
(1) = 0 .
(3.17)
Fillmore and Goldreich [11] integrated Eqs. (3.15) and (3.16) numerically for various values
of ǫ. They also derived analytically the behaviour of M(ξ ) when ξ → 0 using adiabatic
invariants to obtain the motion of the shells. Bertschinger [12] analyzed the case ǫ = 1.
Let us mention in passing that the ratio of the density at the turnaround radius to the
critical density is ρ∗/ρc = 9π 2/[16(3ǫ + 1)].
B. Secondary infall with angular momentum.
As was emphasized earlier, the assumption that the infalling dark matter is devoid of
angular momentum with respect to the galactic center is inadequate for the calculation of
the velocity peaks which are the main topic of this paper. However, it is possible to include
the effect of angular momentum into the secondary infall model while keeping the model
tractable. We will do this in two steps. First we will assign the same value of angular
momentum magnitude to all particles in a given shell. Second, we will assign the particles
in a given shell a distribution of angular momentum magnitudes. The model which results
from the first step is not so realistic but it is easier to explain.
1. Single magnitude of angular momentum for al l particles on a shel l.
Let’s assume first that the particles belonging to shell Mi all have the same magnitude
of angular momentum l(Mi ) and that they all have, at some initial time, the same radial co-
ordinate r(Mi , t) and the same radial velocity vr (Mi , t) = ∂ r(Mi , t)/∂ t. We further assume
13
that at any point ~r(Mi , t) = rr(Mi , t) on shell Mi , all particles in the shell have their trans-
verse velocities ~v⊥ (Mi , r , t) = l(Mi ) ϕ/r(Mi , t) isotropically distributed about r , i. e. each
direction ϕ perpendicular to r is equally much represented. As a result of these assumptions,
each shell remains spherical as it moves through the spherically symmetric mass distribution
M (r, t) due to all other shells. Initially, at time ti , the shell Mi has radial velocity
vr (Mi , t) = H (ti )r(Mi , ti ) =
2r(Mi , ti )
3ti
.
(3.18)
Eqs. (3.2) and (3.3) hold as before. Provided l(Mi ) is not too large, the turnaround radius
r∗(Mi ) and time t∗ (Mi ) are still given by Eqs. (3.7) to a very good approximation. We will
use these equations, neglecting the corrections therein due to l(Mi ) 6= 0 and find that self-
similarity is possible. Note, however, that if the corrections to Eqs. (3.7) due to l(Mi ) 6= 0 are
included one still finds self-similarity to be possible. The reason we neglect the corrections is
not to obtain self-similarity but because these corrections are truly small for realistic values
of angular momentum.
To obtain self-similarity we assume as a necesary, but no longer sufficient, condition that
the initial mass distribution is given by the scale free power law of Eq. (3.11). Eqs. (3.12),
(3.13), (3.14) are then still valid as well. Each dark matter particle satisfies:
Substituting therein
d2r
dt2 =
l2
r3 −
GM (r, t)
r2
.
t∗(Mi ) ! ,
r(Mi , t) = r∗(Mi )λ t
M (r, t) = M (t)M r
R(t) ! ,
one obtains, using Eqs. (3.12) - (3.14):
d2λ
dτ 2 =
l(Mi )2 t∗(Mi )2
r∗ (Mi )4λ3 −
π 2
8
τ 2/3ǫ
λ2 M
τ 2/3+2/9ǫ ! ,
λ
(3.19)
(3.20a)
(3.20b)
(3.21)
where τ = t/t∗(Mi ) as before. To obtain self-similar solutions, we must make the additional
assumption that
l(Mi ) = j
r∗ (Mi )2
t∗ (Mi )
,
where j is a constant. Then
d2λ
dτ 2 =
j 2
λ3 −
π 2
8
τ 2/3ǫ
λ2 M
τ 2/3+2/9ǫ ! .
λ
14
(3.22)
(3.23)
Eq. (3.16) for M(ξ ) and the boundary conditions (3.17) remain unchanged.
The question arises how realistic the model is in the above form. Consider shell Mi near
its turnaround time t∗ (Mi ). The actual angular momenta of the particles in the shell are of
course not distributed as in the model. The model assumes ~v⊥ to be isotropically distributed
about ~r . Instead, angular momentum has a unique value ~l(Mi , ~r) at each point, with ~l(Mi , ~r)
changing from point to point on the shell. As was discussed in Section II, ~l(Mi , ~r) must have
at least two zeros on the shell. This implies that there are necessarily some particles in the
shell which will pass through the center of the galaxy and, by continuity, other particles will
reach the Earth as well. There are two velocity peaks in the spectrum of cold dark matter
particles on Earth due to particles falling into and out of the galaxy for the first time, two
peaks due particles falling into and out of the galaxy for the second time, etc. In contrast,
in the above model, after a certain galactic age none of the particles falling onto the galaxy
for the first time reach the Earth because these particles have too much angular momentum.
Their distance of closest approach to the galactic center exceeds our own distance to the
galactic center. If we turn for a moment to the results of the computer simulations, we see
that for the example of Figs.6 and 7, in which j = 0.2, only particles which are falling in
and out of the galaxy for the n-th time with n > 3 can presently reach us.
2. Distribution of magnitude of angular momenta on a shel l.
In reality, particles at different locations on a given shell have different values of vector
angular momentum. As a result, the time evolution of a shell is not spherically symmetric
when angular momentum is present. This is illustrated by Fig.2 in a special axially sym-
metric case. However, we can restore spherical symmetry by averaging over all possible
orientations of a physical halo. This corresponds to adopting the model of the previous
subsection but with a distribution of magnitudes of angular momentum for the particles in
each shell. Let each shell Mi have a fraction nk (Mi ) of particles with magnitude of angular
momentum lk (Mi ) where k = 1, . . . , K . To obtain self-similar solutions, nk (Mi ) must be
independent of Mi and
lk (Mi ) = jk
r∗ (Mi )2
t∗(Mi )
.
The equations for self-similar solutions are then
rk (Mi , t) = r∗ (Mi )λk t
t∗ (Mi ) ! ,
d2λk
dτ 2 =
j 2
k
k −
λ3
π 2
8
τ 2/3ǫ
k M
λ2
τ 2/3+2/9ǫ ! ,
λk
M(ξ ) =
2
3ǫ
K
Xk=1
nk Z ∞
1
τ 1+2/3ǫ Θ ξ −
dτ
τ 2/3+2/9ǫ ! ,
λk (τ )
15
(3.24)
(3.25a)
(3.25b)
(3.25c)
K
Xk=1
In our calculations, we shall take jk to be distributed according to the density
nk = 1 .
dn
dj
=
2j
j 2
0
exp(−j 2/j 2
0 ) .
(3.25d)
(3.26)
Let us explain this choice, starting with the behaviour of dn/dj near j = 0. The angular
momentum field on a sphere must have at least two zeros. Let us choose the origin (θ = 0)
of polar coordinates to be at one of them. Assuming that the zero is simple, the Taylor
expansion of the magnitude of angular momentum function j (θ, φ) in powers of θ starts
with the term linear in θ: j ∼ θ. Then we have: dn/dj ∼ dn/dθ ∼ θ ∼ j . The cutoff at
large j in Eq. (3.26) was chosen to be Gaussian for the sake of convenience. The actual
distribution likely has a sharp cutoff, with a maximum value of angular momentum, but
that is very similar to a Gaussian. We express our results below in terms of the average over
the distribution, ¯j = √πj0/2.
A benefit of including angular momentum into the secondary infall model is to produce
galactic halos with an effective core radius. The radial infall model, i.e. the model without
angular momentum, produces flat rotation curves for 0 < ǫ < 2/3. Adding angular momen-
tum has the effect of depleting the inner halo and hence of making the halo contribution to
the rotation curve go to zero as r → 0. This is desirable because, in spiral galaxies like our
own, it is the sum of the contributions from the halo, the disk and the bulge that produces
flat rotation curves, and the central parts of the galaxy are known to be dominated by the
bulge and the disk. We define the “effective halo core radius” b as the radius at which half
of the rotation velocity squared is due to the halo. In our galaxy b is estimated to be of
order 10 kpc. We will find below that this implies ¯j ∼ 0.2 in the model.
The secondary infall model with a distribution of angular momentum described in this
subsection still has shortcomings due to the fact that the model averages over all possible
orientations of a physical halo. In particular, the model is only able to produce estimates of
the average sizes of velocity peaks. The averages are over all locations at the same distance
from the galactic center as we are. At some of these locations, a particular velocity peak
may be much larger than average because that location happens to be close to an inner
caustic.
C. Inclusion of baryons
One may also wish to include the effect of the gravitational potential of the galactic
bulge and disk. As was already noted, the disk and bulge of our galaxy are conspiring
with its dark matter halo to produce an everywhere approximately flat rotation curve. We
may reasonably assume that this was also true in the past because most spiral galaxies are
observed to have approximately flat rotation curves and they do not all have the same age.
This suggests a simple way to include the effect of baryons in the self-similar secondary
infall model, to wit: first obtain for given ǫ the mass function M(ξ ) of the model without
16
angular momentum and then use that mass function and Eq.(3.25b) to obtain the phase
space distribution of the dark matter in the model with angular momentum. We will find
below that including the gravitational field of the disk and bulge in this manner does not
have much effect upon the sizes of the highest energy peaks but it does shift their kinetic
energies upward by deepening the potential well at our location.
D. ǫ and the spectrum of initial density perturbations.
The spectrum of the cosmological density perturbations which give rise to galaxies con-
tains information about the likely value of the parameter ǫ. It has been shown [17] that, if
the density perturbations have a Gaussian distribution, the average density profile around
a peak in the density distribution is given simply by
< δ(r) >= δ(0)
ζ (r)
ζ (0)
(3.27)
where δ(~r) ≡ δρ(~r)/ρ and ζ (r) ≡< δ(~r)δ(0) > is the 2-point correlation function. The latter
is related to the power spectrum P (k) by
ζ (r) = Z exp(i~k .~r)P (k)d3k .
The power spectrum is the product P (k) = AknT 2(k) where Akn is the primordial spectrum,
taken for simplicity to be a power law, and T (k) is the transfer function. For cold dark
matter, the transfer function is given by [18]
× h1 + 3.89q + (16.1q)2 + (5.46q)3 + (6.71q)4i−1/4
where q = k Mpc/h2 . The Harrison-Zel’dovich spectrum corresponds to n = 1. Eqs.(3.27)
and (3.28) imply that if P (k) ∼ kα on some momentum scale k , then ζ (r) ∼ r−α−3 and hence
ǫ = (α + 3)/3 on the corresponding length scale r = 1/k . We computed α = d lnP /d lnk for
n = 1 and plotted the resulting ǫ(r) in Fig. 4 for the relevant scales. The figure suggests
that ǫ is of order 0.2 - 0.3 on the galactic scale.
ln(1 + 2.34q)
2.34q
T (k) =
(3.28)
,
(3.29)
IV. NUMERICAL INTEGRATION
In this section, we present the results from numerically integrating Eqs. (3.25b,3.25c)
for various values of the parameter ǫ and various angular momentum distributions, includ-
ing no angular momentum, a single value of angular momentum and the distribution of
Eq.(3.26). The function λ(τ ) gives us the phase-space distribution of the particles through
the equations:
t∗(Mi ) ! = R(t)λ(τ )τ −2/3−2/9ǫ ,
r(Mi , t) = r∗(Mi )λ t
(4.1a)
17
v (Mi , t) =
dr(Mi , t)
dt
=
R(t)
t
τ 1/3−2/9ǫ dλ
dτ
.
(4.1b)
If there is a distribution of angular momentum values, the functions λ(τ ), r(Mi , t) and
v (Mi , t) carry an index k which we have suppressed here to avoid cluttering the equations.
To solve Eq.(3.25b) for the particle tra jectory λ(τ ) we need to know the mass function
M(ξ ). This function, in turn, is given in terms of the tra jectory λ(τ ) by Eq.(3.25c) or,
equivalently, by:
M(ξ ) = Xn=1 (cid:16)τ −2/3ǫ
2n−1 (ξ ) − τ −2/3ǫ
(ξ )(cid:17) ,
2n
where the τj (ξ ) correspond to the moments of time when the tra jectory crosses radius
r = ξR(t), i.e. they are the solutions of λ(τ ) = ξ τ 2/3+2/9ǫ . Following Fillmore and Goldreich
[11], we solve Eqs.(3.25b) and (3.25c) simultaneously by a technique of successive iterations.
Starting with some arbitrary mass profile M(ξ ) (we took M(ξ ) = ξ 2) we find λ(τ ), which
is then used to derive a new mass profile, from which a new tra jectory is derived, and so
on. The procedure is repeated till it converges. We find that the mass profile changes very
little after 5 iterations. Typically we run 10 iterations to get the final results.
(4.2)
Fig. 5 shows the phase-space diagram for the case ǫ = 0.2 and zero angular momentum.
The solid line in that figure shows the location of all the particles in phase space at a
given time, i.e.
it is the set of points (r(Mi , t), v (Mi , t)) for all Mi . The radial distances
are normalized to the turnaround radius R at time t and the velocities are normalized to
qGM (t)/R(t) = πR(t)/√8t which is the rotation velocity at the turnaround radius. Fig. 6
shows the phase space diagram for the case ǫ = 0.2 and a single value of angular momentum
j = 0.2. The particle tra jectory λ(τ ) for that case is shown in Fig. 7.
A convenient way to show the mass distribution is by showing the rotation curve. We
define ν (ǫ, ξ ) by:
rot (r) = GM (r, t)/r ≡ ν 2 ǫ,
v 2
R(t) ! GM (t)
r
R(t)
.
(4.3)
With this definition, we have ν (ǫ, ξ = 1) = 1. The functions ν (ǫ, ξ )2 obtained by numerical
integration are plotted in Fig. 8 for various values of ǫ and j = 0.
To fit the model to our galactic halo, we must choose values of the present turnaround
radius R ≡ R(t) and of M ≡ M (t). Equivalently, we may choose values of the present age t
and of R. M is given in terms of R and t by Eq. (3.13b). t is given in terms of the Hubble
rate H0 = h 100 km s−1 Mpc−1 by the relation t−1 = 3H0/2. We will use h to state the
age of the universe. Then we fix R in terms of h by requiring that the model reproduce the
measured value, vrot = 220 km s−1 , of the rotation velocity in our galaxy. Let us call ν (ǫ)
the value of ν (ǫ, ξ ) in the flat part of the rotation curve, near r = 0.02R for ǫ < 0.4; see Fig.
8. ν (ǫ) is related to vrot by Eq. (4.3). This implies:
R h = 1.32 ν (ǫ)−1 Mpc .
(4.4)
Table I gives ν (ǫ)2 , Rh and M h for various values of ǫ.
18
Our distance to the galactic center is taken to be 8.5 kpc and we define ξs ≡ 8.5 kpc/R.
The contribution of the n-th velocity peak to the local halo density is given by:
ρn =
M
4πR3ξ 2
s
6τ 2/3−4/9ǫ
n
9ǫτn λn − (6ǫ + 2)λn
where λ ≡ dλ/dτ , and λn and τn are the solutions of λ(τ ) = ξsτ 2/3+2/9ǫ . The kinetic energy
(in a frame of reference which is not rotating along with the galactic disk) per unit particle
mass in the n-th peak is given by:
,
(4.5)
En =
1
2
v 2
n =
R2
2t2 τ 2/3−4/9ǫ
n
j 2
λ2
n
+ λn
2! .
(4.6)
We shall express En in units of (300 km s−1)2/2 when presenting our results.
We now discuss in greater detail the results specific to the different types of angular
momentum distributions used.
A. Radial infall
Without angular momentum all particles pass through the origin, r = 0, at each oscilla-
tion. To avoid infinities in the numerical integration, a regulator at small r is required. The
one which is most convenient for us and which we use is to give a small amount of angular
momentum to the dark matter particles. We found j 2 = 10−6 to be small enough for our
purposes.
Fig. 5 shows a typical phase space distribution. Fig. 8 shows the rotation curves
for various values of ǫ. An analytical treatment of the radial infall model using adiabatic
invariants predicted [11] the behaviour of the density near the origin to be: ρ ∝ r−9ǫ/(3ǫ+1)
in the range 2/3 ≤ ǫ ≤ 1 and ρ ∝ r−2 in the range 0 < ǫ ≤ 2/3. These predictions agree
very well with our results. The rotation curves do indeed go to a constant near the origin
when 0 < ǫ ≤ 2/3 except for small logarithmic corrections. The analytical treatment given
in section V suggests the behaviour ρ ∼ 1/(r2qln(1/r)) for small ǫ.
Fig. 9 shows the velocity peaks on Earth predicted by the radial infall model with ǫ = 0.2
and h = 0.7. The rows labeled ¯j = 0.0 in Table II give the density fractions and kinetic
energies of the five most energetic incoming peaks for the cases ǫ = 0.2 and 1.0, and h = 0.7.
For each incoming peak there is an outgoing peak with approximately the same energy and
density fraction. We find that, in the radial infall model, the sizes of the two peaks due to
particles falling in and out of the galaxy for the first time are large, each containing of order
10% of the local halo density for ǫ in the standard CDM model inspired range of 0.15 - 0.4.
As was emphasized already, this spectrum is unrealistic because angular momentum has a
non-negligible effect upon the peak sizes.
19
B. Single non-zero value of angular momentum.
Fig. 6 shows the phase space diagram for the case ǫ = 0.2 and j = 0.2. In the model
with a single non-zero value of angular momentum, the halo distribution has a set of inner
caustics in addition to the usual outer caustics. However, the inner caustics are spheres in
the model whereas they are rings for a physical halo. The density profile for the model is
shown in Fig. 10. There is a break in the logarithmic slope near the first inner caustic, near
r = 0.01R in the figure. For ǫ < 2/3 we find ρ ∝ r−2 outside the first inner caustic (the same
as with j = 0) but ρ ∝ r−γ inside with γ = 9ǫ/(3ǫ + 1). This observation is related to the
fact that we find the function λ(τ ) to be oscillating with constant amplitude and constant
period for large τ when j 6= 0; see Fig. 7. This can happen only if τ 2/3ǫM (cid:16)λ/τ 2/3+2/9ǫ (cid:17) in
Eq. (3.25b) is independent of τ at large τ . For M(r) ∝ rα , this implies α = 3/(3ǫ + 1) or
γ = 3 − α = 9ǫ/(3ǫ + 1). We find that the exponent α does not depend upon j , but the
radius rc at which the break in power law behaviour occurs does depend upon j : the smaller
j , the smaller rc .
These observations may be understood as follows. When angular momentum is present,
the contribution of a single phase-space sheet to the density profile ρ(r) is non-singular near
r = 0. In that case, as a result of self-similarity the mass profile M (r), at r small enough
that the influence of the first one or two phase-space sheets may be neglected, has the same
functional dependence upon r as the dependence of M upon R after eliminating t from the
expressions for M (t) and R(t) in Eqs. (3.13). This yields M ∝ r3/(3ǫ+1) . The fact that ρ(r)
behaves as a negative power of r near r = 0 is in agreement with the results of N-body
simulations [19]. The ǫ-dependence of this power law was already known to White and
Zaritsky [20].
The spectrum of velocity peaks that the model with a single value of angular momentum
predicts is unrealistic. In particular, there are in this model no peaks on Earth associated
with particles falling in or out of the galaxy for the n-th time with n small, because such
particles have too much angular momentum to reach the Earth radius. In the case of Fig.
6, there are only peaks for n > 3. But as we argued at length in section II, there are in
reality peaks on Earth due to particles falling in and out of the galaxy for the n-th time
with n = 1, 2, 3, ... Angular momentum reduces the sizes of the peaks with small n but does
not suppress them completely.
C. Distribution of angular momenta.
This model describes a physical halo averaged over all possible orientations. The particles
are assumed to have the distribution of angular momenta given in Eq.(3.26). For the purpose
of numerical integration, the angular momentum was discretized with a spectrum of four
hundred values. The phase space diagram is like the one of Fig. 6 except that there are four
hundred such diagrams superimposed on one another, one for each value of j .
As was mentioned earlier, angular momentum has the effect of making the contribution
of the halo to the rotational velocity go to zero at the galactic center as shown by Fig. 11.
We define the “effective core radius” b to be such that ν 2 (ǫ, b/R) = ν 2 (ǫ)/2. We find that
20
near r = 0, the density ρ(r) ∝ r−γ with γ = 9ǫ/(3ǫ + 1) as in the case of a single value
of angular momentum. This behaviour is expected for the same reason as we gave for that
case. The only change is that the transition between the region at large r where ρ(r) ∝ r−2
and the region at small r where ρ(r) ∝ r−γ is smoother now because the critical radius rc
where the transition occurs depends upon j and there is now a distribution of j values.
The velocity peaks for ¯j = 0.2, ǫ = 0.2 and h = 0.7 are shown in Fig. 12. The velocity
peaks for the same choice of parameters except ¯j = 0.4 are shown in Fig. 13 to indicate the
sensitivity of the peaks upon ¯j . The spectrum of velocity peaks is also sensitive to the value
of ǫ. It is shown for the cases ǫ = 0.15 and ǫ = 1 on Figs. 14 and 15 respectively.
Table II gives the values of the current turn-around radius R, the effective core radius b,
the halo density at our location ρ(rs ), and the density fractions and kinetic energies of the
five most energetic incoming peaks for various values of ǫ, ¯j and h.
D. Self-similar infall with baryons
As was discussed earlier, the dark matter phase space distribution for the input param-
eters ǫ, h and ¯j is obtained in this case by solving Eq. (3.25c) using the mass distribution
M(ξ ) for the same values of ǫ and h but j = 0. The resulting velocity peaks are shown in
Fig. 16 for ǫ = 0.2, h = 0.7 and different values of ¯j . Since the particles for different values
of ¯j but the same values of ǫ and h are all moving in the same gravitational potential, the
kinetic energies per unit particle mass En of the peaks depend only very weakly upon ¯j .
So, we have combined the spectra for different ¯j values in one figure. Note that the En are
larger in this case than in the case of Figs. 12 and 13 because the gravitational well near
the center of the galaxy is deeper. Fig. 17 shows the results for ǫ = 0.4, h = 0.7.
V. AN ANALYTICAL TREATMENT.
In this section, we derive analytical expressions for the sizes and locations of the velocity
peaks. The treatment involves the following approximations:
1. the mass distribution, including the contributions from both baryons and dark matter,
is taken to be M(ξ ) = ξ for 0 ≤ ξ ≤ 1
2. the dimensionless angular momentum j values are assumed to be small
3. our distance rs to the galactic center is assumed to be small compared to the oscillation
amplitudes of the particles in the peaks under consideration.
The method of adiabatic invariants will be used to obtain the motion of the dark matter
particles.
We first treat the case of a single angular momentum value, i.e the model described in
subsection III.B.1. Since M(ξ ) = ξ , the equation of motion in rescaled variables of a particle
with angular momentum j is:
21
d2λ
dτ 2 =
j 2
λ3 −
π 2
8λ
2
3 ( 2
3ǫ
−1) .
τ
(5.1)
π 2
4
dλ
dτ
Let λM and λm be respectively the amplitude of oscillation and the distance of closest
approach at time τ .
In the spirit of the method of adiabatic invariants, λM and λm are
assumed to be slowly varying functions of τ . This is a valid assumption for all oscillations
except the first one. The first oscillation is only marginally adiabatic. For given λM , the
velocity at time τ and position λ is:
= ±vuut
The adiabatic invariant is:
I (τ , λM ) = Z λM
λm
dxs−
= Z 1
λm/λM
In the limit of small j , λm
λM
Since λM (1) = 1, we have:
M ln x − j 2 (cid:18) 1
x2 − 1(cid:19) .
−1)λ2
may be neglected and we obtain λM (τ )2τ
λ ! − j 2 1
−1) ln λM
λ2 −
dλdλ/dτ
π 2
4
−1) = constant.
M ! .
1
λ2
(5.3)
(5.2)
2
3 ( 2
3ǫ
2
3 ( 2
3ǫ
2
3 ( 2
3ǫ
τ
τ
3 ( 2
λM (τ ) = τ − 1
3ǫ
−1) .
(5.4)
The period of oscillation is:
T (τ , λM ) = 2 Z λM
dλdλ/dτ −1
λm
x2 − 1(cid:19)#−1/2
π 2
dx "−
M ln x − j 2 (cid:18) 1
M Z 1
3 ( 2
2
= 2λ2
−1)λ2
3ǫ
4
λm /λM
Using Eq. (5.4), we have in the limit of small j :
4λM
−1) Z 1
1
3 ( 2
0
3ǫ
Let us introduce the “phase” ϕ(τ ) :
dx
√− ln x
4
√π
3 ( 2
τ − 2
3ǫ
T (τ ) =
−1) .
πτ
=
τ
dϕ = π
dτ
T (τ )
.
Using Eq. (5.6) and setting ϕ(1) = 0, we have:
π√π
4
ϕ(τ ) = 9ǫ
τ
1
3 ( 4
3ǫ +1) − 1
4 + 3ǫ
.
.
(5.5)
(5.6)
(5.7)
(5.8)
The times τn at which the particle passes by the solar radius rs are given by:
22
ϕ(τn ) =
π
2
(2n − 1) + o(rs/R)
(5.9)
There are two peaks for given n, one ingoing and the other outgoing. The differences between
the properties of the two peaks are of order o(rs/R) and are neglected.
The contribution of each of the two n-th velocity peaks to the local density is given by:
where
ρn =
M
4πRr2
s
3 (1− 2
2
3ǫ )
6τ
n
dλ
dτ (τn ) − (6ǫ + 2)λ(τn)
,
9ǫτn
λ(τn ) =
2
3 + 2
τ
9ǫ
n
rs
R
(5.10)
(5.11)
and dλ
dτ (τn ) is given by Eqs. (5.2) and (5.4). We neglect the second term in the denominator
of Eq. (5.10) since it is o(rs/R) relative to the first term. Combining everything, we have
the following estimate for the peak sizes due to dark matter particles with a single value of
angular momentum j :
ρn =
s ǫ
M
6πRr2
π 2
4
τ 4/3ǫ
n
ln
R
1
3 + 4
rs τ
9ǫ
n
− j 2 R2
r2
s
− 2
3
4
τ
9ǫ
n
−1/2
,
(5.12)
where
9ǫ
4+3ǫ
+
τn = " 2
√π
3 (cid:19) + 1#
1
(2n − 1) (cid:18) 4
9ǫ
Note that ρn is the local energy density contributed by each of the two n-th peaks. When the
expression under the square root in Eq. (5.12) is negative, one must set the corresponding
ρn = 0 since this corresponds to the situation where the particles have too much angular
momentum, and hence too large a distance of closest approach to the galactic center, to reach
the solar radius. The kinetic energy per unit particle mass of the dark matter particles in
the n-th peaks is:
(5.13)
.
=
+
En =
2
dτ !2
j 2
dλ
t∗ (cid:19)2
1
(cid:18) r∗
λ2
n
n
n
ln
2
π 2
8 (cid:18) R
R
t (cid:19)
+ o(rs/R)2 .
1
3 + 4
rs τ
9ǫ
n
R and t are determined in terms of h by t = 2/3H0 and Eq. (4.4) as before.
For the case ǫ = 0.2, j = 0.2, h = 0.7, the quantity under the square root in Eq. (5.12)
is negatve and hence ρn = 0 for n=1,2 and 3. This is consistent with the phase space
distribution shown in Fig. 6 which was obtained by numerical integration. We found Eqs.
(5.12 - 5.14) to be consistent with the results from numerical integration at the 30% level
(5.14)
23
or so. The agreement is worse for the case of zero angular momentum, probably because
the logarithmic singularity of the potential at the origin in that case makes the motion non-
adiabatic there. Also, when comparing results for the En values, one should allow for an
overall shift between the two calculations due to a change in the depth of the gravitational
potential at the solar radius. Indeed, the analytical calculation has the mass distribution
M(ξ ) = ξ which implies a perfectly flat rotation curve, whereas the rotation curves for the
numerical calculations are as shown in Fig.8.
The formalism readily accomodates a distribution dn/dj of angular momenta. Eq. (5.14)
for the peak kinetic energies still applies. The expected peak sizes, in the sense of an average
over all locations at distance rs from the galactic center, are given by the convolution of
dn/dj with the expression, Eq. (5.12), for the peak sizes when there is a single value of
angular momentum. Thus:
ρn =
dj
π 2
4
τ 4/3ǫ
n
− 2
3
4
τ
9ǫ
n
R
3 + 4
1
rs τ
9ǫ
n
M
s ǫ Z jn
6πRr2
0
−1/2
dj
ln
− j 2 R2
dn
r2
s
where τn is given by Eq. (5.13) as before and jn is the maximum value of j for which the
argument of the square root is positive:
vuuut
ln
.
For the angular momentum distribution of Eq. (3.26), one has
R
1
3 + 4
rs τ
9ǫ
n
9ǫ + 1
4
τ
3
n
π
2
rs
R
,
(5.15)
jn =
(5.16)
ρn =
M
6πR2rs ǫ
− 2
9ǫ
1
τ
3
n
jn
F (cid:16)(jn /j0)2(cid:17) ,
(5.17)
where
dxe−ux (1 − x)−1/2 .
F (u) ≡ u Z 1
0
A graph of F (u) is shown in Fig. 18. Table 3 shows the peak sizes predicted by Eq.(5.17) for
the first eight peaks in the case ǫ = 0.2, ¯j = 0.2 and h = 0.7. The peak sizes agree with the
results of the numerical integration, given in the fifth line of Table 3, to within 15%. The
peak energies also agree within 15% after one has subtracted an overall shift caused by the
fact that the gravitational potential at our location is considerably deeper in the analytical
calculation than in the numerical one.
(5.18)
VI. CONCLUSIONS
Motivated by the prospect that the spectrum of dark matter particles on Earth may
some day be measured in a direct detection experiment, we have endeavoured to obtain
24
predictions for the properties of that spectrum. Previously, it had generally been assumed
that the spectrum is isothermal. In contrast, we find that there will be large deviations from
an isothermal spectrum in the form of peaks in velocity space associated with particles that
are falling in and out of the Galaxy for the first time and with particles that have fallen in
and out of the Galaxy only a small number of times in the past.
To obtain estimates for the velocity magnitudes of the particles in the peaks and for
the contributions of the individual peaks to the dark matter local density, we have used the
secondary infall model of galactic halo formation. We have generalized the extant version
of that model to include the effect of angular momentum. We forced spherical symmetry
onto the model with angular momentum by averaging over all possible orientations of a
physical halo. As a result, the model can only make predictions for the average properties
of the velocity peaks, the average being over all locations at the same distance from the
galactic center as we are. We found that the model with angular momentum has self-similar
solutions if the angular momentum distribution, as well as the initial overdensity profile, has
a particular scale-free form. The self-similar solutions were analyzed in detail numerically
and analytically.
The model produces a good overall fit to what is known about galaxies like our own. It
produces flat rotation curves when the parameter ǫ is in the range 0 to 2/3. The expected
value of that parameter in models of large scale structure formation with cold dark matter
and a flat (Harrison- Zel’dovich) spectrum of primordial density perturbations is ǫ ∼ 0.25.
The model implies a relationship (cfr. Eq. (4.4) and Table I) between the current turn-
around radius R, the present age of the universe, the galactic rotation velocity and the
paramemter ǫ. This relationship is consistent with observations and ǫ ∼ 0.25.
In the model the galactic rotation curve is approximately flat all the way out to the
turn-around radius R. R is of order 1-2 Mpc for our galaxy. So far, the rotation curves of
individual galaxies have been measured, and have been found to be flat, up to distances of
order 100 kpc, implying masses of order 1012M⊙ or larger. The discovery of flat rotation
curves [21] caused a well-known revolution in our concept of galaxies. Prior to these mea-
surements galaxies were thought to have size of order 10 kpc and mass of order 1011M⊙ .
The model implies that galaxies like our own are of order 20 times bigger even than the
minimum size implied by the rotation curve measurements. The size is 1.8 Mpc and the
mass is 1.7 × 1013M⊙ for the case ǫ = 0.25, h = 0.7. See Table I for other cases. With
galaxies that massive, one has Ω ≃ 1 in galaxies.
We found that the effect of angular momentum is to deplete the inner parts of the halo
with the result that the halo contribution to the rotation curve goes to zero at r = 0.
The model establishes a direct relation between the amount of angular momentum and the
effective core radius b, defined as the radius at which the halo contributes half of the rotation
velocity squared. Table II gives the turn-around radius R, the effective core radius b and
the local halo density ρ as a function of the input parameters: ǫ, the average amount of
dimensionless angular momentum ¯j and the Hubble parameter h. The observed value (220
km/s) of our galaxy’s rotation velocity and our distance to the galactic center (8.5 kpc) were
used as fixed input parameters.
Table II also gives the average values of the peak sizes as fractions of the local halo
density ρ and the kinetic energies per unit particle mass of the particles for the first five
incoming peaks as a function of the variable input parameters ǫ, ¯j and h. For each incoming
25
peak there is an outgoing peak with approximately the same kinetic energy and average local
density. Let us emphasize again that the peak sizes given are averages over all locations at
the same distance (8.5 kpc) from the galactic center as we are. It is not possible to make
more precise predictions for the peak sizes on Earth without assuming a particular angular
momentum field for the infalling dark matter.
ACKNOWLEDGMENTS
We thank J. Ipser, S. Tremaine and D. Eardly for useful discussions. This research was
supported in part by DOE grant DE-FG05-86ER40272 at the University of Florida, DOE
grant DE-AC02-76ER01545 at Ohio State and the DOE and NASA grant NAGW-2381 at
Fermilab.
26
REFERENCES
[1] For dark matter reviews, see e.g., V. Trimble, Ann. Rev. Astron. Astrophys.25 (1987)
425; E.W. Kolb and M.S. Turner, The Early Universe, Addison-Wesley, 1988; M. Sred-
nicki, Editor Particle Physics and Cosmology: Dark Matter, North-Holland, 1990.
[2] J. Ellis, J. S. Hagelin, D. V. Nanopoulos, K. A. Olive, and M. Srednicki, Nucl. Phys. B
238, 453 (1984).
[3] R. D. Peccei and H. Quinn, Phys. Rev. Lett. 38, 1440 (1977); Phys. Rev. D 16, 1791
(1977); S. Weinberg, Phys. Rev. Lett. 40, 223 (1978); F. Wilczek, ibid. 40, 279 (1978).
[4] For reviews of the cosmological and astrophysical properties of axions, see: M. Turner,
Phys. Rep. 197, 67 (1990); G. Raffelt, Phys. Rep. 198, 1 (1990).
[5] J. R. Primack, B. Sadoulet, and D. Seckel, Ann. Rev. Nucl. Part. Sci. B38, 751 (1988);
P. F. Smith and J. D. Levin, Phys. Rep. C 187, 203 (1990).
[6] P. Sikivie, Phys. Rev. Lett. 51 (1983) 1415 and Phys. Rev. D32 (1985) 2988; L.Krauss,
J. Moody, F. Wilczek and D. Morris, Phys.Rev. Lett. 55 (1985) 1797; S. DePanfilis et
al., Phys. Rev. Lett. 59 (1987) 839 and Phys. Rev. D40 (1989) 3151; C. Hagmann et
al., Phys. Rev. D42 (1990) 1297.
[7] K. van Bibber et al.,Int. J. Mod. Phys. D3, Suppl. 1994 33-42.
[8] D. Lynden-Bell, Mon. Not. R. Astron. Soc. 136, 101 (1967).
[9] J. R. Ipser and P. Sikivie, Phys. Lett. B 291, 288 (1992).
[10] J.E. Gunn and J.R. Gott, Ap. J. 176 (1972) 1; J.R. Gott, Ap. J. 201 (1975) 296; J.E.
Gunn, Ap. J. 218 (1977) 592; C. Pryor and M. Lecar, Ap. J. 269 (1983) 513.
[11] J. A. Filmore and P. Goldreich, Ap. J. 281, 1 (1984).
[12] E. Bertschinger, Ap. J. Suppl. 58, 39 (1985).
[13] P. Sikivie, I. I. Tkachev and Yun Wang, Phys. Rev. Lett. 75, 2911 (1995).
[14] C. J. Hogan and M. J. Rees, Phys. Lett. B 205, 228 (1988).
[15] E. Kolb and I. I. Tkachev, Phys. Rev. Lett. 71, 3051 (1993); Phys. Rev D 49, 5040
(1994); Phys. Rev D 50, 769 (1994).
[16] E. Kolb and I. I. Tkachev, astro-ph/9510043; Ap. J. 460, L25 (1996).
[17] A. G. Doroshkevich, Astrophysics 6, 320 (1970); P. J. E. Peebles, Ap. J. 277, 470 (1984);
Y. Hoffmann and J. Shaham, Ap. J. 297, 16 (1985).
[18] J. M. Bardeen, J. R. Bond, N. Kaiser and A. S. Szalay, Ap. J. 304, 15 (1986).
[19] J. Dubinsky and R.G. Carlberg, Ap. J. 378 496 (1991); S. White and D. Zaritsky, Ap.
J. 394, 1 (1992).
[20] S. White, private communication.
27
[21] V.C. Rubin and W.K. Ford, Ap. J. 159, 379 (1970); D.H. Rogstad and G.S. Shostak,
Ap. J. 176, 315 (1972); V.C. Rubin, W.K. Ford and N. Thonnard, Ap. J. 238, 471
(1980).
28
TABLE I. ν 2 (ǫ), Rh (in units of Mpc) and M h (in units of M⊙ ) for different values of ǫ.
TABLES
ǫ
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
ν 2 (ǫ)
0.25
0.6
0.9
1.15
1.35
1.5
1.7
1.8
Rh
2.6
1.7
1.4
1.23
1.14
1.08
1.02
0.98
M h
1.1 × 1014
3.2 × 1013
1.8 × 1013
1.2 × 1013
9.5 × 1012
8.1 × 1012
6.8 × 1012
6.1 × 1012
TABLE II. Density fractions fn and kinetic energies per unit particle mass En of the first five
incoming peaks for various values of ǫ, h and the average dimensionless angular momentum ¯j . Also
shown are the current turn-around radius R in units of Mpc, the effective core radius b in kpc,
and the local density ρ in units of 10−25 g cm−3 . The fn are in percent and the En are in units of
0.5 × (300 km s−1 )2 .
¯j
R
h
ǫ
0.2
0.0
0.7
2.0
0.9
0.7
0.0
1.0
2.4
0.7
0.2
0.15
2.0
0.7
0.1
0.2
”
0.2
0.7
2.0
2.8
0.5
”
”
1.6
0.9
”
”
”
0.4
0.7
2.0
1.8
0.7
0.2
0.25
0.4
0.2
0.7
1.5
f5 (E5 )
1.9 (2.2)
0.7 (2.8)
4.0 (1.3)
2.4 (2.0)
3.6 (1.6)
3.0 (1.7)
3.6 (1.5)
2.6 ( 0.9)
3.1 (1.8)
2.1 (2.5)
f2 (E2 )
5.3 (3.2)
1.1 (3.2)
5.4 (2.3)
7.2 (3.0)
4.1 (2.6)
2.5 (2.7)
5.3 (2.5)
1.6 (1.8)
2.9 (2.8)
1.5 (3.4)
f3 (E3 )
3.3 (2.7)
0.9 (3.0)
5.3 (1.8)
4.9 (2.5)
4.3 (2.1)
2.8 (2.3)
5.1 (2.0)
2.1 (1.4)
3.3 (2.4)
1.8 (3.0)
b
0.0
0.0
13
4.5
12
17
9.3
40
8.5
2.2
ρ
8.1
8.4
5.0
7.6
5.4
4.9
6.0
2.6
5.5
7.7
f1 (E1 )
13 (4.0)
1.6 (3.4)
4.0 (3.1)
7.4 (3.8)
3.1 (3.4)
1.9 (3.5)
4.4 (3.2)
0.8 (2.5)
2.0 (3.5)
1.1 (4.0)
f4 (E4 )
2.4 (2.4)
0.8 (2.9)
4.9 (1.5)
3.2 (2.2)
4.1 (1.8)
2.9 (2.0)
4.5 (1.7)
2.4 (1.1)
3.4 (2.1)
1.9 (2.8)
29
TABLE III. Values of jn , En and ρn given by Eqs.(5.14, 5.17) for the case ǫ = 0.2, ¯j = 0.2,
h = 0.7
n
1
2
3
4
5
6
7
8
jn
0.05
0.11
0.17
0.22
0.26
0.30
0.34
0.38
ρn (10−26 g cm−3 )
1.7
2.5
2.7
2.5
2.2
1.9
1.6
1.4
En ( 1
2 (300 km s−1 )2 )
4.9
3.8
3.3
2.9
2.6
2.4
2.2
2.0
30
FIGURES
FIG. 1. Phase space distribution of the halo dark matter particles at a fixed moment of time.
The solid lines represent occupied phase-space cells. The dotted line corresponds to the observer
position. Each intersection of the solid and dotted lines produces a velocity peak.
FIG. 2. Positions in physical space at successive moments in time t1 < t2 < . . . < t6 of the
particles on a turnaround sphere that is intially rotating rigidly about the vertical axis. The *
indicates the appearance of an inner caustic ring.
31
FIG. 3. A small scale sub-clump falling into the galaxy for the first time.
FIG. 4. The expected value of the ǫ parameter as a function of galaxy size, in models of structure
formation based upon cold dark matter and a flat (Harrison-Zel’dovich) spectrum of primordial
density perturbations. We defined h0.7 ≡ h/0.7
32
FIG. 5. The phase space distribution of halo dark matter particles at a fixed moment of time
for the case ǫ = 0.2 and j = 0. The solid lines represent occupied phase space cells. The dotted
line corresponds to the Sun’s position if h = 0.7.
FIG. 6. The phase space distribution of the dark matter particles in the case ǫ = 0.2, h = 0.7
and a single value of angular momentum j = 0.2
33
FIG. 7. The function λ(τ ) for ǫ = 0.2, j = 0.2.
FIG. 8. Rotational velocity squared curves for different values of ǫ and j = 0
34
FIG. 9. The density fractions fn = ρn/ρ (in percent) and the kinetic energies per unit particle
mass En of the peaks at the Sun’s position, for ǫ = 0.2, j = 0 and h = 0.7. The peaks form pairs.
One member of each pair is due to particles with positive radial velocity and the other is due to
particles with negative radial velocity
FIG. 10. Density profile for the case ǫ = 0.2, h = 0.7 and a single value of angular momentum
j = 0.2.
35
FIG. 11. Rotational curves for the case ǫ = 0.2, with and without angular momentum.
FIG. 12. The same as Fig. 9 but ǫ = 0.2, h = 0.7 and ¯j = 0.2.
36
FIG. 13. The same as Fig. 9 but ǫ = 0.2, h = 0.7 and ¯j = 0.4.
FIG. 14. The same as Fig. 9 but for ǫ = 0.15, h = 0.7 and ¯j = 0.2.
37
FIG. 15. The same as Fig. 9 but ǫ = 1, h = 0.7 and ¯j = 0.2.
FIG. 16. The same as Fig. 9 but including the contribution of baryons to the galactic gravita-
tional potential. ǫ = 0.2 and h = 0.7 in all cases. The peaks are shown explicitly for ¯j = 0.1. The
dashed lines go through the tops of the peaks for the cases ¯j = 0.2 and ¯j = 0.4 while the dotted
line corresponds to j = 0.
38
FIG. 17. The same as Fig. 16 but for ǫ = 0.4.
FIG. 18. The function F (u) defined in Eq. 5.18.
39
|
astro-ph/9804021 | 1 | 9804 | 1998-04-02T16:02:33 | Star Formation in Dwarf Galaxies | [
"astro-ph"
] | We explore mechanisms for the regulation of star formation in dwarf galaxies. We concentrate primarily on a sample in the Virgo cluster, which has HI and blue total photometry, for which we collected H$\alpha$ data at the Wise Observatory. We find that dwarf galaxies do not show the tight correlation of the surface brightness of H$\alpha$ (a star formation indicator) with the HI surface density, or with the ratio of this density to a dynamical timescale, as found for large disk or starburst galaxies. On the other hand, we find the strongest correlation to be with the average blue surface brightness, indicating the presence of a mechanism regulating the star formation by the older (up to 1 Gyr) stellar population if present, or by the stellar population already formed in the present burst. | astro-ph | astro-ph | Star Formation in Dwarf Galaxies
Noah Brosch1
Space Telescope Science Institute
3700 San Martin Drive
Baltimore MD 21218, U.S.A.
and
Ana Heller and Elchanan Almoznino
The Wise Observatory and the School of Physics and Astronomy
Tel Aviv University, Tel Aviv 69978, Israel
To be published in the Astrophysical Journal
Received
;
accepted
8
9
9
1
r
p
A
2
1
v
1
2
0
4
0
8
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1On sabbatical
leave from the Wise Observatory and the School of Physics and
Astronomy, Raymond and Beverly Sackler Faculty of Exact Sciences, Tel Aviv University,
Tel Aviv 69978, Israel
– 2 –
ABSTRACT
We explore mechanisms for the regulation of star formation in dwarf
galaxies. We concentrate primarily on a sample in the Virgo cluster, which
has HI and blue total photometry, for which we collected Hα data at the Wise
Observatory. We find that dwarf galaxies do not show the tight correlation of
the surface brightness of Hα (a star formation indicator) with the HI surface
density, or with the ratio of this density to a dynamical timescale, as found
for large disk or starburst galaxies. On the other hand, we find the strongest
correlation to be with the average blue surface brightness, indicating the
presence of a mechanism regulating the star formation by the older (up to 1
Gyr) stellar population if present, or by the stellar population already formed
in the present burst.
Subject headings: galaxies: irregular -galaxies: stellar content - HII regions -
stars: formation
– 3 –
1.
Introduction
The star formation (SF) is a fundamental process in the evolution of galaxies and
is far from being well understood. The SF is usually characterized by the initial mass
function (IMF) and the total SF rate (SFR), which depends on many factors such as the
density of the interstellar gas, its morphology, its metallicity, etc. According to Larson
(1986), four major factors drive star formation in galaxies:
large scale gravitational
instabilities, cloud compression by density waves, compression in a rotating galactic
disk due to shear forces, and random cloud collisions. In galaxies with previous stellar
generations additional SF triggers exist, such as shock waves from stellar winds and
supernova explosions. In dense environments, such as clusters of galaxies and compact
groups, tidal interactions, collisions with other galaxies, ISM stripping, and cooling
flow accretion probably play some role in triggering the SF process. The triggering
mechanisms were reviewed recently by Elmegreen (1998).
While "global" phenomena, such as the first two SF triggers of Larson (1987),
play a large part in grand design spirals, random collisions of interstellar clouds may
provide the best explanation for dwarf galaxies with bursts of SF. Due to their small
size, lack of strong spiral pattern, and sometimes solid-body rotation (e.g., Martimbeau
et al. 1994, Blok & McGaugh 1997), the star formation in dwarf galaxies is not
triggered by compression due to gravitational density waves or by disk shear. Therefore,
understanding SF in dwarf galaxies should be simpler than in other types of galaxies.
The characterization of the SF processes by a star formation rate (SFR) controlled
by the interstellar gas density as a power law was first introduced by Schmidt (1959).
The volume density of young stars, ρ∗, is related to the volume density of HI gas in the
Galactic disk as ρ∗ = a ρn
gas, where n is a constant, probably ≈2 for spiral galaxies. In
other galaxies the convention is to express the quantities as projected densities of stars
(Σ∗) and of gas, as actually observed: Σ∗ = A Σn
gas. This is usually studied by correlating
– 4 –
the surface density of a young star tracer, such as the Hα surface brightness, with the
gas column density.
The Hα emission from a galaxy measures its ongoing SFR (Kennicutt 1983).
Gallagher et al. (1984) derived an analytic relation between the detected Hα flux and
the present SFR of a galaxy; similar relations were derived by Kennicutt et al. (1994).
The blue luminosity of a galaxy, on the other hand, measures its past star formation
integrated over the last ∼ 109 yrs (Gallagher et al. 1984). The newly formed stars, of
which the more massive produce the Lyman continuum photons which ionize hydrogen
and produce the Hα emission, contribute also to the blue light output of a galaxy. This
contribution is minor in comparison to that from the stars already existing in a galaxy,
unless the SF event is the first in the history of the galaxy or the star burst is unusually
strong. Interestingly, Tresse & Maddox (1998) found recently that the Hα luminosity of
a galaxy correlates with its blue absolute magnitude.
Kennicutt (1998) found that his parametrization of the Schmidt law fitted well the
SF pattern of spiral and IR-selected starburst galaxies. An alternative to the Schmidt
law, proposed by Silk (1997), fitted equally well. In this variant, the SFR per unit
area scales with the ratio of the gas surface density to the local dynamical timescale:
ΣSF R ∝ Σgas
τdyn
∝ Σgas × Ωgas, where Ωgas is the angular rotation speed and the scenario fits
disk configurations. Kennicutt (1998) adopted Ωgas = V (R)
R , where V(R) is the rotation
velocity of the gas at a distance R.
Hunter et al. (1998) tested a set of SF predictors on two small samples of dwarf
galaxies, one measured by them and another derived from de Blok (1997). They found
that the ratio of Σgas to the critical density for the appearance of ring instabilities did not
correlate with the star formation, but that the stellar surface brightness did. From this,
they concluded that possibly the stellar energy input provides the feedback mechanism
for star formation.
– 5 –
We concentrate on a sample of late-type dwarf galaxies in the Virgo cluster (VC).
The reason for selecting dwarfs was to limit the number of possible trigger mechanisms
of SF; these objects are devoid of large-scale SF triggers, as explained above. Having
only VC members limits the sample to a well-defined galaxy background; in addition,
all objects are at ∼the same distance and have HI information from the same source.
We tested for correlations between the Hα emission and other observed quantities, in
order to investigate mechanisms which regulate SF in dwarf galaxies. The justification
to correlate the Hα SFR index against Σgas is the finding of Kennicutt (1998) that a
Schmidt-type law seems to fit large galaxies. If the SFR depends on the gas density
ratioed to the dynamical timescale (Silk 1997, Kennicutt 1998), a correlation with
Σgas × Ωgas is expected. Finally, if the SFR depends on the local population of blue
stars, as found by Hunter et al. (1998), then a dependence on the average blue surface
brightness is expected. We also tested the SFR against the ISM gas velocity, represented
by the width of the HI line profile at 20% of its peak intensity (σ(HI)), and against a
combination of it and the surface density of HI, in a manner similar to that suggested by
Silk (1997).
2. The sample
Our sample consists of 52 late-type dwarf galaxies in the VC selected from Binggeli
et al. (1985, VCC), with HI measurements from Hoffman et al. (1987, 1989). The
sample was constructed in order to enable the detection of weak dependencies of the
star formation properties on hydrogen content and surface brightness. We selected two
sub-samples by surface brightness; one represents a high surface brightness (HSB) group
and is either BCD or anything+BCD, and another represents a low surface brightness
(LSB) sample and includes only ImIV or ImV galaxies. The morphological classification,
which bins the dwarf galaxies in the HSB or LSB groups, is exclusively from the VCC. In
– 6 –
addition, the galaxies are binned by their HI flux integral (FI) from Hoffman et al. (1987,
1989). The HSB sub-sample was selected with galaxies of high HI content (FI>1500 mJy
km s−1) or with low HI content (0<FI<600 mJy km s−1) and is described in Almoznino
& Brosch (1998, AB98). The LSB sub-sample has FI>1000 for the high HI sample or
0<FI<700 mJy km s−1 for the low HI sample (described in Heller et al. 1998, HAB98).
The LSB sub-sample is complete, in the sense that it contains all objects classified ImIV
or ImV in the VCC with mB <17.2 mag. The HSB sub-sample contains 45% of the VCC
galaxies of this type with mB <17.2. Although not complete, it is representative of this
type of object in the VC.
The galaxies were observed at the Wise Observatory (WO) in Mizpe-Ramon from
1990 to 1997, with CCD imaging through the B, V, R, and I broad bands, and narrow
Hα bandpasses in the rest frame of each galaxy. The discussion of all observations and
their interpretation is the subject of other papers (AB98, HAB98). We restrict the
discussion here to the analysis of the integrated Hα flux F(Hα), as it reflects on the
global process of SF. In particular, we concentrate on correlations of this SFR index with
other parameters collected from the literature.
We compare our results with other dwarf galaxies for which we collected published
data. We selected Case galaxies from Salzer et al. (1995, S95) of types HIIH, DHIIH,
BCD, MagIrr, and GIrr, as most similar to our VC sample. We further required that
HI observations would exist for the Case galaxies, and collected seven such objects. As
these galaxies do not have total Hα fluxes listed, we estimated those from (a) the total
blue magnitude mB, (b) the listed equivalent width of the Hβ line, and (c) by assuming
Hα
Hβ =2.9 (Case B, with no extinction). A second comparison sample of eight galaxies
originates from Martin (1997), where each object has an average Hα surface brightness
measure in the 1" wide slit. FI and σ(HI) values were collected from Huchtmeier &
Richter (1989), while total blue magnitudes and sizes originate from NED. No corrections
– 7 –
for Galactic or internal extinction were applied to the data. We also assumed that the
Balmer emission observed spectroscopically is representative of the entire galaxy.
We prefer to use here distance-independent measures, which are not sensitive to the
exact location of a galaxy in the VC, to the value of H0, or to deviations from a smooth
Hubble flow, and to stick, as much as possible, to directly observable quantities. The
observables F(Hα), FI, and mB have, therefore, been normalized to the optical area of
each galaxy, yielding "surface brightness" measures per square arcmin. We calculated
average blue surface magnitudes Σ(B), average HI flux integrals per unit surface Σ(HI),
and average Hα surface brightnesses Σ(Hα) for all objects. The optical area of a galaxy
is defined here as A= πD2
4R , with D the major axis in arcmin and R the axial ratio listed
in VCC or estimated from the image of the object on the Digitized Sky Survey, to yield
Σ(HI).
We used in some correlations σ(HI), and derived Ω = σ(HI)
D
as a representative
gas dynamical property at the outermost optical radius. This definition of Ω is not
purely equivalent to that used by Kennicutt (1998), but it does not require cosmological
assumption in its derivation. We caution at this point that Σ(HI) may overestimate
the surface density of HI in cases where the hydrogen distribution extends beyond the
opical area of an object. Cases where the HI distribution was 3× and more larger than
the optical size of a galaxy were reported by Taylor et al. (1995). However, while very
extended HI distributions do exist, they are not a general characteristic of dwarf galaxies.
Hoffman (private communication) found that only two of the five Virgo cluster BCDs
mapped at Arecibo showed evidence for being extended. In most cases, the Arecibo
beam will cover more than 3× the optical size of one of our objects, implying that
not much HI could have been missed in the measurements we use here. In absence of
synthesis or multi-beam mapping of the HI distributions, we selected to use the coarse
measure of Σ(HI) as defined here, with all caveats mentioned.
– 8 –
The WO sample ranges over more than two orders of magnitude in Σ(HI), over
more than three orders of magnitude in Σ(Hα), and over slightly less that two orders of
magnitude in Σ(B). The comparison sample from Salzer et al. (1995) is more restricted
in the range of Hα, while the galaxies from Martin (1997) have more intense Hα than
the WO objects. In general, galaxies from Salzer et al. are ∼ 2× more distant than the
VC sample, while objects from Martin are ∼ 3× closer than the VC.
3. Star formation correlations
We checked first correlations between global parameters of our dwarf galaxy sample,
such as total blue brightness, total HI content, etc. In all correlations we considered only
detected quantities (no upper limits were included). We did not find that mB and the HI
FIs were correlated in any of the subsamples (for the entire WO sample the correlation
coefficient was 0.57, F=17.4). This scatterplot is shown in the top left panel of Figure
1. Note that some degree of correlation would be expected only from the distance effect,
with both mB and FI being lower for more distant objects.
The plot of Σ(B) vs. Σ(HI), shown in the top right panel of Fig. 1, indicates that
galaxies with more HI per unit area tend also to have higher blue surface brightness, i.e.,
a higher past-averaged SFR, but this correlation was not very significant. We found that
logΣ(Hα) correlates with the HI line width (correlation coefficient 0.61, F=21.3) and
show this in the left middle panel of Fig. 1. Dwarf galaxies with brighter blue surface
brightness tend also to have wider HI profiles (correlation coefficient 0.51, F=13.6), as
the right middle panel of Fig. 1 shows. Σ(HI) correlates also weakly with σ(HI). This is
illustrated in the lower left panel of Fig. 1.
Kennicutt (1998) showed that in a sample of large spirals and starburst galaxies
the average Hα disk surface brightness correlates well with the average molecular and
– 9 –
atomic gas surface density. Dwarf galaxies have very small quantities of molecular gas
(e.g., Gondhalekar et al. 1998), therefore using here Σ(HI) should represent well the
total ISM. This correlation, shown in the lower right panel of Fig. 1, was also weak, and
the combined WO sample had a correlation coefficient of only 0.52 (F=13.1)
A better correlation was found for Σ(Hα) vs. the "Silk"-type parameter Σ(HI)Ω.
Figure 2 shows this for the two VC samples (HSB=filled diamonds, LSB=squares), as
well as for the comparison samples from Salzer et al. (1995; triangles) and Martin (1997;
filled circles). Note that the two VC samples join up nicely, with the HSB galaxies being
brighter and more Hα-intense than the LSB objects. The correlation coefficient for
the combined WO sample is 0.70 (F=34.2) and the slope is 0.93±0.16. The logΣ(Hα)
correlates even better with the blue surface magnitude, as Figure 3 shows. For the
combined WO sample the correlation coefficient is 0.77 and the slope is –0.63±0.09
(F=51.2).
The Salzer et al. (1995) and the Martin (1997) galaxies deviate in both plots
from the trend set by the VC sample. Some of the discrepancy may be the result of
our samples being measured in a uniform and consistent manner, whereas the plotted
parameters for the comparison samples were calculated from published data and some
assumptions (explained above). The Martin galaxies appear consistently above the
location of the WO galaxies; it is probable that their total Hα flux was over-estimated
by assuming that the slit average is representative of the entire galaxy. This is confirmed
for the three objects in common with Marlowe et al. (1997), which have consistently
lower total Hα fluxes than adopted by us here. The Salzer et al. (1995) objects are
generally below the WO objects. They have significant extinction (< cβ >≈ 0.77), which
translates into an under-estimate of the Hα emission when scaling from the Hβ flux. In
addition, the Hβ fluxes were not corrected for underlying absorption; this also causes an
Hα under-estimate. Other reasons for discrepancies may be the different distances to the
– 10 –
two comparison samples, which influence Ω we use here through the angular diameter of
a galaxy, used in the present derivation.
4. Discussion
We mentioned above a number of triggers of star formation. Some, such as shear
and two-fluid instabilities, or spiral density waves, are important mainly in large disk
galaxies and thus are not relevant for dwarfs. The sample studied here is comprised
of fairly isolated galaxies, although this was not a selection criterion. The galaxies are
distant enough from other objects to discount recent (few 107 yrs) interactions as possible
star formation triggers. In these dwarfs the expectation is that the SF may be regulated
only by the gas density, or by the gas density combined with some factor connected
with the stellar content of the galaxy. We checked here various correlations of the star
formation indicator Σ(Hα) with global or specific (per unit area) galaxy parameters.
The "expected" correlations, observed by Kennicutt (1998) to fit well spiral galaxies,
were found to be much weaker in dwarfs. The strongest correlation was with Σ(B), while
the local ISM dynamic indicator Σ(HI)Ω showed the second strongest correlation.
The correlation found for the CFRS survey galaxies (<z>≃0.2, Tresse & Maddox
1998), between the global MB and log L(Hα), can be understood if that survey selected
preferentially galaxies of similar sizes in blue and in hydrogen emission, reducing the
problem to a correlation between area-normalized quantities. These galaxies are much
brighter (MB ≥–21 mag) than the dwarfs discussed here and, being selected on the basis
of their I-band emission, are probably not representative of the "star-forming dwarfs"
class.
Our findings support a scenario whereby the star formation is not controlled by the
gas volume density, by its surface density, or by the ratio of the gas surface density to a
– 11 –
local dynamical timescale. The strongest correlation, based on the correlation coefficient
and the value of the F-statistic, was with the average blue surface magnitude, as found
also by Hunter et al. (1998). There the question was posed whether this was an effect
of the SFR being ∼constant over the last ∼1 Gyr. We can definitely rule out this
possibility, as at least one of our objects (VCC 144; Brosch et al. 1998) seems to exhibit
its first SF burst. Many other galaxies, mainly from the LSB sample, show a number of
small HII regions indicating localized star formation at present. The colors (AB98) are
best fitted by (at least) two stellar populations formed in short bursts, spaced a few 100
Mys to 1 Gyr apart. This indicates that a constant SF is not a serious possibility for the
dwarf galaxies studied here.
5. Conclusions
We tested correlations among parameters related to star formation, gas and stellar
content, and internal dynamics on a sample of dwarf galaxies in the Virgo cluster. We
found that both the Schmidt law and the more recent relation derived by Silk (1997)
do not fit these galaxies as well as they do spirals (Kennicutt 1998). The strongest
correlation of the Hα surface brightness, which measures the present star formation
strength, was with the average blue surface brightness, supporting the proposition of
Hunter et al. (1998) that a feedback mechanism must be at work to regulate the present
SF by the older stellar population.
Acknowledgements
EA is supported by a special grant from the Ministry of Science and the Arts
to develop TAUVEX, a UV imaging experiment. AH acknowledges support from the
US-Israel Binational Science Foundation. NB is grateful for continued support of the
– 12 –
Austrian Friends of Tel Aviv University. Astronomical research at Tel Aviv University is
partly supported by a Center of Excellence award from the Israel Academy of Sciences.
We acknowledge Bruno Binggeli for an updated catalog of the Virgo Cluster and G. Lyle
Hoffman for additional HI information on Virgo galaxies and for comments on a draft
of this paper. This research has made use of the NASA/IPAC Extragalactic Database
(NED) which is operated by the Jet Propulsion Laboratory, California Institute of
Technology, under contract with the National Aeronautics and Space Administration.
We acknowledge some constructive remarks, which improved the presentation, by the
anonymous referee.
– 13 –
References
Almoznino, E. & Brosch, N. 1998, MNRAS, preprint (AB98).
Brosch, N., Almoznino, E. & Hoffman, G.L. 1998, A&A, 331, 873.
Binggeli, B., Sandage, A. & Tamman, G.A. 1985, A. J., 90, 1681.
de Blok, E. 1997, PhD thesis, Rijksuniversiteit Groningen.
de Blok, E. & McGaugh, S.S. 1997, MNRAS, 290, 533.
Gallagher, J.S., Hunter, D.A. & Tutukov, A.V. 1984, ApJ, 284, 544.
Gondhalekar, P.M., Johansson, L.E.B., Brosch, N., Glass, I. & Brinks, E. 1998, A&A,
in press.
Elmegreen, B.G. 1998, in Origins of Galaxies, Stars, Planets and Life (C.E. Woodward,
H.A. Thronson, & M. Shull, eds.), ASP series, in press.
Heller, A., Almoznino, E. & Brosch, N. 1998, in preparation (HAB98).
Hoffman, G.L., Helou, G., Salpeter, E.E., Glosson, J. & Sandage, A. 1987, ApJS, 63,
247.
Hoffman, G.L., Lewis, B.M., Helou, G., Salpeter, E.E. & Williams, H.L. 1989 ApJS, 69,
65.
Huchtmeier, W.R. & Richter, O.-G. 1989, A General Catalog of HI Observations of
Galaxies, New York: Springer Verlag.
Hunter, D.A., Elmegreen, B.G. & Baker, A.L. 1998, ApJ, 493, 595.
Kennicutt, R.C. 1998, ApJ, in press (astro-ph/9712213).
Kennicutt, R.C. 1983, ApJ, 272, 54.
– 14 –
Kennicutt, R.C., Tamblyn, P. & Congdon, C.W. 1994, ApJ, 435, 22.
Marlowe, A.T., Meurer, G.R. & Heckman, T.M. 1997, ApJS, 112, 285.
Martin, C. 1997, ApJ, 491, 561.
Martimbeau, N., Carignian, C. & Ray, J.-R. 1994, AJ, 107, 543.
Salzer, J.J., Moody, J.W., Rosenberg, J.L., Gregory, S.A. & Newberry, M.V. 1995, AJ,
110, 920.
Schmidt, M. 1959, ApJ, 129, 243.
Silk, J. 1997, ApJ, 481, 703.
Taylor, C.L., Brinks, E., Grashuis, R.M. & Skillman, E.D. 1995, ApJS, 99, 427.
Tresse, L. & Maddox, S.J. 1998, ApJ, in press (astro-ph/9709240).
– 15 –
Figure captions
Figure 1: Scatterplots for various observables measured or calculated for the Virgo
cluster sample of dwarf irregular galaxies. The figure shows the total HI flux
integral vs. the total blue magnitude in the top left panel, the logarithm of the HI
surface flux integral (HISFI) vs. the blue surface magnitude in the top right panel,
the logarithm of the Hα surface brightness (Σ(Hα)) vs. the 20% width of the HI
line in km s−1 (σ(HI))in the middle left panel, the blue surface magnitude vs.
σ(HI) in the middle right panel, the HISFI vs. σ(HI) in the lower left panel, and
Σ(Hα) vs. σ(HI) in the lower right panel. The symbols used are filled diamonds
for the BCD galaxies and squares for the LSBs.
Figure 2: Relation between the SFR per unit area [logΣ(Hα)] and logΣ(HI)Ω, a quantity
proportional to the gas surface density times a global gas-dynamical measure. The
additional symbols with respect to Fig. 1 are filled circles for objects from Martin
(1997) and triangles for galaxies from Salzer et al. (1995).
Figure 3: Relation between the average SFR per unit area logΣ(Hα) and the average
blue surface magnitude. The symbols are as in Fig. 2.
– 16 –
−9.0
−11.0
)
α
Η
(
Σ
−13.0
−15.0
−17.0
3.0
–
1
7
–
4.0
5.0
Σ(ΗΙ)Ω
6.0
7.0
8.0
−9.0
−11.0
)
α
Η
(
Σ
−13.0
−15.0
−17.0
12.0
–
1
8
–
13.0
14.0
Σ(Β)
15.0
16.0
17.0
|
astro-ph/0601542 | 1 | 0601 | 2006-01-24T16:13:24 | Collapse of Positronium and Vacuum Instability | [
"astro-ph",
"hep-ph",
"hep-th",
"quant-ph"
] | A hypercritical value for the magnetic field is determined, which provides the full compensation of the positronium rest mass by the binding energy in the maximum symmetry state and disappearance of the energy gap separating the electron-positron system from the vacuum. The compensation becomes possible owing to the falling to the center phenomenon. The structure of the vacuum is described in terms of strongly localized states of tightly mutually bound (or confined) pairs. Their delocalization for still higher magnetic field, capable of screening its further growth, is discussed. | astro-ph | astro-ph |
,
COLLAPSE OF POSITRONIUM AND VACUUM INSTABILITY
A.E. Shabad a
P.N. Lebedev Physics Institute, Leninsky prospekt 53, Moscow 119991, Russia
V.V. Usovb
Center for Astrophysics, Weizmann Institute of Science, Rehovot 76100, Israel
Abstract.A hypercritical value for the magnetic field is determined, which provides
the full compensation of the positronium rest mass by the binding energy in the
maximum symmetry state and disappearance of the energy gap separating the
electron-positron system from the vacuum. The compensation becomes possible
owing to the falling to the center phenomenon. The structure of the vacuum
is described in terms of strongly localized states of tightly mutually bound (or
confined) pairs. Their delocalization for still higher magnetic field, capable of
screening its further growth, is discussed.
It is accepted that magnetic fields are stable in pure quantum electrody-
namics (QED), and other interactions (weak or strong) or magnetic monopoles
have to be involved to make the magnetized vacuum unstable [1]. In this Talk
(see Ref. [2] for a detailed version) we establish, within the frame of QED, that
there exists a hypercritical value of the magnetic field
B(1)
hpcr ≃
exp(cid:26) π3/2
√α
+ 2CE(cid:27) ≃ 1028B0 ≃ 1042G .
m2
4e
(1)
Here α = e2/4π ≃ 1/137, B0 = m2/e = 4.4 × 1013G, m is the electron
mass. The value (1) is less than the magnetic field ∼ (1047 − 1048) G expected
to be present near superconductive cosmic strings [3] and the one (∼ 1047G)
produced at the beginning of the inflation [4]. The hypercritical magnetic field
leads to the shrinking of the energy gap between an electron and positron
that takes place due to the falling to the center. The latter is caused by the
ultraviolet singularity of the photon propagator (or Coulomb potential) [5] plus
the dimensional reduction in the magnetic field [6] - [8]. We discuss the vacuum
structure around the hypercritical magnetic field and its possible decay under
a further growth of the magnetic field - that may cause its screening.
We rely on the theory of the falling to the center developed in [9] that implies
deviations from the standard quantum theory manifesting themselves when
extremely large electric fields near the singularity become important. In that
theory the singularity in the Schrodinger-like equation yields a singular mea-
sure in the scalar product and hence the geometry of a black-hole-like object.
(Stress, that the geometry is induced: no interaction of gravitational origin is
included.)
ae-mail: [email protected]
be-mail: [email protected]
We proceed from the (3+1)-dimensional Bethe-Salpeter (BS) equation in an
approximation, which is the ladder approximation once the photon propagator
(in the coordinate space) is taken in the Feynman gauge: Dij(x) = gijD(x2),
x2 ≡ x2
0 − x2, gii = (1,−1 − 1 − 1). In an asymptotically strong magnetic field
this equation may be written [2] in the following (1+1)-form covariant under
the Lorentz transformations along the axis 3:
(i−→∂k −
= i8πα Xi=0,3
Pk
2 − m)Θ(t, z)(−i←−∂k −
D(cid:18)t2 − z2 −
P 2
⊥
Pk
2 − m)
(eB)2(cid:19) giiγiΘ(t, z)γi,
(2)
0 − xp
3−xp
where Θ(t, z) is the 4×4 (Ritus transform of) BS amplitude, t = xe
0 and
z = xe
3 are the differences of the coordinates of the electron (e) and positron
(p) along the time x0 and along the magnetic field B = (0, 0, B3 = B). Pk and
P⊥ are projections of the total (generalized) momentum of the positronium onto
the (0,3)- subspace and the (1,2)-subspace. Only two Dirac gamma-matrices,
γ0,3, are involved, ∂k = ∂tγ0 − ∂zγ3, Pk = P0γ0 − P3γ3.
Equation (2) is valid in the domain, where the argument of D is greater than
the electron Larmour radius squared (LB)2 = (eB)−1. When B = ∞, this
domain covers the whole exterior of the light cone z2 − t2 ≥ 0.The argument
of the original photon propagator (xe − xp)2 has proved to be replaced in (2)
by t2 − z2 − (exe
⊥ are the center of
orbit coordinates of the two particles in the transversal plane. Now that after
the dimensional reduction this subspace no longer exists these substitute for
the transversal particle coordinates themselves: exe,p
⊥ are not coordinates, but
quantum numbers of the transverse momenta. The mechanism of replacement
of a coordinate by a quantum number is the same as in [7].
⊥/(eB)2, where exe,p
⊥)2 = t2 − z2 − P 2
⊥ −exp
In deriving equation (2) the expansion over the complete set of Ritus matrix
eigenfunctions [10] was used in [2] that accumulate the dependence on the
transversal spacial and spinorial degrees of freedom. This expansion yields an
infinite set of equations, where different pairs of Landau quantum numbers ne,
np are entangled, Eq. (2) being the equation for the (ne = np = 0)-component
that decouples from this set in the limit B = ∞.
most symmetric Ansatz Θ = IΦ, where I is the unit matrix.
In the ultra-relativistic limit P0 = P3 = P⊥ = 0 equation (2) is solved by the
The ultraviolet singularity on the light cone (x2 = 0) of the free photon
propagator, D(x2) = −(i/4π2)(1/x2), after this expression is used in eq. (2),
leads to falling to the center in the Schrodinger-like differential equation
d2Ψ(s)
ds2 +(cid:18)m2 −
1
4s2(cid:19) Ψ(s) =
−
4α
π
1
s2 Ψ(s),
(eB)−1/2 ≪ s ≤ ∞,
(3)
to which the radial part of equation (2) is reduced in the most symmetrical case,
when the wave function Φ(x) = s−1/2Ψ(s) does not depend on the hyperbolic
angle φ in the space-like region of the two-dimensional Minkowsky space, t =
s sinh φ, z = s cosh φ, s = √z2 − t2.
The solution that decreases at s → ∞ is given by the McDonald function with
ν = i 2pα/π ≃ 0.096 i , oscillating
imaginary index: Ψ(s) = √s Kν(ms),
when s → 0. The falling to the center [11] holds for any positive α.
According to [9], the singular equation (3) should be considered as the gen-
eralized eigenvalue problem with respect to α. The operator in the left-hand
side is self-adjoint provided the standing wave condition is imposed,
Ψ(s)s=LB = 0 ,
(4)
that treats the Larmour radius as the lower edge of the normalization box.
The discrete eigenvalues αn(LB) condense in the limit B = ∞ to become a
continuum of states that form the (rigged) Hilbert space of vectors orthogonal
with the singular measure s−2ds. The latter fact allows to normalize them to
δ-functions and interpret as free particles emitted and absorbed by the singular
center. As long as the Larmour radius is much smaller than the only character-
istic length in eq.(3), electron Compton length, LB ≪ m−1 ≃ 3.9 × 10−11 cm.,
the small-distance asymptotic regime is reached, and nothing "behind the hori-
zon," s < LB, - where the two-dimensional equations (2) and (3) are not valid
- may affect the problem. In this way the existence of the limit of vanishing
regularization length, impossible in the standard theory, is achieved.
For sufficiently large magnetic fields LB becomes so small that the region
where Kν(ms) oscillates gets inside the domain of validity of equation (3). The
value of the magnetic field when this happens for the first node is just (1). The
corresponding Larmour radius is about fourteen orders of magnitude smaller
than m−1 and makes ∼ 10−25 cm. Eq.
(1) tells how large the magnetic
field should be in order that the boundary problem (3), (4) might have a
solution, in other words, that the point P0 = P = 0 might belong to the
spectrum. Therefore, if the magnetic field exceeds the first hypercritical value
the positronium ground state exists c with its rest energy compensated for by
the mass defect.
The ultra-relativistic state Pµ = 0 has the internal structure of what was
called a "confined state", belonging to kinematical domain called "sector III"
in [9], i.e. the one whose wave function behaves as a standing wave combination
of free particles near the lower edge of the normalization box and decreases as
cA relation like (1) is present in [8]. There, however, a different problem is studied and,
correspondingly, a different meaning is attributed to that relation:
it expresses the mass
gained dynamically - in the course of spontaneous breakdown of the chiral invariance in
massless QED - by a massless Fermion in terms of the magnetic field applied to it. Later,
in [12] the authors revised that relation in favor of a different approximation. Supposedly,
the revised relation may be of use in the problem of ultimate magnetic field dealt with here.
exp(−ms) at large distances. The effective "Bohr radius", i.e. the value of
s that provides the maximum to the wave function makes smax = 0.17m−1.
This is certainly much less than the standard Bohr radius (e2m)−1. Taken
at the level of 1/2 of its maximum value, the wave function is concentrated
within the limits 0.006 m−1 < s < 1.1 m−1. But the effective region occupied
by the confined state is still much closer to s = 0, since - in accord with the
aforesaid - the probability density of the confined state is the wave function
squared weighted with the measure s−2ds singular in the origin [9] and is
hence concentrated near the edge of the normalization box s ≃ 10−25cm, and
not in the vicinity of the maximum of the wave function. The electric fields
at such distances are about 1043 Volt/cm. Certainly, there is no evidence that
the standard quantum theory should be valid under such conditions. This
fact encourages the use of a theory that admits deviations from the standard
quantum theory that close to the singularity point.
At B = B(1)
hpcr the total energy and momentum of a positronium in the ground
state are zero. This state is not separated from the vacuum by an energy gap,
and has maximum symmetry in the coordinate and spin space. Hence, it may
be related to the vacuum and responsible for its structure.
At B > B(1)
hpcr the eigenvalues of the BS equation (2) for the total 2-momentum
components P0,3 of the e+-e− system are expected to shift into the space-like
region (we keep P⊥ = 0), whereas for B < B(1)
hpcr the c.m. 2-momentum of the
real pair was, naturally, time-like. If P0,3 6= 0, at least for far space-like Pk,
the situation can be modelled by the same equation as (3), but with the large
negative quantity m2 + P 2
substituted for m2. Then the wave function
would contain two oscillating exponentials for large space-like intervals,
k /4
exp(cid:26)iPk
xe + xp
2 (cid:27) ,
s ≫ (−P 2
k )−1/2,
(5)
P 2
k
4 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
±isvuut(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
exp
m2 +
and two oscillating exponentials exp(±2i ln spα/π) for small ones, s ∼ LB.
In the Lorentz frame, where P0 = 0, P3 6= 0, and with the time arguments
in the two-time BS amplitude equal to one another: xe
0 , the solution
3−xp
oscillates along the magnetic field with respect to the relative coordinate xe
3
3 + xp
coordinate xe
(mutually free particles) and with respect to the c.m.
(vacuum lattice).
0 = xp
3
We are now in the kinematical domain called sector IV, or deconfinement
sector in Refs. [9]. Here the constituents are free at large intervals and near the
singular point s = 0. The wave incoming from infinity is partially reflected,
and partially penetrates to the singular point, the probability of creation of the
delocalized (free) states being determined by the barrier transmission coefficient
[9]. Such states may exist if one succeeds to satisfy e.g. periodic conditions, to
be imposed on the lower and upper boundaries of the normalization volume,
instead of condition (4), appropriate in sector III. The possibility to obey them
is provided again by the falling to the center. Now, the delocalized states in
two-dimensional Minkowsky space correspond to electron and positron that
circle along Larmour orbits with vanishing radii in the plane orthogonal to the
magnetic field and simultaneously perform, when the interval between them is
large, a free motion along the magnetic field. They have magnetic moments
and seem to be capable of screening the magnetic field. This provides the
mechanism that may prevent the classical magnetic field from being larger
than a second hypercritical field, for which the delocalization first appears. No
sooner than the delocalized states are found in our present problem one may
definitely claim the instability of the vacuum with the second hypercritical
magnetic field or - which is the same - the instability of such field under the
pair creation that might provide the mechanism for its diminishing. For the
present, we state that the first hypervalue (1) is such a value of the magnetic
field, the exceeding of which would already cause restructuring of the vacuum.
Acknowledgments
This work was supported by the Russian Foundation for Basic Research (project
no 05-02-17217) and the President of Russia Programme (LSS-1578.2003.2), as
well as by the Israel Science Foundation of the Israel Academy of Sciences and
Humanities.
References
[1] M. Bander and H.R. Rubinstein, Phys.Lett. B 280, 121 (1992); 289, 385
(1992); R.C. Duncan, astro-ph/0002442;
[2] A.E. Shabad and V.V. Usov, hep-th/0512236.
[3] E. Witten, Nucl.Phys. B 249, 557 (1985); E.M. Chudnovsky, G.B. Field,
and D.N. Spergel, Phys.Rev. D 34, 944 (1986).
[4] S. Kawati and A. Kokado, Phys.Rev. D 39, 3612 (1989); D. Grasso and
H.R. Rubinstein, Phys.Rep. 348, 163 (2001).
[5] J.S. Goldstein, Phys.Rev. 91, 1516 (1953); P.I. Fomin, V.P. Gusynin,
V.A. Miransky and Yu.A. Sitenko, Rivista Nuovo Cimento 6, numero 5
(1983); N. Seto, Progr.Theor.Phys.Suppl. 95, 25 (1988).
[6] R. Loudon, Am.J.Phys. 27, 649 (1959); Yu.M. Loskutov and V.V.
Skobelev, Phys.Lett. A 56, 151 (1976); V.N. Oraevskii, A.I. Rez and
V.B. Semikoz, Zh.Eksp.Teor.Fiz. 72, 821 (1977); L.B. Leinson and V.N.
Oraevskii, Sov.J.Nucl.Phys. 42, 254 (1985).
[7] A.E. Shabad and V.V. Usov, Astrophys. and Space Sci. 117, 309 (1985);
ibid 128, 377 (1986).
[8] V.P. Gusynin, V.A.Miransky and I.A. Shovkovy, Phys.Rev. D 52, 4747
(1995); Nucl.Phys. B 462, 249 (1996); C.N. Leung, Y.J. Ng and A.W.
Ackley, Phys.Rev. D 54, 4181 (1996);
[9] A.E. Shabad, hep-th/0403177; J.Phys.A: Math.Gen. 38, 7419 (2005).
[10] V.I. Ritus, Sov.Phys.JETP 48, 778 (1978).
[11] L.D. Landau and E.M. Lifshits, "Quantum M echanics", (Pergamon
Press, Oxford) 1991.
[12] V.P. Gusynin, V.A.Miransky and I.A. Shovkovy,Phys. Rev. Lett. 83,
1291 (1999); Nucl. Phys. B 563, 361 (1999); C.N. Leung and S.-Y. Wang,
hep-ph/0510066
|
0710.1431 | 1 | 0710 | 2007-10-07T16:36:02 | High-resolution polarimetry of Parsamian 21: revealing the structure of an edge-on FU Ori disc | [
"astro-ph"
] | We present the first high spatial resolution near-infrared direct and polarimetric observations of Parsamian 21, obtained with the VLT/NACO instrument. We complemented these measurements with archival infrared observations, such as HST/WFPC2 imaging, HST/NICMOS polarimetry, Spitzer IRAC and MIPS photometry, Spitzer IRS spectroscopy as well as ISO photometry. Our main conclusions are the following: (1) we argue that Parsamian 21 is probably an FU Orionis-type object; (2) Parsamian 21 is not associated with any rich cluster of young stars; (3) our measurements reveal a circumstellar envelope, a polar cavity and an edge-on disc; the disc seems to be geometrically flat and extends from approximately 48 to 360 AU from the star; (4) the SED can be reproduced with a simple model of a circumstellar disc and an envelope; (5) within the framework of an evolutionary sequence of FUors proposed by Green et al. (2006) and Quanz et al. (2007), Parsamian 21 can be classified as an intermediate-aged object. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- 15 (2007)
Printed 2 November 2018
(MN LATEX style file v2.2)
High-resolution polarimetry of Parsamian 21: revealing the
structure of an edge-on FU Ori disc⋆
´A. K´osp´al1†, P. ´Abrah´am1, D. Apai2,3, D. R. Ardila4, C. A. Grady5,
Th. Henning6, A. Juh´asz6, D. W. Miller7 and A. Mo´or1
1Konkoly Observatory of the Hungarian Academy of Sciences, P.O. Box 67, H-1525 Budapest, Hungary
2Steward Observatory, The University of Arizona, 933 N. Cherry Avenue, Tucson, AZ 85721, USA
3NASA Astrobiology Institute
4Spitzer Science Center, California Institute of Technology, Pasadena, CA 91125, USA
5Eureka Scientific and Goddard Space Flight Center, Code 667, Greenbelt, MD 20771, USA
6Max-Planck-Institute for Astronomy, Konigstuhl 17, 69117 Heidelberg, Germany
7Department of Physics and Astronomy, University of Louisville, Louisville, KY 40292, USA
Accepted date. Received date; in original form date
ABSTRACT
We present the first high spatial resolution near-infrared direct and polarimetric ob-
servations of Parsamian 21, obtained with the VLT/NACO instrument. We comple-
mented these measurements with archival infrared observations, such as HST/WFPC2
imaging, HST/NICMOS polarimetry, Spitzer IRAC and MIPS photometry, Spitzer
IRS spectroscopy as well as ISO photometry. Our main conclusions are the follow-
ing: (1) we argue that Parsamian 21 is probably an FU Orionis-type object; (2)
Parsamian 21 is not associated with any rich cluster of young stars; (3) our mea-
surements reveal a circumstellar envelope, a polar cavity and an edge-on disc; the disc
seems to be geometrically flat and extends from approximately 48 to 360 AU from the
star; (4) the SED can be reproduced with a simple model of a circumstellar disc and
an envelope; (5) within the framework of an evolutionary sequence of FUors proposed
by Green et al. (2006) and Quanz et al. (2007), Parsamian 21 can be classified as an
intermediate-aged object.
Key words: stars: circumstellar matter -- stars: pre-main sequence -- stars: individual:
Parsamian 21 -- infrared: stars -- techniques: polarimetric.
7
0
0
2
t
c
O
7
]
h
p
-
o
r
t
s
a
[
1
v
1
3
4
1
.
0
1
7
0
:
v
i
X
r
a
1
INTRODUCTION
The most important process of low-mass star formation is
the accretion of circumstellar material onto the young star.
According to current models, the mass accumulation is time-
dependent with alternating episodes of high and low accre-
tion rates (Vorobyov & Basu 2006; Boley et al. 2006). The
high phase may correspond to FU Orionis objects (FUors), a
unique class of low-mass pre-main sequence stars that have
undergone a major outburst in optical light of 4 mag or more
(Herbig 1977). Currently about 20 objects have been clas-
sified as FU Ori-type (for a list see ´Abrah´am et al. 2004).
Records of the outbursts are not available for all of these
objects: some were classified as FUors since they share many
⋆ The results published in this paper are based on data collected
at the European Southern Observatory in the frame of the pro-
gramme P073.C-0721(A).
† E-mail: [email protected]
spectral properties with the FUor prototypes. According
to the most widely accepted model (Kenyon & Hartmann
1991), the observed spectral energy distribution (SED) is
reproduced by a combination of an accretion disc and an
envelope. The energy of the outburst originates from the
dramatic increase of the accretion rate. On the other hand,
Herbig et al. (2003) favour a model in which the outburst
occurs in an unstable young star rotating near breakup ve-
locity.
Parsamian 21 is a system consisting of a central ob-
ject, HBC 687 (α2000 = 19h 29m 0.s87, δ2000 = 9◦ 38′ 42.′′69),
and an extended nebula first listed in the catalogue of
Parsamian (1965). Though no optical outburst was ever
observed, Parsamian 21 was classified as a FUor on the
basis of its optical spectroscopic and far-infrared proper-
ties (Staude & Neckel 1992). According to Henning et al.
(1998) Parsamian 21 is situated in a molecular cloud named
"Cloud A", very close to the Galactic plane. Indeed, the dis-
tance of 400 pc and the radial velocity of vlsr = +27 ±
2
´A. K´osp´al et al.
15 km s−1 (Staude & Neckel 1992) agree very well with
the respective values for Cloud A (vlsr = +27 ± 4 km
s−1, d = 500 ± 100 pc, Dame & Thaddeus 1985). Re-
cently Quanz et al. (2007) questioned the FUor nature of
Parsamian 21 referring to mid-infrared spectral properties,
which better resemble those of post-AGB stars (this issue
will be discussed in Sect. 4).
The elongated nebulosity around the central source ex-
tends ≈ 1 arcmin to the north, while it is less developed
to the south. It is visible in the optical and the near-
infrared, though the source becomes unresolved at 10.8
and 18.2 µm (Polomski et al. 2005). Polarimetric observa-
tions by Draper et al. (1985) revealed a centrosymmetric
polarization pattern and an elongated band of low po-
larization perpendicular to the axis of the bright nebula.
Bastien & M´enard (1990) interpreted the polarization map
of Parsamian 21 in terms of multiple scattering in flattened,
optically thick structures and derived an inclination angle
of 80-85◦ and a size of 30 × 8 arcsec for this disc-like struc-
ture. The existence of a disc is also supported by the discov-
ery of a short bipolar outflow oriented along the polar axis
of the nebula (Staude & Neckel 1992). Emission at submm
wavelengths was detected by Henning et al. (1998) and
Polomski et al. (2005). The surroundings of Parsamian 21
have been searched for companions several times, but close-
by stars seen in J, H and K-band images proved to be field
stars (Li et al. 1994), and no close-by sources have been
found at longer wavelengths so far (Polomski et al. 2005).
Recently several studies were published on the cir-
cumstellar environment of FUors using high-resolution in-
frared and interferometric techniques (e.g. Green et al.
2006; Malbet et al. 1998, 2005; Millan-Gabet et al. 2006a;
Quanz et al. 2007). Such observations could prove or dis-
prove the basic assumptions of FUor models. In this paper
we present the highest resolution imaging and polarimetric
data yet on Parsamian 21 from the VLT/NACO instrument,
allowing the inspection of the circumstellar material and disc
at spatial scales of ≈ 0.07 arcsec. The edge-on geometry of
Parsamian 21 represent an ideal configuration to separate
the components of the circumstellar environment (disc, en-
velope) and understand their role in the FUor phenomenon.
2 OBSERVATIONS AND DATA REDUCTION
The log of observations presented in this paper is summa-
rized in Table 1. In the following subsections we describe
these measurements in detail.
2.1 VLT/NACO observations
We have acquired ground-based near-infrared imaging and
polarimetric observations in visitor mode using the NACO
instrument mounted on the UT4 of ESO's Very Large
Telescope VLT at Cerro Paranal, Chile on 2004 June
18. NACO consists of the NAOS adaptive optics system
and the CONICA near-infrared camera (Lenzen et al. 1998;
Hartung et al. 2000; Rousset et al. 2003). The weather con-
ditions were excellent, the typical optical seeing was ≈ 0.72
arcsec with two short peaks of 1.3 arcsec during the night.
For imaging we used the visual dichroic and a camera with
13 mas/pixel scale to obtain H (λc = 1.66 µm) and KS
(λc = 2.18 µm) images and a camera with 27 mas/pixel
scale to obtain L′ (λc = 3.8 µm) images. In each filter a
4-point dithering was applied with small-amplitude random
jitter at each location. The adaptive optics configuration was
fine-tuned for each filter to ensure the best possible correc-
tion. The H-band polarimetric observations were obtained
using the 27 mas/pixel scale camera.
VLT/NACO imaging. The data reduction was carried
out using self-developed IDL routines. For the H and KS
band filters lampflats were taken, while for the L′ filter,
skyflats were acquired. For each filter, these images were
combined to a final flatfield, and bad pixel maps were pro-
duced. Then the raw images were flatfield corrected, and bad
pixels were removed. Sky frames were calculated by taking
the median of all images taken with the same filter, then im-
ages were sky subtracted. Individual frames were shifted to
the same position and their median was taken to obtain the
final mosaic. Shifts were calculated by computing the cross-
correlation of the images, which gives a more precise result
in case of extended sources, than a simple Gaussian-fitting to
the peak. Dithering resulted in a final mosaic of 21.4 × 21.4
arcsec in case of H and KS band, and 43.8 × 43.8 arcsec in
L′ band. The photometric standard was S889-E for the H
and KS filters (H = 11.662 ± 0.004 mag, KS = 11.585 ± 0.005
mag, Persson et al. 1998), and HD 205772 for the L′ filter
(L′ = 7.636 ± 0.027 mag, Bouchet et al. 1991).
VLT/NACO polarimetry. Polarimetric
observations
were obtained using the differential polarimetric imaging
technique (DPI, see e.g. Draper et al. 1985; Kuhn et al.
2001; Apai et al. 2004). The basic idea of this method is to
take the difference of two orthogonally polarized, simultane-
ously acquired images of the same object in order to remove
all non-polarized light. As the non-polarized light mainly
comes from the central star, after subtraction only the po-
larized light, such as the scattered light from the circum-
stellar material remains. This standard way of dual-beam
imaging polarimetry combined with adaptive optics at the
VLT is a powerful way of doing high-contrast mapping. We
obtained polarimetric images with NACO through the H fil-
ter, using a Wollaston prism with a 2 arcsec Wollaston mask
to exclude overlapping beams of orthogonal polarization.
Parsamian 21 was observed at four different rotator angles of
0◦, 45◦, 90◦ and 135◦, providing a redundant sampling of the
polarization vectors. At each angle a 3-point dithering was
applied. The polarimetric calibrator was R CrA DC No. 71
(Whittet et al. 1992). The polarimetric data reduction was
done in IDL using previously developed software tools which
we presented in detail in Apai et al. (2004). According to the
NACO User's Manual, instrumental polarization is generally
about 2%. Based on our polarimetric calibration measure-
ments, we found a typical instrumental polarization value of
3% (position angle: 91◦ east of north). Considering the ex-
pected high polarization for the Parsamian 21 nebula, we can
conclude that instrumental polarization can be neglected in
the present case, except in limited areas of low polarization.
2.2 GFP Hα imagery
Parsamian 21 was observed with the Goddard Fabry-Perot
(GFP) interferometer at the Apache Point Observatory 3.5m
High-resolution polarimetry of Parsamian 21
3
Table 1. Log of the observations of Parsamian 21. Top part: data from this paper; Bottom part: archival data.
Instrument
Filter (central λ)
Mode
Date
Exp. Time
Field of View Pixel scale
YY/MM/DD
[arcsec × arcsec]
[arcsec]
VLT/NACO
VLT/NACO
VLT/NACO
VLT/NACO
H (1.66 µm)
KS (2.18 µm)
L′ (3.80 µm)
H (1.66 µm)
Imaging
Imaging
Imaging
Polarimetry
04/06/18
04/06/18
04/06/18
04/06/18 72×10 s, 24×80 s
8×30 s
8×20 s
48×0.2 s
13.6×13.6
13.6×13.6
27.8×27.8
27.8×3.1
GFP
6563 and 6590A
Imaging
07/05/15
900 s
130 circular
HST/NICMOS
HST/WFPC2
2 µm
Polarimetry
F814W (0.80 µm)
Imaging
97/11/12
01/07/30
18×31 s
2×500 s
19.5×19.3
79.5×79.5
0.013
0.013
0.027
0.027
0.38
0.076
0.1
FWHM
[arcsec]
0.07
0.07
0.11
0.12
1.1
0.16
0.17
Spitzer/IRAC 3.6, 4.5, 5.8 and 8.0 µm Imaging
Spitzer/MIPS
Imaging
Spitzer/IRS
04/04/21
04/04/13
Spectroscopy 04/04/18
24 and 70 µm
5 − 40 µm
848 s
318×318
1.2
73 s, 462 s
444×480, 174×192 2.5, 4.0
1.9 -- 2.7
6, 18
1121 s
ISO/ISOPHOT
65 and 100 µm
Imaging
96/09/28
3542 s
360×420
15
44, 47
telescope on 2007 May 15. The observations were made in
direct imaging mode. The on-band image has a central wave-
length of 6563 A (Hα) and a width of 120 km s−1. The off-
band image has a central wavelength of 6590 A and a width
of 120 km s−1. The pixel scale is 0.38 arcsec per pixel. The
observations were made through patchy cirrus and are not
photometric, with seeing near 1 arcsec. The instrument and
data reduction are described in Wassell et al. (2006) and
references therein. For the Hα on-band image, conspicuous
night sky rings contaminate the region near Par 21. In order
to remove these rings, we created a data cube consisting of
subimages centred on the GFP optical axis, rotated by dif-
ferent angles. After excluding the area around Parsamian 21,
we azimuthally medianed the subimages, and subtracted the
resulting image from the original measurement. For the off-
band image, the night sky rings had displaced beyond the
location of the Par 21 nebulosity, thus no ring subtraction
was necessary. The on- and off-band images were shifted and
scaled to match each other by measuring the positions and
fluxes of seven stars in the vicinity of Parsamian 21. The
continuum-subtracted Hα image can be seen in Fig. 3.
2.3 HST archival data
HST/NICMOS polarimetric observations were obtained
on 1997 November 12 using the NIC2 camera in MULTI-
ACCUM mode, and the POL0L, POL120L and POL240L
polarizing filters. These filters have a bandpass between 1.9
and 2.1 µm. Raw data files were calibrated at the STScI
with the calnica v.4.1.1 pipeline. Since these data have not
been published yet, we downloaded the pipeline-processed
files and we did further processing using the IDL-based Po-
larizer Data Analysis Software, which is available through
the STScI website and is described by Mazzuca & Hines
(1999). This software package combines the images obtained
through the three polarizers using an algorithm described
in Hines et al. (2000). We used coefficients appropriate for
the pre-NCS measurements from Hines & Schneider (2006).
A 6-point dither pattern was applied, covering in all about
24 × 34 arcsec. The central star and the inner parts of the
Parsamian 21 nebula can only be seen in two of the six im-
ages, which we shifted and co-added.
HST/WFPC2. Parsamian 21 was observed with the
WFPC2 on 2001 July 30 through the F814W filter.
Parsamian 21 is in the middle of the image on the WF3 chip
(Fig. 1), which has a pixel scale of 0.1 arcsec. Since these
data have not been published yet, we used the High-Level
Science Product with the association name 'U6FC2801B',
available at the HST archive.
2.4 Spitzer archival data
Observations of Parsamian 21 were obtained on 2004 April
13 (MIPS), April 18 (IRS) and April 21 (IRAC). The MIPS
and IRAC, as well as part of the IRS data are still unpub-
lished. The IRAC images cover an area of 5.3 × 5.3 arcmin
centred on Parsamian 21. The 'high-dynamic-range' mode
was used to obtain 15 frames in each position, 5 with 0.4 s
exposure time and 10 with 10.4 s. The frames were processed
with the SSC IRAC Pipeline v14.0 to Basic Calibrated Data
(BCD) level. The BCD frames were then processed using the
IRAC artifact mitigation software, and finally the artifact-
corrected frames were mosaicked with MOPEX. Photometry
for Parsamian 21 was done in IDL. We used the short expo-
sure images, because Parsamian 21 did not saturate these
frames. We used an aperture of 6 arcsec and a sky annu-
lus between 6 and 12 arcsec. The aperture corrections were
1.061, 1.064, 1.067 and 1.089 for the four channels, respec-
tively1. Colour correction was applied by convolving the ob-
served SED with the IRAC filter profiles in an iterative way.
The resulting colour-corrected fluxes can be seen in Table 2.
Photometry for field stars was obtained in the long
exposure images, because most field stars are fainter than
Parsamian 21 and did not saturate even the long exposure
frames. For this purpose we used StarFinder, which is an
IDL-GUI based program for crowded stellar fields analysis
(Diolaiti et al. 2000). PSF-photometry was calculated for all
1 IRAC Data Handbook,
http://ssc.spitzer.caltech.edu/irac/dh/iracdatahandbook3.0.pdf
available
version
3.0,
at
4
´A. K´osp´al et al.
stars having S/N > 3 and aperture corrections correspond-
ing to "infinite" aperture radius were applied (0.944, 0.937,
0.772 and 0.737 for the four channels, respectively1). Then,
we cross-identified the sources in the four bands and found
that 100 field stars are present in all four IRAC images. We
plotted these sources as well as Parsamian 21 on a [3.6]−[4.5]
vs. [5.8]−[8.0] colour-colour diagram in Fig. 10. This plot will
be discussed in Sect. 4.2.
The MIPS 24 and 70 µm images were taken in 'compact
source super resolution' mode. At 24 µm 14 frames (each
with 2.62 s exposure time), at 70 µm 44 frames (each with
10.49 s exposure time) were combined into a final mosaic
using MOPEX. At 24 µm Parsamian 21 saturated the de-
tector, thus we used a model PSF to determine which pixels
are still in the linear regime. Then we fitted the PSF only
to these pixels. The 70 µm image is not saturated, therefore
we simply calculated aperture photometry in IDL using an
aperture radius of 16 arcsec, sky annulus between 39 and 65
arcsec, and an aperture correction of 1.7412. We found that
the source was point-like at both 24 and 70 µm. Colour cor-
rection was applied by convolving the observed SED with
the MIPS filter profiles in an iterative way. The resulting
colour-corrected fluxes can be seen in Table 2.
The IRS spectroscopy was carried out in 'staring mode'
using the Short Low, the Short High and the Long High
modules. Measurements were taken at two different nod po-
sitions and at each nod position three exposures were taken.
We started from BCD level and extracted spectra from the
2D dispersed images using the Spitzer IRS Custom Extrac-
tion software (SPICE). In the case of the Short Low channel
we extracted a spectrum from a wavelength dependent, ta-
pered aperture around the star, and also from a sky position.
In the case of the high-resolution channels, we extracted
spectra from the full slit. For the Short High channel, the
target was off-slit and most of the stellar PSF fell outside
of the slit. We corrected this spectrum for flux loss using
the measured IRS beam profiles (for more details, see the
Appendix). The resulting complete 5 − 35 µm spectrum can
be seen in Figs. 5 and 6.
2.5 ISO archival data
ISO measurements of Parsamian 21 at 65 and 100 µm
were obtained in 'oversampled map' (PHT32) mode on
1996 September 28. These observations were reduced us-
ing a dedicated software package (P32TOOLS) developed
by Tuffs & Gabriel (2003). This tool provides adequate cor-
rection for transients in PHT32 measurements. Absolute cal-
ibration was done by comparing the source flux with the on-
board fine calibration source. In order to extract the flux of
our target, a point spread function centred on Parsamian 21
was fitted to the brightness distribution on each map. At
65 µm the PSF could be well fitted by the ISOPHOT the-
oretical footprint function. At 100 µm, however, the source
turned out to be extended. In this case the brightness dis-
tribution was modelled as the convolution of the standard
theoretical footprint function with a two-dimensional ellip-
tical Gaussian of 40 arcsec × 20 arcsec, and a position angle
2 MIPS Data Handbook,
at
http://ssc.spitzer.caltech.edu/mips/dh/mipsdatahandbook3.2.pdf
available
version
3.2,
Table 2. Photometry for Parsamian 21. All fluxes are colour-
corrected.
Instrument
Wavelength [µm]
Flux [mJy]
Spitzer/IRAC
Spitzer/IRAC
Spitzer/IRAC
Spitzer/IRAC
Spitzer/MIPS
Spitzer/MIPS
ISO/ISOPHOT
ISO/ISOPHOT
3.6
4.5
5.8
8.0
24
70
65
100
143 ± 8
187 ± 10
303 ± 16
829 ± 42
5 530 ± 310
13 300 ± 1 300
12 600 ± 1 300
14 200 ± 1 900
Figure 1. HST/WFPC2 image of Parsamian 21 taken through
the F814W filter. The intensity scale is square root and brightness
increases from light to dark. The white square marks the area
shown in the upper left panel of Fig. 2.
of 114◦. The obtained fluxes were colour corrected by con-
volving the observed SED with the ISOPHOT filter profiles
in an iterative way. The results are displayed in Table 2.
3 RESULTS
3.1 Broad-band imaging
Figure 1 displays the HST/WFPC2 image taken at 0.8 µm.
The image reveals the structure of the Parsamian 21 neb-
ula with an unprecedented spatial resolution and detail. We
note that strong emission lines associated with the bipolar
outflow (see Sect. 3.2) are not included in the bandpass, so
the HST image provides a clean picture of the reflection neb-
ula. The overall dimensions of the nebula are approximately
60 × 20 arcsec. The northern part has a characteristic el-
liptic, loop-like shape, with the inside of the loop having a
relatively low surface brightness. The outer edge of the neb-
ula is rather well-defined, while the inner edge is more fuzzy.
Apart from the loop itself, there is also a wide, faint SE-NW
oriented stripe about 30 arcsec from the star. The nebula is
clearly bipolar, although very asymmetric: the southern part
High-resolution polarimetry of Parsamian 21
5
5
0
RA offset [arcsec]
-5
-10
-15
)
c
e
s
c
r
a
(
t
e
s
f
f
o
C
E
D
5
0
-5
-10
Knot 1
Knot 3
Knot 2
5
0
-5
-10
5
0
RA offset [arcsec]
-5
-10
-15
Figure 3. Continuum-subtracted Hα image of Parsamian 21. The
peak at position (0,0) is the residuum of the central source. Three
Herbig-Haro knots could be identified; these were named Knot 1,
2 and 3 by Staude & Neckel (1992).
Knot 2
Knot 3
Knot 1
y
t
i
s
n
e
t
n
I
1200
1000
800
600
400
200
0
-10
-5
0
5
Angular distance [arcsec]
Figure 4. South-north oriented cut across Parsamian 21 in the
continuum-subtracted Hα image. 0 marks the position of the cen-
tral star. The peaks marked by solid vertical lines at −10.5, −3
and 5 arcsec correspond to Knot 2, 3 and 1, respectively. Dashed
vertical lines mark the positions of the same knots as measured by
Staude & Neckel (1992). Arrows demonstrate the direction where
the knots propagate.
source, we could identify three knots (see Fig. 3): Knots 1,
2 and 3 are located 5 arcsec north, 10.5 arcsec south and
3 arcsec south, respectively. The knots correspond to those
identified by Staude & Neckel (1992), but all three knots
have moved outwards since then (Fig. 4). Assuming a dis-
tance of 400 pc, and supposing that the projected velocities
are close to the velocities of the knots along the polar axis,
one can derive 120 km s−1 for Knots 1 and 3, and 530 km
s−1 for Knot 2. This is in good agreement with the velocities
calculated by Staude & Neckel (1992) using radial velocity
measurements. Such outflow velocities imply kinematic ages
of approximately 80 yr for Knot 1 and 40 yr for Knots 2 and
3.
3.3 Spectral energy distribution
Table 2 contains our Spitzer and ISO photometry for
Parsamian 21. These data, complemented with previously
published measurements between 1983 and 1999 are plot-
ted in Fig. 5, showing the complete UV-to-mm spectral en-
ergy distribution (SED) of the object. Comparing data taken
Figure 2. HST/WFPC2 (with filter F814W) and VLT/NACO
(with filters H, KS and L′) images of Parsamian 21. Circles at
the left bottom corners indicate the FWHM at the centre of the
corresponding image. In the KS-band image, arrows mark the
positions of the two closest stars to Parsamian 21 (see Sect. 4.2).
is much less developed than the northern one. We note that
while the southeastern arc is brighter close to the star (see
the central part of the HST image in Fig. 2), the southwest-
ern arc is more conspicuous farther from the star (Fig. 1).
It is remarkable that field stars are visible both to the north
and to the south of Parsamian 21. Supposing that these stars
are in the background, this implies that there is no signifi-
cant difference in the extinction between the northern and
southern part.
Figure 2 shows the central parts of the HST/WFPC2
and the VLT/NACO images (note that the innermost 0.57
arcsec of the HST image is saturated). The bipolar nature of
the nebula is most conspicuous at the shortest wavelength
(0.8 µm), where all 4 arcs (roughly to the northeast, north-
west, southwest and southeast) are visible. In the KS and H
bands the nebula is more triangle-shaped. In all three NACO
bands the central source is extended, with deconvolved sizes
of 0.20, 0.12 and 0.08 arcsec in the H, KS and L′ bands,
respectively.
3.2 Narrow-band imaging
Figure 3 displays the continuum-subtracted Hα image of
Parsamian 21. Conspicuous Herbig-Haro knots are seen to
the north and the south of the central star. The north-
ern knot is compact, resembling the knot described in
Staude & Neckel (1992). The southern knot, however, has
a distinctly bowed shape. The southern extent of this knot
has a S/N of 10 over background per pixel, while the flank-
ing structures have S/N ≈ 7. The northern knot has S/N
in excess of 15 per pixel. To measure the angular distance
of the knots from the central star in the Hα imagery, we
binned the data in a 3 pixel (1.1 arcsec) wide swath ori-
ented north-south. Apart from the residual of the central
6
´A. K´osp´al et al.
Figure 5. Complete UV-to-mm SED of Parsamian 21. Source of
data: 2MASS All-Sky Catalog of Point Sources, MSX6C Infrared
Point Source Catalog, ´Abrah´am et al. (2004), Polomski et al.
(2005), Neckel & Staude (1984), Henning et al. (1998) and this
work. The model SED is discussed in Sect. 4.4.
m
W
[
ν
]
2
-
F
ν
g
o
L
PAH
-12.2
-12.4
-12.6
-12.8
-13.0
PAH
PAH
CO2 ice
Silicates
14.6 15.0 15.4 15.8
6
8
10
Wavelength [µm]
12
14
16
18
Figure 6. 5 − 18 µm part of the Spitzer/IRS spectrum of
Parsamian 21. The inset displays in details the continuum-
subtracted spectral region around the 15.2 µm CO2 ice feature
of Parsamian 21 with solid line and as a comparison HH 46 IRS
with dashed line (for discussion see Sect. 4.5).
between 1983 and 1999, ´Abrah´am et al. (2004) found that
there is no long-term flux variation in the 1 − 100 µm wave-
length range. The new Spitzer data from 2004 reveal that
the mid-infrared brightness of Parsamian 21 stayed constant
since then at a 15% level. Although Parsamian et al. (1996)
reported 2 − 3 mag brightness variations in the B band be-
tween 1966 and 1990 (in Fig. 5 we plotted optical measure-
ments from 1980 as representative values), no significant flux
changes can be seen at longer wavelengths within the mea-
surement uncertainties.
At wavelengths shorter than 3 µm, the SED is highly
reddened (Polomski et al. 2005), probably due to the com-
bined effect of interstellar extinction and self-shadowing by
circumstellar material. The fact that the central source is
extended at these wavelengths (see Sect. 3.1) indicates that
the photometry is contaminated by scattered light from
the inner part of the circumstellar environment. Between
3 and 25 µm, the SED is rising towards longer wavelengths
as νFν ∝ λ. The 5 − 18 µm Spitzer spectrum (Fig. 6) dis-
plays strong PAH emission features at 6.3, 8.2 and 11.3 µm
and amorphous silicate emission around 10 µm (Quanz et al.
2007, see also the spectrum of Polomski et al. 2005). Weaker
absorption features can also be seen between 7 and 8 µm,
which cannot be safely identified with any ice or PAH
feature. The 10 − 19 µm channel of the Spitzer/IRS spec-
trum was not shown by Quanz et al. (2007) due to data re-
duction difficulties. We made an attempt to evaluate also
this channel (for details, see Appendix). As can be seen
in the inset of Fig. 6, the spectrum shows a strong CO2
ice absorption feature at 15.2 µm, which has not been re-
ported in the literature so far. Between 25 and 100 µm the
SED is flat (νFν ∝ const.), while the submillimetre shape
is νFν ∝ λ−3.5 (corresponding to β = 0.5, assuming optically
thin emission and a dust opacity law of κν ∝ ν β). Using a dis-
tance of 400 pc and an interstellar reddening of AV = 2 mag
(Hillenbrand et al. 1992), the bolometric luminosity com-
puted as the integral of the SED from 0.44 to 1300 µm is
10 L⊙ (assuming isotropic radiation field).
3.4 Polarimetry
Figure 7 shows images of Parsamian 21 calculated from the
H-band NACO polarization data, as well as from the 2 µm
NICMOS measurements. The left column shows the total
intensity (I) obtained as the sum of the intensities of the two
orthogonal polarization states. The middle column displays
the polarized intensity (P I) calculated as the square root
of the quadratic sum of the Q and U Stokes components.
The right column shows the degree of polarization (% pol.),
i.e. the ratio of the polarized to the total intensity.
The NACO total intensity maps (Fig. 7, left) are very
similar to the H-band map in Fig. 2. Thus, the Wollaston-
prism data can reproduce well the direct images taken with-
out the prism. While the total intensity is relatively smooth,
the polarized intensity (Fig. 7, middle) reveals many fine de-
tails not visible in the total intensity map. The most strik-
ing feature is a dark horizontal lane across the star, where
the polarized intensity is very low. Above this dark lane,
high intensity regions can be seen, primarily emphasizing
the walls of the upper lobe. Below the dark lane, the south-
eastern arc is very pronounced. The degree of polarization
(Fig. 7, right) again shows features different from the previ-
ous two maps. Here, the central star almost disappears, as
the stellar light is not polarized. The horizontal dark lane
of low polarization across the star is even more pronounced:
there the degree of polarization is around ≈ 5% in the H-
band NACO images. We note that the polarization pattern
in this area is probably affected by instrumental polarization
(Sec. 2.1), though this effect should not alter our discussion
in Sec. 4.3.2. The polarization of the central source itself,
measured in a 0.08 arcsec radius aperture, is 9% and its po-
sition angle is 70◦ east of north. The long exposure NACO
images (Fig. 7, middle row), showing the outer regions with
higher signal-to-noise ratio, are consistent with the short ex-
posure maps (Fig. 7, upper row). It is remarkable that the
NICMOS images, which represent independent observations
(different resolution, different instrument, different polariza-
tion technique), show strikingly similar features, although
the degree of polarization is slightly higher.
Figure 8 shows the polarization pseudovectors overlaid
on total intensity contours. The NACO map, along with the
degree of polarization map in Fig. 7 right shows that the
High-resolution polarimetry of Parsamian 21
7
I
I
1
0
-1
1
0
-1
1
0
-1
PI
% pol.
VLT/NACO
1.65 µm
1
1
1
1
0
0
0
0
-1
-1
-1
-1
1
1
0
0
-1
-1
1
1
0
0
-1
-1
1
1
0
0
-1
-1
PI
% pol.
VLT/NACO
1.65 µm
1
1
1
1
0
0
0
0
-1
-1
-1
-1
1
0
-1
1
0
-1
1
0
-1
1
I
0
-1
1
PI
0
-1
1
% pol.
0
-1
HST/NICMOS
2.0 µm
1
1
1
1
0
0
0
0
-1
-1
-1
-1
1
0
-1
1
0
-1
1
0
-1
80%
60%
40%
20%
80%
60%
40%
20%
80%
60%
40%
20%
]
"
[
t
e
s
f
f
o
C
E
D
1
0
-1
1
0
-1
1
0
-1
1
0
-1
1
0
-1
1
0
-1
RA offset ["]
Figure 7. Left: total intensity (I); middle: polarized intensity (P I); right: degree of polarization (% pol.), with scalebar. Upper row:
short exposure H-band NACO images; middle row: long exposure H-band NACO images (the saturated central parts are masked out);
lower row: 2 µm NICMOS images of the same area. The intensity scale is logarithmic and brightness increases from dark to light.
horizontal band of low polarization across the star is very
well-confined, narrow (about 0.15 − 0.20 arcsec wide at the
5% polarization level) and can be discerned from 0.12 to
1.2 arcsec to the star (and likely beyond, as indicated by
the long exposure maps, see also the polarization maps of
Draper et al. 1985 and Hajjar et al. 1997). The position an-
gle of this band is 78◦ ± 4◦ east of north, consistent with
that measured by Draper et al. (1985) (75◦ ± 4◦). This im-
plies that the band in our high spatial resolution images are a
direct inward continuation of the band seen by Draper et al.
(1985) and Hajjar et al. (1997) in their lower resolution im-
ages. The degree of polarization in this band is a few percent,
and it is oriented parallel with the band itself.
Most of the northern part of the nebula shows a
centrosymmetric pattern, characteristic of reflection neb-
ulae with single scattering. This is consistent with what
Draper et al. (1985) and Hajjar et al. (1997) measured for
Parsamian 21 itself. It is also similar to what can be seen in
other bipolar nebulae (see e.g. Meakin et al. 2005 for neb-
ulae of young stars or Scarrott et al. 1993 for nebulae of
evolved stars). The highest polarization in the northern part
can be measured in the northeastern wall (25 − 30%) and
also in a spot about 0.9 arcsec west and 0.6 arcsec north
to the star (20 − 25%). The southern part of the nebula
also follows the regular centrosymmetric pattern seen in the
northern part. Here the highest polarization can be found
in the southeastern arc (60 − 70%).
The NICMOS vector map at 2 µm (Fig. 8 bottom) re-
peats all the main features seen by NACO in the H-band
(Fig. 8 top). We note nevertheless the presence of two small,
8
´A. K´osp´al et al.
120
100
80
]
c
e
s
c
r
a
[
t
e
s
f
f
o
C
E
D
60
40
20
0
0
1
0
-1
]
"
[
t
e
s
f
f
o
C
E
D
1
0
-1
VLT/NACO
50%
20
40
60
RA offset [arcsec]
80
100
120
1
0
-1
RA offset ["]
40
1
0
-1
HST/NICMOS
]
"
[
t
e
s
f
f
o
C
E
D
1
0
-1
30
]
c
e
s
c
r
a
[
20
t
e
s
f
f
o
C
E
D
10
0
0
10
RA offset [arcsec]
20
30
1
0
-1
RA offset ["]
50%
1
0
-1
1
0
-1
40
Figure 8. Polarization pseudovectors overlaid on intensity con-
tours. Top: VLT/NACO, H-band, with 5 × 5 pixels binning, bot-
tom: HST/NICMOS, 2 µm, with 2 × 2 pixels binning.
localized depolarization areas on either side of the central
source at positions (0.′′9, 0.′′0) and (−0.′′8, −0.′′1).
4 DISCUSSION
4.1 Parsamian 21: an FU Orionis-type star
no
optical
ever
outburst was
Although
observed,
Parsamian 21 shares many properties characteristic of
FUors. Its spectral type in the literature ranges from A5
to F8 supergiant (e.g. Staude & Neckel 1992), similar to
the typical FUor spectral types (F -- G supergiant). In ad-
dition, it shows strong infrared excess and drives a bipolar
outflow, similarly to many FUors. However,
in a recent
paper by Quanz et al. (2007) the pre-main sequence nature
1.4
1.2
1.0
H
-
J
0.8
0.6
0.4
0.2
Parsamian 21
0.0
0.2
0.4
0.6
H-K
0.8
1.0
Figure 9. 2MASS colour-colour diagram of 188 sources in an area
of 5.3 × 5.3 arcmin centred on Parsamian 21. The main-sequence
and giant branch are marked by solid lines (Koornneef 1983), the
reddening path with dashed lines (Cardelli et al. 1989) and the
T Tauri locus with dotted line (Meyer et al. 1997).
and the FUor status of Parsamian 21 were questioned,
mainly because its PAH emission bands are untypical for
young stars. They suggest that Parsamian 21 is either an
intermediate mass FUor object, or an evolved star sharing
typical properties with post-AGB stars.
An important parameter which may discriminate be-
tween a pre-main sequence object and a post-AGB star is
the luminosity. The typical luminosity of post-AGB stars
is in the order of 103 − 104 L⊙ (Kwok 1993), while FUors
are typically fainter than a few 100 L⊙. The luminosity of
Parsamian 21, 10 L⊙ (see Sect. 3.3), is at least a factor
of 100 lower than that of post-AGB stars. Even adopt-
ing the largest distance estimate mentioned in the liter-
ature (1800 pc, Staude & Neckel 1992), the luminosity of
Parsamian 21 is still too low. The observed proper motion of
the Herbig-Haro knots, however, prefers the lower distance
(400 pc), otherwise the outflow velocities (≈ 2400 km s−1 at
1800 pc) would be unusually high for a young stellar object
(typically up to 600 km s−1, see e.g. Mundt et al. 1987;
Hessman et al. 1991). Moreover, the existence of Herbig-
Haro outflows are usually tracers of low-mass star formation
(Reipurth & Bally 2001).
In the following discussion, we consider Parsamian 21
to be a FUor and we discuss its properties in the context of
young eruptive stars. Nevertheless we note that most of our
results on the morphology and circumstellar structure are
valid regardless of the nature of the central object.
4.2 Parsamian 21: an isolated young star?
Parsamian 21 is situated close to the Galactic plane (l =
45.8◦, b = −3.8◦), in a molecular cloud called Cloud A. This
cloud was identified by Dame & Thaddeus (1985) in their
CO survey of molecular clouds in the northern Milky Way.
The cloud occupies an area of 8 square degrees in the sky,
High-resolution polarimetry of Parsamian 21
9
]
5
.
4
[
-
]
6
.
3
[
0.4
0.2
0.0
-0.2
-0.4
Parsamian 21
B
C
A
-0.5
0.0
[5.8]-[8.0]
0.5
1.0
Figure 10. IRAC colour-colour diagram of 100 sources located
in the same area as in Fig. 9. The gray square in the upper
right corner marks the approximate domain of Class II sources
(Allen et al. 2004).
Figure 11. Sketch of the morphology of circumstellar material
around Parsamian 21, overlaid on the HST/WFPC2 image. The
central star is surrounded by an edge-on disc. Perpendicular to the
disc, the star drives a bipolar outflow that excavates an outflow
cavity in the dense circumstellar material. Light from the central
star illuminates the walls of the cavity.
Table 3. VLT/NACO photometry for the two closest stars. Un-
certainties are about 0.1 mag.
α2000
δ2000
H mag KS mag
L′ mag
19h 29m 0.s96
19h 29m 0.s70
9◦ 38′ 42.′′11
9◦ 38′ 44.′′78
20.2
20.6
19.1
19.5
>14.6
>14.4
and the only known young star associated with it is the
T Tauri star AS 353 (Dame & Thaddeus 1985).
In order to check whether FUors are usually associated
with star forming regions, we searched the literature and
found that most FUors are located in areas of active star
formation (e.g. Henning et al. 1998). To find out whether
there is star formation in the vicinity of Parsamian 21,
we searched for pre-main sequence stars. For this purpose
we constructed a 2MASS J−H vs. H−KS and an IRAC
[3.6]−[4.5] vs. [5.8]−[8.0] colour-colour diagram for sources
found in our 5.′3 × 5.′3 IRAC field of view (Fig. 9, 10). Our
selection criteria in the case of 2MASS was S/N > 10 and
uncertainties < 0.1 mag in all J, H and KS bands, while
in the case of IRAC S/N > 3 and detectability at all four
bands were required. The 2MASS diagram revealed that
most of the nearby objects are reddened main sequence or
giant stars. On the IRAC diagram, however, there are three
objects (apart from Parsamian 21 itself), which display in-
frared excess at 8 µm (marked by A, B and C in Fig. 10).
According to the classification of Allen et al. (2004), Class
II sources exhibit colours of [3.6]−[4.5] > 0.0 and [5.8]−[8.0]
> 0.4, thus one of these stars (A) might be a Class II source,
while B and C are more likely Class III/main sequence
sources. The nature of source A and its possible relation-
ship to Parsamian 21 is yet to be investigated. Nevertheless,
Parsamian 21 seems to be rather isolated compared to most
FUors and certainly not associated with any rich cluster of
young stellar objects.
We also searched for possible close companions of
Parsamian 21 in the WFPC2 and NACO direct images. In
order to establish a detection limit for source detection, we
measured the sky brightness on the NACO images (before
sky-subtraction), and estimated a limiting magnitude for
each filter. The resulting values are 22.8, 21.6, and 15.2 mag
in H, KS and L′, respectively. In case of the HST/WFPC2
image, the larger (80 × 80 arcsec) field of view made it pos-
sible to estimate a limiting magnitude using star counts;
the resulting value is 23.5 mag. Due to the bright reflec-
tion nebula, the detection limit is somewhat lower close to
the star. The two closest objects we found are the follow-
ing: one star to the southeast, at a distance of 1.4 arcsec
(560 AU at 400 pc), and another one to the northwest, at
a distance of 3.3 arcsec (1320 AU at 400 pc). These sources
are marked with arrows in Fig. 2. Neither of the stars are
visible at 3.8 or 0.814 µm, although we can give an upper
limit for their L′ brightness. Their positions and photom-
etry are given in Table 3. As these sources are very red,
they can equally be heavily reddened background stars, or
stars with infrared excess (indicating that they might be as-
sociated with Parsamian 21). Supposing that they are red-
dened main sequence stars, one can estimate an extinction
of AV ≈ 10 − 15 mag. Further multifilter observations may
help to clarify the nature of these objects and their possible
relationship to Parsamian 21.
4.3 The circumstellar environment of
Parsamian 21
The appearance and polarization properties of the nebula
around Parsamian 21 can be understood in the following
way: the star drives an approximately north-south oriented
bipolar outflow, which had excavated a conical cavity in
the dense circumstellar material (Fig. 11). The star illu-
minates this cavity and the light is scattered towards us
mainly from the walls of the cavity. The outflow direction
is perpendicular to an almost edge-on dense circumstellar
disc. This picture is supported by the following facts: (a) the
centrosymmetric polarization pattern is characteristic of re-
flection nebulae with single scattering; (b) the morphology
and limb brightening suggest a hollow cavity (as opposed
to an "outflow nebula", where the lobes are composed of
dense material ejected by the central source); and (c) the
low-polarization lane across the star strongly suggests the
presence of an edge-on circumstellar disc, where multiple
10
´A. K´osp´al et al.
0.5
0
-0.5
]
"
[
t
e
s
f
f
o
C
E
D
60
50
40
]
c
e
s
c
r
a
[
t
e
s
f
f
o
C
E
D
30
20
10
0
0
10%
1
0.5
0
-0.5
-1
FWHM
0.5
0
-0.5
20
40
60
RA offset [arcsec]
80
100
120
1
0.5
0
-0.5
-1
RA offset ["]
Figure 14. VLT/NACO polarization map of Parsamian 21 overlaid on H-band total intensity contours. The circle in the upper right
corner displays the FWHM of the polarization measurement. Polarization vectors are displayed at full resolution, only showing the central
low polarization band, where the polarization vectors are aligned. Such arrangement is expected when multiple scattering occurs in an
edge-on disc.
y
t
i
s
n
e
t
n
i
e
v
i
t
a
l
e
R
10000
1000
100
10
F814W
H
KS
1
1
2
3
4
Angular distance [arcsec]
]
s
t
n
u
o
C
[
s
s
e
n
t
h
g
i
r
B
1000
100
10
1
20
40
60
80
100
Position [pixel]
Figure 12. Brightness profiles of Parsamian 21 at 0.8 µm and in
the H and KS bands. Dotted lines mark Hubble's relation.
Figure 13. South-north cut at 0.6 arcsec east from the star. Solid
line: total intensity; dashed line: polarized intensity.
scattering occurs. In general the Parsamian 21 system shows
similarities to the NGC 2261 nebula associated with R Mon.
This object also consists of a northern cometary nebula and
a southern jetlike feature (Warren-Smith et al. 1987).
4.3.1 Envelope/Cavity
We characterised the opening of the upper lobe by mark-
ing the ridge along the northeastern and northwestern arcs
which we interpret as the walls of the cavity. As viewed from
the star northwards, the cavity starts as a cone with an open-
ing angle of ≈ 60◦, giving the nebula in Fig. 2 a characteris-
tic equilateral triangle-shape. Farther away from the star the
cavity deviates from the conical shape, becomes narrower.
The whole cavity occupies an area of 8 000 × 24 000 AU (at a
distance of 400 pc). The sharp outer boundary of the nebula
implies a significant density contrast between the cavity and
the surrounding envelope.
In Fig. 12 we plotted radial brightness profiles at 0.8 µm
and in the H and KS bands, starting from the star north-
wards. The slope of the intensity profiles in the inner ∼ 1.6′′
(640 AU) follows closely Hubble's relation (Hubble 1922),
i.e. the brightness is proportional to r−2. This trend can be
clearly followed above the noise out to about 5 arcsec in our
H and KS images, giving an estimate of 2000 AU for the
outer size of the envelope. The fact that the brightness pro-
files follow Hubble's relation implies that the nebula is pro-
duced by isotropic single scattering, in accordance with the
centrosymmetric polarization pattern and the high degree
of polarization. Around ∼ 1.6′′ the profiles becomes steeper,
probably due to decreased density in the inside of the cav-
ity. Similar steepening in the outer part of the nebula was
already mentioned by Li et al. (1994). The profiles are sim-
ilar at all observed wavelengths and do not show significant
dependence on the position angle.
4.3.2 Disc
As mentioned in Sect. 3.4, both the NACO and NICMOS
polarimetric images show a lane of low polarization oriented
nearly east-west across the star (Fig. 7). The drop in the
degree of polarization can also be clearly seen in Fig. 13,
where a north-south cut at 0.′′6 east to the star is plot-
ted. Following Bastien & M´enard (1990), we interpret the
low polarization by multiple scattering in an edge-on disc
(possible other explanations for the origin of the low po-
larization areas are discussed in e.g. Lucas & Roche 1998).
According to models of such circumstellar structures (e.g.
Whitney & Hartmann 1993; Fischer et al. 1996) the polar-
ization vectors are oriented parallel to the disc plane. As can
be seen in Fig. 14, despite the low degree of polarization, the
predicted alignment of the vectors can be clearly seen in the
case of Parsamian 21 too. The NICMOS polarization map
shows a similar effect. The dark lane in Fig. 7 can be fol-
lowed inwards to as close as 48 AU from the star. This is an
upper limit for the inner radius of the circumstellar disc.
An interesting feature of the mentioned models is two
depolarized areas on either side of the central star, which
mark the outer end points of an edge-on disc. The depolar-
ization is due to a transition from the linearly aligned to the
centro-symmetic polarization pattern. These depolarized ar-
eas can be seen in Fig. 8 bottom, on both sides at about 0.9′′
from the star. At the distance of Parsamian 21 this corre-
sponds to 360 AU, and can be adopted as the outer radius
of the dense part of the disc, where multiple scattering at
near-infrared wavelengths is dominant. It is interesting that
the NACO map does not show clear depolarized areas at the
same position, but seems to resolve the transition in vector
orientation, while the larger beam of NICMOS averaged the
differently oriented vectors, resulting in depolarized spots.
The polarized intensity and degree of polarization im-
ages in Fig. 7 suggest a slight asymmetry in the dense disc:
the eastern (left) side is straight, while the western (right)
side shows a kink and also has a different position angle than
that on the other side. The thickness of the disc can be mea-
sured on the area where the polarization vectors are aligned
(Fig. 14). This approximately corresponds to the area where
the degree of polarization is below ≈ 10%. The resulting
thickness is approximately 0.1′′ (40 AU) to the east and is
somewhat larger, 0.2′′ (80 AU), to the west. One should note
that, since these values are close to the spatial resolution of
the polarimetric images, these numbers should be consid-
ered as upper limits for the thickness of the circumstellar
disc around Parsamian 21. They are upper limits also be-
cause if the inclination is not exactly 90◦, the thickness of
the disc can be even less. The thickness does not show signifi-
cant increase with radial distance, suggesting the picture of a
flat, rather than a flared disc, at least considering the dense,
multiple-scattering part. The smooth brightness and polar-
High-resolution polarimetry of Parsamian 21
11
Table 4. Model parameters. Parameters in italics are fixed, while
the others were fitted.
Parameter
Variable
Value
Inner disc radius
Outer disc radius
Temperature at 1 AU
Power-law index for temperature
Power-law index for surface density
Disc mass
Inclination
Inner envelope radius
Outer envelope radius
Temperature at 5 AU
Power-law index for temperature
Power-law index for surface density
Envelope mass
Interstellar Extinction
R1
R2
Td,0
qd
pd
Md
i
R3
R4
Te,0
qe
pe
Me
AV
3.5 R⊙
360 AU
285 K
0.75
1.6
0.02 M⊙
86◦
5.4 AU
2000 AU
368 K
0.4
0.4
0.02 M⊙
2 mag
ization distribution between the disc and the surrounding
envelope (Fig. 13), however, implies that there is a contin-
uous density transition between the two components. From
the ratio of the horizontal to vertical sizes of the disc a lower
limit of 84◦ for the inclination of the system (the angle be-
tween the normal of the disc and the line of sight) can be
derived.
Fischer et al. (1996) computed a grid of polarization
maps of young stellar objects with the aim of help-
ing the interpretation of polarimetric imaging observa-
tions. They consider five different models, four with mas-
sive, self-gravitating discs and one with a massless Ke-
plerian disc. Since the mass of circumstellar material of
Parsamian 21 derived from submillimetre observations is rel-
atively low (< 0.3 M⊙, Henning et al. 1998; Polomski et al.
2005; Sandell & Weintraub 2001; Hillenbrand et al. 1992),
the most appropriate model
for our case is a Keple-
rian disc. Indeed the polarization pattern as computed by
Fischer et al. (1996) for an inclination of 87◦ (their Fig. 1)
looks remarkably similar to our Fig. 8 (the differences might
be explained by the narrower cavity and flatter disc of
Parsamian 21). Thus, the geometry and structure assumed
by Fischer et al. (1996) in their Keplerian model could be a
good starting point for further radiative transfer modelling
of Parsamian 21.
4.4 Modelling the circumstellar environment
Our observations provide some direct measurements of the
geometry of the circumstellar structure (disc size and thick-
ness, inclination, envelope size). In the following we discuss
the consistency of this picture with the observed SED. Our
approach is to construct a simple disc+envelope model, in
which we fix those parameters whose values are known from
our NACO observations or from other sources (outer disc
radius from this work; power-law index for disc temperature
from Shakura & Sunyaev 1973; disc mass was set in order to
ensure that the whole disc is optically thick; inclination from
this work; outer envelope radius from this work, the power-
law index for envelope temperature is a typical value for
optically thin envelopes containing larger than interstellar
grains, e.g. Hartmann 2000, Eqn. 4.13; interstellar extinction
12
´A. K´osp´al et al.
from Hillenbrand et al. 1992). Then we check whether the
SED can be fitted by tuning the remaining parameters. We
adopted an analytical disk model (Adams et al. 1987), which
has been successfully used to model FUors (Quanz et al.
2006; ´Abrah´am et al. 2006). Our model consists of two com-
ponents, an optically thick and geometrically thin accretion
disc (Shakura & Sunyaev 1973) and an optically thin en-
velope (no cavity is assumed). No central star is included
in the simulation, partly because in outbursting FUors the
star's contribution is negligible compared to that of the in-
ner disc (Hartmann & Kenyon 1996), and partly because of
the edge-on geometry where the star is obscured by the disc.
This assumption is supported by the fact that the shape of
the SED at optical wavelengths is broader than a stellar pho-
tosphere. The model also does not take into account internal
extinction and light scattering, thus it cannot reproduce any
of the near-IR imaging and polarimetric observations.
The temperature and surface density distribution in the
disc are described by power-laws:
T (r) = Td,0(cid:16) r
Σ(r) = Σd,0(cid:16) r
1 AU(cid:17)−qd
1 AU(cid:17)−pd
,
.
(1)
(2)
Similar power-laws were assumed for the envelope. The ob-
served flux at a specific frequency is given by
Fν =
2πr(1 − e
−Σ
d κν
cos i )Bν (Td)dr +
(3)
(4)
cos i
D2 Z R2
D2 Z R4
1
R1
R3
2πr(1 − e−Σe κν )Bν(Te)dr.
The first term describes the emission of the accretion disc,
the second term describes the radiation of the optically thin
envelope. For the dust opacity we used a constant value
of κν = 1 cm2g−1 at λ > 1300 µm, κν = κ1300µm (cid:0)
between 1300 and 100 µm and again a constant value of
κν = κ(100 µm) at λ < 100 µm.
1300µm(cid:1)−1
λ
We fitted the SED via χ2 minimisation using a genetic
optimization algorithm PIKAIA (Charbonneau 1995). This
algorithm performs the maximization of a user defined func-
tion, for which purpose we used the inverse χ2. Since there
are many photometric measurements in the mid-infrared do-
main, but just a few in the far-infrared, the mid-infrared
region has a higher weight during the fit, compared to the
far-infrared domain. Therefore, in order to ensure an equally
good fit at all wavelengths, we divided the SED into four re-
gions and weighted the χ2 of each domain with the inverse
of number of photmetric points the region contained. Then
the final χ2 was the sum of the χ2 of all regions. The regions
we used were: 0.3 − 3, 3 − 30, 30 − 300 and 300 − 3000 µm.
The parameters of the best-fit model, which gives a weighted
χ2 of 0.67, are listed in Table 4. The fitted model SED as
well as the disc and envelope components are overplotted in
Fig. 5.
The model SED is consistent with the observed fluxes.
This shows that the picture of a thin accretion disc and
an envelope is consistent with both the measured SED and
the geometry and disc/envelope parameters inferred from
our polarimetric observations. Detailed modelling of the sil-
icate, PAH, and ice spectral features, as well as the correct
treatment of internal extinction and scattering would re-
quire radiative transfer modelling, which will be the topic of
a subsequent paper.
4.5 The evolutionary status of Parsamian 21
The geometry of FUor models discussed in the literature
(e.g. Hartmann & Kenyon 1996; Turner et al. 1997) usually
consist of a central star surrounded by an accretion disc
and an infalling envelope with a wind-driven polar hole.
These assumptions are supported by the fact that they
fit well the SED (Green et al. 2006; Quanz et al. 2007),
the interferometric visibilities (Millan-Gabet et al. 2006b;
´Abrah´am et al. 2006) and the temporal evolution of the SED
( ´Abrah´am et al. 2004). In this paper we present the first di-
rect imaging of these circumstellar structures in a FUor. Our
polarimetric measurements of Parsamian 21 show the exis-
tence of a circumstellar disc which extends from at least 48
to 360 AU. The most striking feature of the disc is its flat-
ness over the whole observed range. The short-wavelength
part of the SED could be well reproduced using a radial
temperature profile of r−0.75 (Sect. 4.4). This profile is ex-
pected from both a geometrically thin accretion disc and a
flat reprocessing disc. An envelope was also seen in the po-
larization maps of Parsamian 21 and it was also a necessary
component for the SED modelling. Envelopes are involved
in many FUor models and in this paper we present a direct
detection of this model component. Our images reveal that
the envelope can be followed inwards as close to the star as
the disc. FUor models often assume a polar cavity in the
envelope, created by a strong outflow or disc wind. The di-
rect images of Parsamian 21 clearly show the presence of
such a cavity and we also detected a bipolar outflow in the
Parsamian 21 system.
In the recent years, as new interferometric and infrared
spectroscopic observations were published for FUors, the
group turned out to be more inhomogeneous in physical
properties than earlier assumed, when mainly optical pho-
tometry and spectroscopy had been available. Quanz et al.
(2007) proposed that some differences might be understood
as an evolutionary sequence. They suggest that FUors con-
stitute the link between embedded Class I objects and the
more evolved Class II objects. Members of the group exhibit-
ing silicate absorption at 10 µm are younger and more em-
bedded (Category 1, e.g. V346 Nor); while objects with pure
silicate emission are more evolved (Category 2, e.g. FU Ori
and Bran 76). There are objects showing a superposition of
silicate absorption and emission, which are probably in an
intermediary evolutionary stage (e.g. RNO 1B). Green et al.
(2006) also sorted FUors, based on the ratio of the far-
infrared excess and the luminosity of the central accre-
tion disc, fd (Equ. 7 in their paper). A large relative ex-
cess (fd > 5%) indicates an envelope of large covering frac-
tion (V1057 Cyg and V1515 Cyg), while low relative excess
means a tenuous or completely missing envelope (Bran 76
and FU Ori). This is also an evolutionary sequence, as young,
more embedded objects have large envelopes, while around
more evolved stars, the envelope has already dispersed. fd
can also be used to calculate the opening angle of the en-
velope, thus a prediction of this scheme is that the opening
angle is becoming wider during the evolution, probably due
to strong outflows during the repeated FUor outbursts.
The two classification schemes are not inconsistent and
one can merge them into the following evolutionary se-
quence: (1) the youngest objects exhibit silicate absorp-
tion and large far-infrared excess (V346 Nor, probably also
OO Ser and L1551 IRS 5 belong here); (2) intermediate-aged
objects, where the silicate feature is already in emission but
there is still a significant far-infrared excess (V1057 Cyg,
V1515 Cyg, probably also RNO 1B and V1647 Ori); (3) the
most evolved objects show pure silicate emission and low far-
infrared excess (FU Ori, Bran 76). We note, however, that
this classification has some weak points. As Quanz et al.
(2007) already mentioned, an edge-on geometry in a more
evolved system may appear as a younger one. Moreover,
during an outburst and the subsequent fading phase, cer-
tain spectral features as well as the global shape of the SED
may change.
Parsamian 21 can be placed in this evolutionary scheme,
though one should keep in mind that because of the nearly
edge-on geometry, the classification of this object is some-
what uncertain. Parsamian 21 displays silicate emission
(Fig. 6). Integrating the flux of the two components in our
simple model (Sect. 4.4), and correcting the apparent disc
luminosity for inclination effect (i = 86◦, Table 4) using
Equ. 6 of Green et al. (2006), we obtained a large relative
far infrared excess of fd = 75%. These two properties place
Parsamian 21 into the intermediate-aged category (though
because of its inclination, it may actually seem younger than
it is). Following Equ. 7 of Green et al. (2006), from the fd
value, we also computed the opening angle of the envelope.
The resulting opening angle of 60◦ agrees well with the angle
measured in the direct NACO images (Sect. 4.3.1).
Parsamian 21 was placed into the evolutionary scheme
using two parameters: the silicate feature and the relative far
infrared excess. In the following we discuss whether its other
physical characteristics match with those of other FUors.
(i) Our observations revealed that the circumstellar disc
of Parsamian 21 is very flat. Due to lack of similar di-
rect measurements for other FUors, we can only specu-
late that perhaps all FUors with envelopes have such flat
discs. On the other hand, the most evolved FUor, FU Ori,
seems to have no envelope but its disc is probably flared
(Kenyon & Hartmann 1991; Green et al. 2006; Quanz et al.
2006). This might suggest that disc flaring develops at later
stages, when illumination from the central source may heat
the disc surface more directly.
(ii) In a nearly edge-on system like Parsamian 21, one
expects to see the 10 µm silicate feature in absorption.
The fact that Parsamian 21 has silicate emission indi-
cates that the line of sight towards the central region is
not completely obscured. Using the optical depth of the
15.2 µm CO2 ice feature, we calculated an AV = 8 mag
(AV = 38.7A15.2µm, Savage & Mathis 1979). This value
is surprisingly low compared to V1057 Cyg (AV ∼ 50 −
100 mag, Kenyon & Hartmann 1991). This indicates a much
more tenuous envelope, which is also supported by the low
envelope mass of 0.02 M⊙ in our modelling.
(iii) Following Quanz et al. (2007), we analysed the profile
of the 15.2 µm CO2 ice feature of Parsamian 21. The inset
in Fig. 6 shows that the feature has a characteristic double-
peaked sub-structure, very similar to HH 46 IRS, an embed-
ded young source (Boogert et al. 2004). HH 46 IRS is a refer-
ence case for processed ice. The presence of processed ice in
High-resolution polarimetry of Parsamian 21
13
Parsamian 21 indicates heating processes and the segrega-
tion of CO2 and H2O ice, already at this evolutionary stage.
Other FUors exhibiting this kind of profile are L1551 IRS 5,
RNO 1B and RNO 1C (Quanz et al. 2007).
The evolutionary state of a young stellar object can also
be estimated following the method proposed by Chen et al.
(1995). According to their Equ. (1) we calculated a bolo-
metric temperature of Tbol = 410 K for the measured SED.
We compared this value with the distribution of correspond-
ing values among young stellar objects in the Taurus and
ρ Ophiuchus star forming regions (Chen et al. 1995). From
this check we can conclude that Parsamian 21 seems to be a
class I object, and its age is ∼ 105 yr. However, Green et al.
(2006) argued that the apparent SED of the disc component
depends on the inclination. Thus we computed Tbol also for
a face-on disc configuration and obtained Tbol = 1160 K,
corresponding to a Class II object. In fact, Parsamian 21 is
probably close to the Class I / Class II border, in accordance
with the proposal of Quanz et al. (2007).
5 SUMMARY
We present the first high spatial resolution near-infrared di-
rect and polarimetric observations of Parsamian 21, with
the VLT/NACO instrument. We complemented these mea-
surements with archival
infrared observations, such as
HST/WFPC2 imaging, HST/NICMOS polarimetry, Spitzer
IRAC and MIPS photometry, Spitzer IRS spectroscopy as
well as ISO photometry. Our main conclusions are the fol-
lowing:
(1) We argue that Parsamian 21 is probably an FU Orionis-
type object;
(2) Parsamian 21 is not associated with any known rich
cluster of young stars;
(3) our measurements reveal a circumstellar envelope, a po-
lar cavity and an edge-on disc; the disc seems to be geomet-
rically flat and extends from at least 48 to 360 AU from the
star;
(4) the SED is consistent with a simple circumstellar disc
+ envelope model;
(5) within the framework of an evolutionary sequence of
FUors proposed by Quanz et al. (2007) and Green et al.
(2006), Parsamian 21 can be classified as an intermediate-
aged object.
ACKNOWLEDGMENTS
We are grateful for the Paranal staff for their support dur-
ing the observing run and thank our support astronomers
O. Marco and N. Ageorges. For the GFP observations and
data reduction we thank G. M. Williger, G. Hilton and B.
Woodgate. Observing time at the Apache Point Observa-
tory 3.5m telescope was provided by a grant of Director's
Discretionary Time. Apache Point is operated by the Astro-
physical Research Consortium. The Goddard Fabry-Perot
is supported under NASA RTOP 51-188-01-22 to GSFC.
CAG is also supported as part of the Astrophysics Data
Program under NASA Contract NNH06CC28C to Eureka
14
´A. K´osp´al et al.
Scientific. This research made use of the SIMBAD astro-
nomical database. This material is partly based upon work
supported by the National Aeronautics and Space Admin-
istration through the NASA Astrobiology Institute under
Cooperative Agreement No. CAN-02-OSS-02 issued through
the Office of Space Science. The work was partly supported
by the grant OTKA K 62304 of the Hungarian Scientific Re-
search Fund.
REFERENCES
´Abrah´am P., K´osp´al ´A., Csizmadia S., et al., 2004, A&A,
428, 89
´Abrah´am P., Mosoni L., Henning T., et al., 2006, A&A,
449, L13
Adams F. C., Lada C. J., Shu F. H., 1987, ApJ, 312, 788
Allen L. E., Calvet N., D'Alessio P., et al., 2004, ApJS,
154, 363
Hines D. C., Schmidt G. D., Schneider G., 2000, PASP,
112, 983
Hines D. C., Schneider G., 2006, in Proceedings of the
2005 HST calibration workshop Polarimetry with nicmos.
p. 153
Hubble E. P., 1922, ApJ, 56, 400
Kenyon S. J., Hartmann L. W., 1991, ApJ, 383, 664
Koornneef J., 1983, A&A, 128, 84
Kuhn J. R., Potter D., Parise B., 2001, ApJ, 553, L189
Kwok S., 1993, ARA&A, 31, 63
Lenzen R., Hofmann R., Bizenberger P., et al., 1998, in
Proc. SPIE Vol. 3354, p. 606-614, Infrared Astronomi-
cal Instrumentation, Albert M. Fowler; Ed. CONICA: the
high-resolution near-infrared camera for the ESO VLT. pp
606 -- 614
Li W., Evans N., Harvey P., et al., 1994, ApJ, 433, 199
Lucas P. W., Roche P. F., 1998, MNRAS, 299, 69
Malbet F., Berger J.-P., Colavita M. M., et al., 1998, ApJ,
507, L149
Apai D., Pascucci I., Brandner W., et al., 2004, A&A, 415,
Malbet F., Lachaume R., Berger J.-P., et al., 2005, A&A,
671
437, 627
Bastien P., M´enard F., 1990, ApJ, 364, 232
Boley A. C., Mej´ıa A. C., Durisen R. H., et al., 2006, ApJ,
651, 517
Mazzuca L., Hines D., 1999, Instruemnt Science Report
NICMOS ISR-99-004
Meakin C. A., Hines D. C., Thompson R. I., 2005, ApJ,
Boogert A. C. A., Pontoppidan K. M., Lahuis F., et al.,
634, 1146
2004, ApJSS, 154, 359
Meyer M. R., Calvet N., Hillenbrand L. A., 1997, AJ, 114,
Bouchet P., Schmider F. X., Manfroid J., 1991, Ap&SS, 91,
288
409
Millan-Gabet R., Monnier J. D., Akeson R. L., et al., 2006a,
Cardelli J. A., Clayton G. C., Mathis J. S., 1989, ApJ, 345,
ApJ, 641, 547
245
Charbonneau 1995, ApJS, 101, 309
Chen H., Myers P. C., Ladd E. F., et al., 1995, ApJ, 445,
377
Dame T., Thaddeus P., 1985, ApJ, 297, 751
Decin L., Morris P., Appleton P. N., Charmandaris V., Ar-
Millan-Gabet R., Monnier J. D., Akeson R. L., et al., 2006b,
ApJ, 641, 547
Mundt R., Brugel E. W., Buehrke T., 1987, ApJ, 319, 275
Neckel T., Staude H. J., 1984, A&A, 131, 200
Parsamian E., 1965, Izv. Akad. Nauk Armyan. SSR., Ser.,
18, 146
mus L., Houck J. R., 2004, ApJS, 154, 408
Parsamian E. S., Gasparian K. G., Ohanian G. B., 1996,
Diolaiti E., Bendinelli O., Bonaccini D., et al., 2000,
Astrophysics, 39, 121
Proc. SPIE, 4007, 879
Persson S. E., Murphy D. C., Krzeminski W., et al., 1998,
Draper P. W., Warren-Smith R. F., Scarrott S. M., 1985,
AJ, 116, 2475
MNRAS, 212, 1P
Fischer O., Henning T., Yorke H. W., 1996, A&A, 308, 863
Green J. D., Hartmann L., Calvet N., et al., 2006, ApJ,
648, 1099
Polomski E. F., Woodward C. E., Holmes E. K., et al.,
2005, AJ, 129, 1035
Quanz S., Henning T., Bouwman J., et al., 2007, ApJ, in
press, ApJ preprint doi: 10.1086/'521219'
Hajjar R., Bastien P., Nadeau D., 1997, Canada-France-
Quanz S. P., Henning T., Bouwman J., et al., 2006, ApJ,
Hawaii Telescope Information Bulletin, 37, 3
648, 472
Hartmann L., 2000, Accretion Processes in Star Formation.
Cambridge University Press
Hartmann L., Kenyon S. J., 1996, ARA&A, 34, 207
Hartung M., Bizenberger P., Boehm A., et al., 2000, in
Proc. SPIE Vol. 4008, p. 830-841, Optical and IR Tele-
scope Instrumentation and Detectors, Masanori Iye; Alan
F. Moorwood; Eds. First test results and calibration meth-
ods of CONICA as a stand-alone device. pp 830 -- 841
Henning T., Burkert A., Launhardt R., et al., 1998, A&A,
336, 565
Herbig G. H., 1977, ApJ, 217, 693
Herbig G. H., Petrov P. P., Duemmler R., 2003, ApJ, 595,
384
Hessman F. V., Eisloeffel J., Mundt R., et al., 1991, ApJ,
Reipurth B., Bally J., 2001, ARA&A, 39, 403
Rousset G., Lacombe F., Puget P., et al., 2003, in Adaptive
Optical System Technologies II. Edited by Wizinowich,
Peter L.; Bonaccini, Domenico. Proceedings of the SPIE,
Volume 4839, pp. 140-149 (2003). NAOS, the first AO
system of the VLT: on-sky performance. pp 140 -- 149
Sandell G., Weintraub D. A., 2001, ApJ, 134, 115
Savage Mathis 1979, ARA&A, 17, 73
Scarrott R. M. J., Scarrott S. M., Wolstencroft R. D., 1993,
MNRAS, 264, 740
Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337
Staude H. J., Neckel T., 1992, ApJ, 400, 556
Tuffs R. J., Gabriel C., 2003, A&A, 410, 1075
Turner N. J. J., Bodenheimer P., Bell K. R., 1997, ApJ,
370, 384
480, 754
Hillenbrand L. A., Strom S. E., Vrba F. J., et al., 1992,
ApJ, 397, 613
Vorobyov E. I., Basu S., 2006, ApJ, 650, 956
Warren-Smith R. F., Draper P. W., Scarrott S. M., 1987,
Table A1. Log of IRS calibration measurements.
Channel
AOR
Date
Target
Short Low
Short High
Long Low
Long High
16295168
19324160
16294912
16463104
16101888
2005-Nov-21 HR 7341
2006-Jul-5
HR 7341
2005-Nov-21 HR 6688
2005-Dec-19 HR 6606
HR 2491
2005-Oct-18
High-resolution polarimetry of Parsamian 21
15
SL1 @ 12 µm
LL1 @ 27 µm
1.0
0.8
0.6
x
u
l
f
e
v
i
t
a
l
e
R
0.4
0.2
0.0
-5
0
5
-20
-10
0
10
20
Distance from the center of the slit in perpendicular direction ["]
SH @ 15 µm
LH @ 24.4 µm
1.0
0.8
0.6
x
u
l
f
e
v
i
t
a
l
e
R
0.4
0.2
0.0
x
u
l
f
e
v
i
t
a
l
e
R
x
u
l
f
e
v
i
t
a
l
e
R
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
-5
0
5
10
Distance from the center of the slit in perpendicular direction ["]
-15
0
-10
-5
5
15
Figure A2. Beam profiles of the four Spitzer/IRS channels, per-
pendicular to the slit, at selected wavelengths. Filled dots show
the measured profile, while open dots indicate the pixelized model
PSF.
the measured star (Decin et al. 2004). The result is a data
cube with two spatial dimensions and one wavelength di-
mension. We resampled these data cubes to finer spatial
grid and smoothed them in wavelength.
We checked whether the measured beam profiles are
consistent with the PSFs provided by Spitzer's Tiny Tim
(J. Krist). In Fig. A2 we plotted the ratio between the ob-
served and the model spectrum at different distances from
the slit centre at a certain wavelength. We found slight dif-
ferences: the measured profiles were in general narrower than
the model PSFs, and in some cases they were not centred at
zero. We examined several effects that can cause a difference
between the model and the measured profiles and we found
that the differences can be attributed to pixelization effects
(pixel sizes are 1.8′′ for SL, 2.3′′ for SH, 5.1′′ for LL and 4.5′′
for LH) and to small (few tenths of arcsecond) uncertainties
in the exact position of the calibration star with respect to
the slit. Based on our analysis, we decided to use the mea-
sured profiles to correct our science measurements, and to
use the TinyTim model profiles to estimate the uncertainty
of the correction. Using this correction, the absolute flux
level of IRS spectra has an uncertainty of 10%, but individ-
ual measurements can be much more precise if the source is
well-centred in the slit and the necessary correction is small.
More details on IRS beam profiles will be given in a future
paper.
Figure A3 shows the position of the IRS slits overplot-
ted on the HST/WFPC2 image of Parsamian 21. While the
target is well-centred in the Short Low slit, it is slightly
off-centred in the Long High and it is practically off-slit in
the Short High. Considering our measured beam profiles, we
found that no correction is necessary in the Short Low chan-
nel, a small correction (5 − 20%, depending on wavelength)
is necessary in the Long High channel, and a significant cor-
rection must be made in the Short High channel, where the
corrected flux is approximately 3-4 times the original one
(see Fig. A4). After the correction, the different IRS chan-
nels match fairly well in the overlapping wavelength regimes.
Figure A1. Top: Dots show the measured spectrum of HR 6688,
obtained with Spitzer/IRS through the Short High slit; the con-
tinuous line show a model spectrum of the same star (from
Decin et al. 2004). Bottom: Scheme of the data grid; the + sign
marks the slit centre; dots show the position of the star relative
to the slit centre; the rectangle indicates the slit itself; the × sign
marks the position where the spectrum plotted in the top panel
was taken.
ApJ, 315, 500
Wassell E., Grady C.A.and Woodgate B., et al., 2006, ApJ,
650, 985
Whitney B. A., Hartmann L., 1993, ApJ, 402, 605
Whittet D. C. B., Martin P. G., Hough J. H., et al., 1992,
ApJ, 386, 562
APPENDIX A: REFINED CALIBRATION OF
THE SPITZER/IRS BEAM PROFILES
In those cases when the source is not well centred in the
slit, part of the stellar PSF falls outside of the slit, lead-
ing to flux loss. When the precise absolute flux level of the
spectrum is important, one should correct for this flux loss.
Correction for this kind of flux loss is not implemented in
the Spitzer data reduction pipeline. For point sources flux
loss in off-centred sources can be efficiently corrected if we
know the beam profiles of the different IRS slits. In order
to construct the beam profiles for all four IRS slits, we re-
duced and analysed dedicated IRS calibration measurements
taken in spectral mapping mode in a regular grid. Table A1
shows a log of these calibration measurements. We used the
pipeline processed post-BCD files (pipeline version S14.0.0).
At each spatial position, we divided the measured spectrum
with a synthetic MARCS model spectrum appropriate for
16
´A. K´osp´al et al.
Figure A3. Positions of the Spitzer/IRS slits with respect to
Parsamian 21 (marked with an asterisk). Black rectangle: Short
Low, small grey rectangle: Short High, big grey rectangle: Long
High.
Figure A4. Short High channel spectrum of Parsamian 21.
|
astro-ph/0102498 | 1 | 0102 | 2001-02-28T19:03:49 | An Extremely Slow Nova? | [
"astro-ph"
] | The binary companion to the peculiar F supergiant HD 172481 is shown to be a Mira variable with a pulsation period of 312 days. Its characteristics are within the normal range found for solitary Miras of that period, although its pulsation amplitude and mass-loss rate (M(dot) ~ 3 x 10^{-6} solar masses per year) are higher than average. Reasons are given for suspecting that the F supergiant, which has L ~ 10^4 solar luminosities, is a white dwarf burning hydrogen accreted from its companion. | astro-ph | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- ?? (2001)
Printed 24 December 2018
(MN LATEX style file v1.4)
An Extremely Slow Nova?
Patricia Whitelock⋆ & Freddy Marang
SAAO, PO Box 9, Observatory, 7935, South Africa (email: [email protected]).
24 December 2018
1
0
0
2
b
e
F
8
2
1
v
8
9
4
2
0
1
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
ABSTRACT
The binary companion to the peculiar F supergiant HD 172481 is shown to be a Mira
variable with a pulsation period of 312 days. Its characteristics are within the normal
range found for solitary Miras of that period, although its pulsation amplitude and
mass-loss rate M ∼ 3 × 10−6 M⊙yr−1 are higher than average. Reasons are given for
suspecting that the F supergiant, which has L ∼ 104 L⊙, is a white dwarf burning
hydrogen accreted from its companion.
Key words: binaries, novae, infrared: stars.
1
INTRODUCTION
3 NATURE OF THE COOL STAR
HD 172481 is a high galactic latitude F-type supergiant
which has drawn some attention since the IRAS survey
showed it to have an infrared dust excess (Odenwald 1986).
Reyniers & Van Winckel (2000, henceforth RVW) have
since shown that the star is a spectroscopic binary with a
cool M-giant companion. Their detailed spectroscopic anal-
ysis of the F supergiant indicates a moderate metal defi-
ciency ([Fe/H]= -- 0.55), no CNO-enhancement, a slight en-
hancement of s-process elements and high lithium content
(log ǫ(Li) = 3.6). Essentially nothing was known about the
M star prior to the present paper.
Since the IRAS measurements most authors have as-
sumed that the F supergiant is a post-AGB star (RVW and
references therein). This paper first discussed the nature of
the M giant (section 3) and then, more speculatively, pro-
vides a possible explanation for the F supergiant (section 4).
2 PHOTOMETRY
The published Stromgren and Geneva photometry (Houk
& Mermilliod 1998) indicate V between 8.9 and 9.1 while
Geneva photometry from RVW shows that the visual mag-
nitude is variable with a peak to peak amplitude of 0.22
mag. RVW show that the M star contributes about 10 per-
cent of the flux at V when the Mira is at maximum light
(see below). The value, V = 9.09 mag, given in the simbad
database is used in the following analysis as representative
of the mean light of the F star.
Forty-five JHKL measurements were made between
1988 and 1999 using the MkII photometer on the 0.75-m
telescope at SAAO Sutherland. These are on the SAAO
system (Carter 1990) and are listed in Table 1. They are
accurate to about ±0.03 mag at JHK and ±0.05 mag at L.
c(cid:13) 2001 RAS
A Fourier analysis of the data in Table 1 reveals a clear
period of 312 days and light curves folded on this period
are shown in Fig. 1. The peak to peak amplitudes of the
Fourier fits to these variations are: ∆J = 0.53, ∆H = 0.68,
∆K = 0.62 and ∆L = 0.58 mag. These of course are lower
limits to the amplitudes of the M giant as the F supergiant
will contribute to the observed magnitude, particularly at
the shorter wavelengths. Corrected amplitudes are discussed
in section 3.1 and listed in Table 2.
These large amplitude variations clearly demonstrate
that the M giant is a Mira variable and hence at the very
tip of the Asymptotic Giant Branch (Feast & Whitelock
1987).
3.1 Infrared Colours
The reddening from HD 172481 has been estimated as E(B−
V ) = 0.44 (RVW) and 0.27 (Bersier 1996). In the following
we calculate the magnitudes and colours using both of these
values, generally giving the value for E(B − V ) = 0.44 first
and following it with the value associated with E(B − V ) =
0.27, in brackets.
The first two lines of Table 2 list the magnitudes for
HD 172481 at the Fourier maximum and minimum (note
that, as seen in Fig. 1, the phase shifts between the JHKL
light-curves are negligible). The next three lines have been
corrected assuming a reddening of E(B − V ) = 0.44 and
give, respectively, the reddened corrected magnitudes of the
F2Ia star (from Johnson 1966) on the assumption that this
is the only significant contributor at V , the reddening cor-
rected mean magnitude of the Mira after the contribution
from the F star has been removed, and the peak to peak
amplitude of the Mira variations, also after removing the
contribution of the F star. These three lines are then re-
2
Patricia Whitelock & Freddy Marang
Table 1. Observed Magnitudes
JD
−2440000
7252.63
7270.65
7328.54
7334.49
7356.41
7362.43
7372.43
7383.42
7393.35
7422.34
7461.30
7617.64
7685.52
7698.48
7709.45
7719.46
7737.37
7742.45
7760.33
7767.31
7816.24
7823.28
7836.26
8054.57
8066.56
8093.53
8115.43
8166.35
8172.35
8451.48
8453.48
8853.36
9146.65
9223.42
9583.44
9615.39
9641.32
10238.57
10589.61
10716.35
10755.25
11100.27
11299.61
11388.40
11675.55
11711.57
J
H
K
L
(mag)
6.693
6.671
6.740
6.742
6.874
6.934
7.005
7.059
7.187
7.196
7.097
6.687
7.069
7.106
7.141
7.167
7.102
7.128
7.111
7.095
7.030
6.971
6.904
7.211
7.204
7.139
7.149
6.856
6.788
7.017
7.006
6.617
6.745
6.876
7.195
7.229
7.219
7.251
7.160
6.607
6.723
6.965
6.754
6.788
6.758
6.904
5.957
5.918
5.967
5.997
6.156
6.224
6.317
6.384
6.551
6.544
6.436
5.878
6.388
6.437
6.494
6.518
6.473
6.477
6.455
6.437
6.386
6.313
6.192
6.580
6.554
6.479
6.495
6.130
6.057
6.341
6.326
5.820
5.948
6.174
6.438
6.589
6.584
6.615
6.477
5.795
5.924
6.239
5.986
6.068
5.983
6.178
5.644
5.606
5.611
5.639
5.775
5.821
5.898
5.965
6.111
6.094
5.996
5.499
5.940
6.002
6.050
6.092
6.042
6.050
6.027
6.020
5.977
5.926
5.807
6.133
6.111
6.050
6.073
5.751
5.703
5.937
5.919
5.461
5.595
5.793
6.033
6.167
6.135
6.201
6.068
5.429
5.553
5.808
5.645
5.684
5.584
5.759
5.197
5.138
5.166
5.175
5.303
5.422
5.444
5.510
5.606
5.553
4.922
5.376
5.357
5.507
5.557
5.507
5.541
5.481
5.459
5.327
5.291
5.170
5.559
5.614
5.437
5.488
5.177
5.165
5.336
5.372
4.953
5.134
5.276
5.709
5.568
5.650
5.691
5.579
4.915
5.071
5.398
5.173
5.202
5.163
5.222
Table 2. Mean Magnitudes
component
max
min
V
-
(9.09)
−E(B − V ) = 0.44
7.75
F2Ia
Mira (mean)
∆
−E(B − V ) = 0.27
8.27
F2Ia
Mira (mean)
∆
-
-
-
-
J
6.65
7.18
7.21
7.58
1.3
7.73
7.29
0.9
H
(mag)
5.85
6.53
7.04
6.58
1.1
7.56
6.41
0.9
K
L
5.49
6.11
7.00
6.08
0.9
7.52
5.98
0.8
4.99
5.57
6.95
5.47
0.7
7.47
5.39
0.7
Figure 1. The infrared observations for HD 172481 phased on a
period of 312 days, with an arbitrary zero-point of JD 2440000;
each point is plotted twice.
peated using the alternative, E(B − V ) = 0.27, reddening
correction.
It is instructive to compare the colours deduced for
this Mira with those of normal isolated Miras. Whitelock,
Marang & Feast (2000, henceforth WMF) provide JHKL
observations of Miras with thin dust shells, while Whitelock
et al. (1994) do the same for IRAS selected Miras with rel-
atively thick shells. The mean colours of the Mira are listed
in Table 3 together with those of a typical 312 day Mira,
calculated from equations 1 to 4 of WMF.
The deduced colours are obviously somewhat uncertain,
particularly at J, because of the uncertain correction for the
F star. Nevertheless, they fall within the range shown by nor-
mal unreddened Miras. In particular, K −L, although larger
than the mean calculated from WMF equation 3, is within
the range shown by normal Miras. WMF note a strong cor-
relation between the pulsation amplitude and K − L, and
indeed the pulsation amplitude of this Mira is high.
This Mira is of particular interest in view of the fact
that we can infer its metallicity; only a very small number
of Miras, those associated with globular clusters, have reli-
able metallicities. RVW estimated [Fe/H]= -- 0.55 for the F
supergiant which we might reasonably assume also applies to
the Mira. For comparison, Table 3 gives the colours (from
WMF) of the Mira V3 in NGC 5927 which has a 311 day
period and [Fe/H]= -- 0.37 (Harris 1996).
c(cid:13) 2001 RAS, MNRAS 000, 1 -- ??
Table 3. Mean Colours of various components
star
J − H H − K K − L
J − K E(B − V )
(mag)
Mira
Mira
P = 312
N5927 V3
1.00
0.88
0.94
0.95
0.50
0.43
0.44
0.39
0.61
0.59
0.51
0.52
1.50
1.31
1.38
1.34
0.44
0.27
3.2 Mass-loss Rate
As RVW discussed, HD 172481 was detected by IRAS and
shows an obvious infrared excess. The colour corrected
12 µm mag is [12]=1.76, so K − [12] = 4.3 (4.2). This is
larger than values shown by any of the stars discussed by
WMF, but is within the range found for IRAS Miras by
Whitelock et al. (1994). This K − [12] would imply a mass-
M ∼ 3 × 10−6 M⊙yr−1 (Whitelock et al. 1994 fig
loss rate of
21) which is high for a 312 day Mira, but higher mass-loss
rates are usually associated with larger pulsation amplitudes
and again we note that the pulsation amplitude of this star
is high.
It would therefore seem possible that the dust shell,
which is responsible for large K −[12] and presumably for the
circumstellar extinction, originates entirely from the Mira.
While it is not essential to assume that any of the dust is
associated with the F supergiant, we cannot rule out the
possibility that some of it is. Neither can we rule out the
possibility of dust trapped in a circumbinary disk, as has
been proposed for some other supposed post-AGB binary
systems (e.g. Waters et al. 1993; Van Winckel 1999)
Note also that the IRAS colours of HD 172481 are
within the range found for D-type symbiotics (Whitelock
1988), illustrating the similarity of their dust shells (see sec-
tion 4).
3.3 Distance and Luminosity
The distance modulus of the Mira can be calculated via the
period luminosity relation (Whitelock & Feast 2000):
MK = −3.47 log P + 0.84,
to give (K0 − MK ) = 13.9 (13.8) or a distance of 6.0 (5.7)
kpc, and a height above the galactic plane of -- 1.1 ( -- 1.0) kpc.
The bolometric magnitude of the Mira can be es-
tablished by fitting a blackbody to the JHKL flux (see
Robertson & Feast 1981). This gives a bolometric magni-
tude at mean light of 9.30 (9.09) and at maximum light
of 8.72 (8.64). Or in absolute terms at mean light Mbol =
−4.59 (−4.69) and LM = 5.3 × 103 (5.8 × 103)L⊙.
4 THE F SUPERGIANT
RVW established that the relative luminosities of the F su-
pergiant and the M star are LF /LM = 1.8, for a reddening
of E(B − V ) = 0.44 (or LF /LM = 1.0 for E(B − V ) = 0.2).
The near-infrared measurement which they used to de-
fine the flux of the system was obtained on JD 2448850
(K = 5.48 Van Winckel, private communication), i.e. at
maximum light for the Mira. Thus the luminosity of the F
star is LF = 1.6 × 104 L⊙ (0.9 × 104 L⊙).
RVW argue that the F supergiant is a post-AGB star
c(cid:13) 2001 RAS, MNRAS 000, 1 -- ??
An Extremely Slow Nova?
3
on the basis of its high galactic latitude, large radial velocity
and moderate metal deficiency. In support of this they also
mention the dust, a variable Hα profile and a slight over-
abundance of s-process elements. Our identification of the
M star as a Mira variable confirms the binary as a low mass
system. There is, however, an alternative to the post-AGB
star scenario. The F supergiant could be a white dwarf, with
hydrogen-shell burning of material accreted from the Mira
wind providing a pseudo-photosphere and the observed lu-
minosity.
Iben & Tutukov (1996) discuss the evolution of mass ac-
creting white dwarfs in wide binaries systems, such as sym-
biotic stars. Of particular interest in the present context are
the D-type symbiotics where a white dwarf accretes from the
wind of a Mira variable in a binary system with a period in
excess of about 10 years (Whitelock 1987). In certain cases
the accretion rates are such that a thermonuclear runaway
gives rise to a very slow nova, e.g. RR Tel and V1016 Cyg.
These symbiotic stars, with their high excitation emission
lines, are apparently very different from HD 172481. How-
ever, it is clear from Iben & Tutukov that relatively small
differences, e.g. in the accretion rate of the white dwarf, can
give rise to very different phenomena.
The luminosity derived above, LF ∼ 104 L⊙, is that
expected for a cool ∼ 0.6 M⊙ white dwarf in a hydrogen shell
burning phase following a hydrogen shell flash (see Iben &
Tutukov fig. 5).
A very rough estimate of the accretion rate of the white
dwarf can be derived from Iben & Tutukov's equation 2. If
we assume the mass-loss rate from the Mira (section 3.2), a
0.6 M⊙ white dwarf, a total stellar mass (white dwarf plus
Mira) of 1.5 M⊙, a wind velocity of 20 km s−1 and an orbital
period of 10 years, the derived accretion rate onto the white
dwarf is 2 × 10−7M⊙yr−1.
It is clear from fig. 7 of Iben & Tutukov that this rate
would in fact result in steady hydrogen burning on the sur-
face of the white dwarf. However, for the parameters as-
sumed this rate should probably be regarded as an upper
limit, as Iben & Tutukov's equation 2 is based on Bondi-
Hoyle accretion and other estimates lead to significantly (up
to ten times) lower values (e.g. Theuns, Boffin & Jorissen
1996; Mastrodemos & Morris 1999). Furthermore, plausible
changes in the period or the wind velocity could increase or
decrease the rate. Nevertheless it seems likely that a white
dwarf paired with this Mira would accrete mass in such a
way as to either burn it steadily or in shell flashes, for a
large range of stellar separations.
In discussing the nova phenomenon Iben & Tutukov
point out that in wide binaries there is no reason to expect
the loss of the hydrogen envelope after the explosion. So
that it is possible for the newly created supergiant to last
for several decades. This might explain the present condition
of the F supergiant in HD 172481.
The high abundance of lithium in the supergiant spec-
trum (RVW) may well be key to understanding this sys-
tem. Arnould & Nørgaard (1975) and Starrfield et al. (1978)
predict the formation of large quantities of lithium during
hydrogen-burning themonuclear runaways. However, their
calculations are not obviously applicable to the situation
under discussion and they also predict an overabundance
of carbon and nitrogen which are not found in HD 172481
(RVW). More recently Hernanz et al. (1996) predict signif-
4
Patricia Whitelock & Freddy Marang
icant lithium production from some nova models, but these
seem to underestimate the ejecta masses. Clearly more the-
oretical work is required in this area.
ACKNOWLEDGMENTS
Our thanks to Hans Van Winckel for providing a preprint
of their paper and details of their unpublished infrared pho-
tometry. We are also grateful to Michael Feast for helpful
discussions. This paper is based on observations made from
the South African Astronomical Observatory and also made
use of information from the simbad data-base. The com-
ments and suggestions of an anonymous referee were also
helpful.
REFERENCES
Arnould M., Nørgaard H., 1975, AA, 42,55
Carter B. S. C., 1990, MNRAS, 242, 1
Bersier D., 1996, A&A, 308, 514
Feast M. W., Whitelock P. A., 1987, in: Late Stages of Stellar
Evolution, (eds.) S. Kwok, S. R. Pottasch, Reidel, Dordrecht,
p. 33
Harris W. E., 1996, AJ, 112, 1487
(http://physun.physics.mcmaster.ca/Globular.html)
Hauk B., Mermilliod M., 1998, A&AS, 129, 431
Hernanz M., Jos´e J., Coc A., Isern J., 1996, ApJ, 465, L27
Iben I., Tutukov A. V., 1996, ApJS, 105, 145
Johnson H. L., 1966, ARAA, 4, 193
Mastrodemos N., Morris M., 1999, ApJ, 523, 357
Odenwald S. F., 1986, ApJ, 307, 711
Reyniers M., Van Winckel H., 2000, A&A, in press,
astro-ph/0010486 (RVW)
Robertson B. S. C., Feast M. W., 1981, MNRAS, 196, 111
Starrfield S., Truran J. W., Sparks W. M., Arnould M., 1978,
ApJ, 222, 600
Theuns T., Boffin H. M. J., Jorissen A., 1996, MNRAS, 280, 1264
Van Winckel H., 1999, in: Asymptotic Giant Branch Star, IAU
Symp. 191, p. 465
Waters L. B. F. M., Waelkens C., Mayor M., Trams N. R., 1993,
A&A, 269, 242
Whitelock P. A., 1987, PASP, 99, 573
Whitelock P. A., 1988, in: The Symbiotic Phenomenon, (eds.) J.
Mikolajewska et al., Kluwer, Dordrecht, p. 47
Whitelock P. A., Feast M. W., 2000, MNRAS, 319, 759
Whitelock P. A., Menzies J. W., Feast M. W., Marang F., Carter
B., Roberts G., Catchpole R. M., Chapman J., 1994, MNRAS,
267, 711
Whitelock P. A., Marang F., Feast M. W., 2000, MNRAS, 319,
728 (WMF)
c(cid:13) 2001 RAS, MNRAS 000, 1 -- ??
|
astro-ph/0411308 | 1 | 0411 | 2004-11-11T22:10:16 | The Ultraviolet Galaxy Luminosity Function from GALEX data: Color Dependent Evolution at Low Redshift | [
"astro-ph"
] | We present measurements of the FUV (1530A) and NUV (2310A) galaxy luminosity functions (LF) at low redshift (z<0.2) from GALEX observations matched to the 2dF Galaxy Redshift Survey. We split our FUV and NUV samples into two UV-bj color bins and two redshift bins. As observed at optical wavelengths, the local LF of the bluest galaxies tend to have steeper faint end slopes and fainter characteristic magnitudes M* than the reddest subsamples. We find evidence for color dependent evolution at very low redshift in both bands, with bright blue galaxies becoming dominant in the highest redshift bin. The evolution of the total LF is consistent with an 0.3 magnitude brightening between z=0 and 0.13, in agreement with the first analysis of deeper GALEX fields probing adjacent and higher redshifts. | astro-ph | astro-ph | Draft version June 27, 2018
Preprint typeset using LATEX style emulateapj v. 6/22/04
4
0
0
2
v
o
N
1
1
1
v
8
0
3
1
1
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
THE ULTRAVIOLET GALAXY LUMINOSITY FUNCTION FROM GALEX DATA: COLOR DEPENDENT
EVOLUTION AT LOW REDSHIFT
Marie Treyer1,2, Ted K. Wyder1, David Schiminovich1, St´ephane Arnouts2, Tam´as Budav´ari4, Bruno Milliard2,
Tom A. Barlow1, Luciana Bianchi4, Yong-Ik Byun3, Jos´e Donas2, Karl Forster1, Peter G. Friedman1, Timothy
M. Heckman4, Patrick N. Jelinsky5, Young-Wook Lee3, Barry F. Madore6, Roger F. Malina2, D. Christopher
Martin1, Patrick Morrissey1, Susan G. Neff7, R. Michael Rich8, Oswald H. W. Siegmund5, Todd Small1, Alex
S. Szalay4, and Barry Y. Welsh5
Draft version June 27, 2018
ABSTRACT
We present measurements of the FUV (1530A) and NUV (2310A) galaxy luminosity functions (LF)
at low redshift (z ≤ 0.2) from GALEX observations matched to the 2dF Galaxy Redshift Survey.
We split our FUV and NUV samples into two UV−bj color bins and two redshift bins. As observed
at optical wavelengths, the local LF of the bluest galaxies tend to have steeper faint end slopes
and fainter characteristic magnitudes M∗ than the reddest subsamples. We find evidence for color
dependent evolution at very low redshift in both bands, with bright blue galaxies becoming dominant
in the highest redshift bin. The evolution of the total LF is consistent with an ∼ 0.3 magnitude
brightening between z ∼ 0 and 0.13, in agreement with the first analysis of deeper GALEX fields
probing adjacent and higher redshifts.
Subject headings: ultraviolet: galaxies -- galaxies: luminosity function, evolution
1. INTRODUCTION
The importance of estimating luminosity functions
(LF) as a function of redshift, color, environment, wave-
length, etc., to understand galaxy formation and evolu-
tion, has been well emphasized in the literature. Mul-
tivariate luminosity functions are crucial to constrain
theoretical models on small scales where the physics is
most complex. The shape of the galaxy LFs results from
non trivial physical processes (Benson et al. 2003, Bin-
ney 2004) and understanding their evolution is a chal-
lenging task for numerical simulations (e.g. Nagamine
et al. 2001). The Galaxy Evolution Explorer (GALEX)
mission has recently opened a new field of constraints
by allowing such statistical measures as LFs to be per-
formed in the rest-frame Ultraviolet (UV) in what is now
considered the low redshift Universe (z < 1 − 2), where
most of the evolution of galaxies is thought to have taken
place.
In this paper, we use data from the GALEX All-Sky
Imaging Survey matched to spectroscopic data from the
2dF Galaxy Redshift Survey to investigate the color
properties of local UV selected galaxies, and the color
dependence and evolution of their LF at very low red-
shift (z < 0.2). Whereas much effort has been devoted
1 California Institute of Technology, MC 405-47, 1200 East Cal-
ifornia Boulevard, Pasadena, CA 91125; [email protected]
2 Laboratoire d'Astrophysique de Marseille, BP 8, Traverse du
Siphon, 13376 Marseille Cedex 12, France
3 Center for Space Astrophysics, Yonsei University, Seoul 120-
749, Korea
4 Department of Physics and Astronomy, The Johns Hopkins
University, Homewood Campus, Baltimore, MD 21218
5 Space Sciences Laboratory, University of California at Berke-
ley, 601 Campbell Hall, Berkeley, CA 94720
6 Observatories of the Carnegie Institution of Washington, 813
Santa Barbara St., Pasadena, CA 91101
7 Laboratory for Astronomy and Solar Physics, NASA Goddard
Space Flight Center, Greenbelt, MD 20771
8 Department of Physics and Astronomy, University of Califor-
nia, Los Angeles, CA 90095
to estimate the evolution of galaxy light over the widest
and furthest range of redshift possible, very few surveys
have allowed to probe evolution at recent epochs (Love-
day 2004), and none so far at UV wavelengths. This work
is a companion paper to Wyder et al. (2004) who present
the total LFs of FUV and NUV selected galaxies in the
local (z ≤ 0.1) Universe (hereafter Paper I). It also com-
plements the work of Arnouts et al. (2004), Schiminovich
et al. (2004) and Budav´ari et al. (2004) who quantify
the evolution of the UV LF at higher redshift using the
GALEX Medium and Deep Imaging Surveys.
The data are summarized in Section 2. Section 3
presents the color properties of the samples. We es-
timate and discuss the dependence of the UV LFs on
galaxy color and redshift at recent epoch in Section
4. We assumed a flat ΛCDM cosmology with H0 =
70 km s−1Mpc−1, ΩM = 0.3 and ΩΛ = 0.7.
2. DATA
The GALEX field-of-view has a diameter of 1.2◦and
each pointing is imaged simultaneously in both the FUV
and NUV bands with effective wavelengths of 1530 A
and 2310 A, respectively. The GALEX instruments
and mission are described by Morissey et al. (2004) and
Martin et al. (2004).
The specific data analyzed in this paper are presented
in Paper I. In brief, they consist of 133 GALEX All-Sky
Imaging Survey pointings overlapping the 2dF Galaxy
Redshift Survey (Colless et al. 2001) in the South Galac-
tic Pole region. As the NUV images are substantially
deeper than the FUV, we used the NUV images for de-
tection and measured the FUV flux in the same aper-
ture as for the NUV. The apparent magnitudes were cor-
rected for foreground extinction using the Schlegel et al.
(1998) reddening maps and assuming the extinction law
of Cardelli et al. (1989).
The GALEX catalogs were matched with the 2dF cat-
alog using a search radius of 6′′. We restricted the sam-
2
Treyer et al.
6
5
4
3
2
1
0
0
1
0
1
Fig. 1. -- Color-color diagrams: F U V − N U V vs F U V − bj for
the FUV (left panel) and the NUV (right panel) selected samples
respectively. The thin solid lines are a set of models by Bruzual &
Charlot (2003) covering a range of star-formation histories, ages,
metallicities and extinctions. The thick dashed lines are dust free
models by Poggianti (1997), from Elliptical to Starburst.
ple to areas with effective exposure times tef f > 60 sec,
and removed sources with magnitude errors greater than
0.4 or contaminated by artifacts from bright stars. We
also removed overlap regions by restricting the coverage
to the inner 0.45◦ of each field. Finally, we excluded
GALEX sources where the 2dF redshift completeness
was less than 80% as well as areas of the sky excluded
from the 2dF input catalog. Our final working area of
GALEX-2dF overlap is 56.73 deg2 in both bands.
We applied a faint magnitude limit of 20 in both
bands corresponding to the bluest U V −bj color observed
and beyond which the redshift completeness begins to
drop. (N.B.: bj was shifted by −0.051 into the AB sys-
tem. All quoted magnitudes are AB). We also applied a
bright magnitude limit of 17 to avoid extended sources
with potential photometric problems. We visually in-
spected the resulting catalog of 2dF spectra and further
removed 27 objects with broad emission lines (the ob-
vious AGN/QSO's). Our final catalogs consist of 1292
FUV-selected galaxies and 1869 NUV-selected galaxies
with 17 ≤ U V ≤ 20 and available bj magnitude and
2dF redshift (with 2dF redshift quality parameter ≥ 3).
Both redshift distributions extends to z ∼ 0.25 (Paper
I). FUV fluxes are available for all but one of the galaxies
in the NUV-selected sample, although uncertainties can
be quite large for the faintest objects.
3. COLOR PROPERTIES
Figure 1 shows color-color diagrams (F U V − N U V vs
F U V − bj) for the FUV and NUV selected samples (left
and right panel, respectively). Typical error bars are
shown in the upper left corners. The thin lines are a
set of 42 spectral energy distribution (SED) models by
Bruzual & Charlot (2003) (hereafter BC), from old el-
lipticals (in red) to young irregular galaxies, assuming
a range of star-formation histories (τ = 1 Gyr to ∞),
ages (0.1 to 12 Gyr), metallicities (0.2 Z⊙ to Z⊙) and
Fig. 2. -- Model k-corrections in the bj, FUV and NUV bands.
The thin solid lines show our set of BC models and the thick dashed
lines the Poggianti SEDs, as in Fig. 1.
extinctions (AV = 0 to 1.5). The thick dashed lines are
dust free SED models by Poggianti (1997). The tracks
show the color evolution of the various galaxy types from
z = 0 to 0.25 (roughly counter-clockwise).
The bulk of both samples is consistent with bona fide
late-type galaxies, with only a handful of objects dis-
playing elliptical-like colors. We do not observe any
source with "U V -optical excess" colors (bluer than the
bluest model) such as were observed in the FOCA sam-
ple (Treyer et al. 1998, Sullivan et al. 2000), suggest-
ing that the so-called excess was to be found in the es-
timate of the UV flux of these objects rather than in
some physical explanation. On the other hand we find a
small fraction of galaxies with F U V −N U V colors some-
what bluer than any model (∼ 15% in the FUV selected
sample, ∼ 7% in the NUV selected sample). This is
partly due to newly measured, small but opposite offsets
in the FUV and NUV calibrations, adding up to +0.13 in
F U V −N U V (new in-flight calibrations, P. Morissey, pri-
vate communication), not applied in the current catalogs.
But shortcomings in the models at UV wavelengths can-
not be ruled out. Despite the wide coverage in parameter
space, the data are not fully accounted for (e.g. the red-
dest sources in both F U V −N U V and F U V −bj). Fig. 1
suggests that UV selected galaxies may require alterna-
tive model ingredients (e.g. more erratic star-formation
histories as suggested by Sullivan et al. 2003).
We assign a best fit SED (drawn from the combined
and interpolated BC and Poggianti libraries) to each
galaxy using its redshift and two independent colors
(F U V − N U V and F U V − bj). Figure 2 shows the
model k-corrections in the bj, FUV and NUV bands (thin
solid lines for BC and thick dashed lines for Poggianti).
UV k-corrections are small enough for the late-type, low-
redshift galaxies dominating the sample (as opposed to
k-corrections in optical bands) that uncertainties on the
assigned SEDs have negligible impact on the resulting
LFs.
We can also fit a power-law through the 2 band-
The UV Galaxy Luminosity Function at low redshift
3
Schechter function parameters (φ∗ is in Mpc−3).
TABLE 1
Band
z
Type
M∗
−α
log φ∗
-
-
All
-
-
-
-
-
−18.04 ± .11
FUV 0.0-0.1
−18.31 ± .06
0.1-0.2
0.0-0.1 Blue −17.89 ± .15
0.1-0.2
−18.44 ± .09
0.0-0.1 Red −18.19 ± .22
−18.15 ± .08
0.1-0.2
−18.23 ± .11
NUV 0.0-0.1
0.1-0.2
−18.53 ± .05
0.0-0.1 Blue −18.04 ± .16
0.1-0.2
−18.71 ± .08
0.0-0.1 Red −18.28 ± .18
0.1-0.2
−18.36 ± .06
-
-
-
-
-
All
-
-
-
-
-
-
-
1.22 ± .07 −2.37 ± .06
−2.33 ± .07
1.29 ± .09 −2.54 ± .09
−2.73 ± .10
1.10 ± .17 −2.80 ± .12
−2.50 ± .10
1.16 ± .07 −2.26 ± .06
−2.27 ± .05
1.25 ± .10 −2.48 ± .09
−2.82 ± .08
0.97 ± .14 −2.56 ± .09
−2.37 ± .06
-
-
-
passes and derive a spectral index for each galaxy (β =
(F U V −N U V )/0.425−2, assuming Fλ ∝ λβ). The mean
spectral indices of the FUV and NUV selected samples
are hβi ∼ −1.6 and −1.3 respectively. The UV slope
β has been shown to be a good tracer of dust attenua-
tion in galaxies undergoing a strong starburst, but not so
reliably in others (Bell 2002, Kong et al. 2004). In par-
ticular, Buat et al. (2004) found that local NUV selected
galaxies tend to have lower attenuations than starburts
for a given slope. As our two samples exhibit slopes cor-
responding to low attenuations in pure starburst galax-
ies, we conclude that they must be very little affected by
dust. (See Paper I for a quantitative discussion).
4. COLOR AND Z DEPENDENT LUMINOSITY FUNCTIONS
We split both samples into two rest-frame color bins:
(F U V − bj)0 ≤ and ≥ 2 (roughly corresponding to an Sb
spectral type and to the mean of the color distribution),
referred to as blue and red subsamples respectively. We
further split each of these subsamples into two redshift
bins: z ≤ 0.1 and 0.1 ≤ z ≤ 0.2, referred to as low
and 'higher' redshift subsamples respectively. The mean
redshifts in these two bins are 0.055 and 0.12 respectively
for the FUV sample, and 0.057 and 0.13 for the NUV
sample.
As in Paper I, the redshift completeness is defined as
the ratio of UV objects matched to a 2dFGRS counter-
part with (high quality) redshift to the total number of
UV galaxies in a given magnitude bin, as computed by
Xu et al. (2004) using the SDSS star/galaxy separation
criteria in overlap regions. It is roughly constant over the
range 17 ≤ U V ≤ 20 in both bands, with mean values
of ∼ 79% in the NUV and 92% in the FUV (cf Fig. 1 in
Paper I). We computed binned luminosity functions us-
ing the traditional Vmax estimator (Felten 1976) in the
12 subsamples (blue, red, and total at low and higher z
in both bands), and derived best fit Schechter functions
(Schechter 1976) for each of them. Since the magnitude
range in the highest redshift slice is too bright and nar-
row to constrain the slope meaningfully, we fixed it to
the low redshift value in each case.
Results are shown in Fig. 3 and the best-fit Schechter
function parameters are listed in Table 1. The FOCA
2000A LF (Sullivan et al. 2000) at a mean redshift
¯z ∼ 0.15 is also plotted for comparison (converted to the
AB magnitude system and to our H0 value but without
accounting for the difference in bandpasses). We refer to
Paper I for a discussion on the discrepancy between the
FOCA and GALEX LFs. As observed at optical wave-
lengths (Madgwick et al. 2002, Nakamura et al. 2003),
the local LFs of the bluest galaxies tend to have steeper
faint end slopes and fainter characteristic magnitudes M∗
than the reddest subsamples (although our definition of
red would qualify as blue in optically selected samples).
The fact that the latest types are observed to fainter ab-
solute magnitudes than the earlier types induces a small
bias in the estimate of the faint end slope of the full sam-
ple. The dotted lines in the upper panels of Fig. 3 show
that the sum of the blue and red LFs is slightly steeper
at the faint end than estimated for the galaxy popula-
tion as a whole, due to the shortage of red galaxies in
the faintest bins.
The total FUV and NUV LFs in the highest redshift
bin (lower panels in Fig. 3) are both consistent with ∼ 0.3
magnitude brightening in M∗ with respect to the LFs at
z < 0.1. This evolution is quite similar to that found in
optically selected samples at the same redshifts (Loveday
2004). Evolution can also be inferred from the steeply
rising V/Vmax distribution of both z-unlimited samples
(with hV/Vmaxi = 0.54 in both bands instead of the
0.5 expected from a non evolving population). As the
redshift distributions show two strong peaks around 0.06
and 0.11 (Paper I), the appearance of evolution could be
attributed to large scale structure effects, but removing
these two structures from the samples did not affect our
results.
Evolution also seems to occur predominantly in the
blue subsamples, with blue galaxies dominating the
bright end of the LF in the highest z bin. We checked
that no color trend was observed as a function of bj, as
might be expected if the 2dFGRS was strongly biased
against, e.g., faint blue low surface brightness galaxies
(LSBG), making bright blue galaxies seem dominant at
higher redshift. The issue of missing galaxies in the 2dF-
GRS, in particular LSBGs, has been addressed in a num-
ber of papers, most recently by Cross et al. (2004). The
authors find that the survey does indeed miss ∼ 15% of
the galaxy population, of which 6% only are identified
as LSBGs. If this particular class of objects is unlikely
to affect our results, the remaining missing galaxies may
bias our analysis, should their redshift distribution look
different from the one actually measured. A truly UV-
selected spectroscopic survey is underway to address this
question.
The present results are consistent with the FUV analy-
sis of a much deeper GALEX field probing adjacent and
higher redshifts (Arnouts et al. 2004, Schiminovich et
al. 2004). They are also in rough agreement with the rate
of evolution derived from a photometric-redshift study of
GALEX/SDSS data at low z (Budav´ari et al. 2004), al-
though our FUV LF in the range 0.1 ≤ z ≤ 0.2 seems
brighter than the photo-z estimate at overlapping z.
Both analysis agree very well in the NUV however (we
postpone attempts at explanations for the FUV discrep-
ancy until more comparable datasets are available). We
note that evolution seems to be limited by the observed
galaxy number counts, especially in the FUV band (Xu
et al. 2004), so that the rate of evolution detected at
z ≤ 1 in the GALEX data is expected to slow down.
This appears to be the case based on a photometric-
4
Treyer et al.
-1
-2
-3
-4
-5
-6
-2
-4
-6
-2
-4
-6
-2
-4
-6
-20
-18
-16
-14
-12
All
-1
-2
-3
-4
-5
-6
-2
-4
-6
-2
-4
-6
-2
-4
-6
-20
-18
-16
-14
-12
All
-21
-20
-19
-18
-17
-21
-20
-19
-18
-17
Fig. 3. -- Top panels : The FUV and NUV LFs at z ≤ 0.1 (left and right panel, respectively). The blue and red lines (and associated
filled circles and asteriks) represent the blue and red subsamples respectively. The black solid lines (and associated empty circles) show the
LFs of the full samples (Paper I), while the dotted lines are the sum of the blue and red Schechter functions. The green dotted line is the
FOCA 2000A LF at ¯z ∼ 0.15 (Sullivan et al. 2000). Bottom panels : The FUV and NUV LFs at 0.1 ≤ z ≤ 0.2. Same symbols and lines
as above.
redshift analysis of the HDF North and South probing
1.75 ≤ z ≤ 3.4 (Arnouts et al. 2004).
MT thanks the GALEX team at Caltech for its great
hospitality. GALEX is a NASA Small Explorer, launched
in April 2003. We gratefully acknowledge NASA's sup-
port for its construction, operation, and science analysis,
as well as the cooperation of the French Centre National
d'Etudes Spatiales and of the Korean Ministry of Science
and Technology.
REFERENCES
Arnouts S., et al. 2004, ApJ, submitted
Bell , E. F. 2002, ApJ, 577, 150
Benson, A. J., Bower, R. G., Frenk, C. S., Lacey, C. G., Baugh,
C. M., & Cole, S. 2003, ApJ, 599, 38
Binney, J. 2004, MNRAS, 347, 1093
Bruzual, G. & Charlot, S. 2003, MNRAS, 344, 1000
Buat V. et al. 2004, ApJ, submitted
Budav´ari T., et al. 2004, ApJ, submitted
Cardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245
Colless, M., et al. 2001, MNRAS, 328, 1039
Cross, N.J.G., et al. 2004, MNRAS, 349, 576
Felten, J. E. 1976, ApJ, 207, 700
Kong, X., Charlot, S., Brinchmann, J., & Fall, M. 2004,
astro-ph/0312474
Loveday, J. 2004, MNRAS, 347, 601
Madgwick, D. S., et al. 2002, MNRAS, 333, 133
Martin, D. C. et al. 2004, ApJ, submitted
Morissey, P. et al. 2004, ApJ, submitted
Nakamura, O., et al. 2003, AJ, 125, 1682
Nagamine, K., Fukugita, M., Cen, R., & Ostriker, J. P. 2001,
MNRAS, 327, L10
Poggianti, B. M. 1997, A&AS, 122, 399
The UV Galaxy Luminosity Function at low redshift
5
Schechter, P. 1976, ApJ, 203, 297
Schiminovich, D., et al. 2004, ApJ, submitted
Schlegel, D. J., Finkbeiner, D. P., & David, M. 1998, ApJ, 500, 525
Sullivan, M., Treyer, M. A., Ellis, R. S., Bridges, T., Milliard, B.,
& Donas, J. 2000, MNRAS, 312, 442
Sullivan, M., Treyer, M. A., Ellis, R. S., Mobasher, B. 2004,
MNRAS, 350, 21
Treyer, M. A., Ellis, R. S., Milliard, B., Donas, J., Bridges, T. J.
1998, MNRAS, 300, 303
Wyder, T. K., et al. 2004, ApJ, submitted
Xu, C.K., et al. 2004, ApJ, submitted
|
0808.0473 | 1 | 0808 | 2008-08-04T18:00:06 | Anomalous Emission from HII regions | [
"astro-ph"
] | Spinning dust appears to be the best explanation for the anomalous emission that has been observed at $\sim 10-60$ GHz. One of the best examples of spinning dust comes from a HII region in the Perseus molecular cloud. Observations of other HII regions also show tentative evidence for excess emission at frequencies $\sim 30$ GHz, although at lower emissivity levels. A new detection of excess emission at 31 GHz in the HII region RCW175 has been made. The most plausible explanation again comes from spinning dust. HII regions are a good place to look for spinning dust as long as accurate radio data spanning the $\sim 5-100$ GHz range is available. | astro-ph | astro-ph |
CMB Component Separation and the Physics of Foregrounds
14-18 July 2008, Pasadena, California
Anomalous emission from HII regions
Infrared Processing & Analysis Center, California Institute of Technology
Clive Dickinson1
Abstract
Spinning dust appears to be the best explanation for the anomalous
emission that has been observed at ∼ 10 − 60 GHz. One of the best ex-
amples of spinning dust comes from a HII region in the Perseus molecular
cloud. Observations of other HII regions also show tentative evidence for
excess emission at frequencies ∼ 30 GHz, although at lower emissivity
levels. A new detection of excess emission at 31 GHz in the HII region
RCW175 has been made. The most plausible explanation again comes
from spinning dust. HII regions are a good place to look for spinning
dust as long as accurate radio data spanning the ∼ 5 − 100 GHz range is
available.
1. Introduction
Anomalous emission appears to be "real" i.e. there is an additional
component of diffuse Galactic microwave emission that is not traditional
synchrotron, free-free or thermal dust emissions (e.g. Davies et al. 2006).
The anomalous emission emits strongly in the frequency range ∼ 10 −
60 GHz and is strongly correlated with FIR dust. A number of physical
mechanisms could be responsible for the anomalous emission including
hard synchrotron, hot free-free emission, cold dust, magneto-dipole radi-
ation, fullerenes, to name a few. However, so far, the best explanation for
the physical mechanism responsible for the anomalous emission appears to
be electro-dipole radiation from ultra-rapidly spinning dust grains, often
referred to as "spinning dust" (Draine & Lazarian 1998a,b).
2. Why HII regions?
HII regions may be a good place to look for anomalous emission.
Most importantly, they are typically associated with large column densi-
ties of dust, and therefore emit a detectable spinning dust signal. Even
though they emit strong free-free emission from the warm ionized gas
(Te ≈ 10000 K), this component can be quantified either by its well-
defined spectrum (a power-law with a spectral index βf f = −0.15 ± 0.05)
or alternatively, from recombination lines (e.g. Hα, RRLs) (Dickinson et
al. 2003). Draine & Lazarian (1998b) considered six typical environments
(CNM,WNM,WIM etc.) and they found that the the emissivity of spin-
ning dust is not strongly dependent on the details of the environment,
as shown in Fig. 1. There is of course a contrary argument regarding
small grain destruction in the environments around hot starts that form
HII regions, particularly the PAHs that could be spinning at fast rates.
However, this is usually only in the central region of a compact HII re-
gion and the dense dust population in the surrounding material (which
is usually not resolved by telescopes with beams >> arcmin) is typically
not strongly affected. It is interesting to note that the dust temperature
1
CMB Component Separation and the Physics of Foregrounds
14-18 July 2008, Pasadena, California
is typically 30-50 K, compared to the ≈ 18 K dust found in the diffuse
ISM.
Figure 1: Model dust emissivities from Draine & Lazarian (1998) with previous
observational data, reproduced from Finkbeiner et al. (2004). Note how the
spinning dust models for different environments (CNM, WNM, WIM etc.) peak
in the range ∼ 20 − 40 GHz and typically vary by only a factor of a few in
emissivity.
3. Detections from "compact" Galactic objects
The best two examples of spinning dust are the HII region G159.6-
18.5 in the Perseus molecular cloud (Watson et al. 2005) and the dark
cloud LDN1622 (Finkbeiner et al, 2002,2004; Casassus et al. 2006). Both
of these show spectra that are consistent with canonical spinning dust
models. Furthermore, LDN1622 shows a better morphological correlation
with the shorter wavelengths of IRAS (12/25 µm), as expected from the
smaller dust grains.
Of the two spinning dust candidates of Finkbeiner et al. (2002), the
strongest candidate was initially LPH[96]201.663+1.643. Follow-up obser-
vations with the Cosmic Background Imager (CBI) showed that the flux
density at 31 GHz was completely consistent with optically thin free-free
emission. The original Finkbeiner et al. (2002) data at 5-10 GHz and
the CBI 31 GHz data point can still be fit by an anomalous component
with a very narrow spectrum (N. Ysard, priv. comm.). However, more
recent GBT 100 m observations (D. Finkbeiner, priv. comm.) could not
2
CMB Component Separation and the Physics of Foregrounds
14-18 July 2008, Pasadena, California
Table 1: Comparison of 100 µm dust emissivities for HII regions and cooler dust
clouds, from data at or near 30 GHz. Emissivities, in units µK (MJy/sr)−1, have
been normalised to 31 GHz.
Source
Dust emissivity Reference
µK (MJy/sr)−1
HII regions
6 HII regions (mean)
LPH96
Cool dust clouds
15 regions WMAP
All-sky WMAP
LDN1622
G159.618.5
3.3 ± 1.7
5.8 ± 2.3
Dickinson et al. (2007)
Dickinson et al. (2006)
11.2 ± 1.5
10.9 ± 1.1
24.1 ± 0.7
17.8 ± 0.3
Davies et al. (2006)
Davies et al. (2006)
Casassus et al. (2006)
Watson et al. (2005)
reproduce the original Green Bank 5 − 10 GHz data, suggesting that the
earlier data were spurious.
Dickinson et al. (2007) observed 6 bright southern HII regions with the
CBI at 31 GHz to search for anomalous emission. Using low frequency
data from the literature, a simple free-free power-law was fitted to the
spectrum of each object. The 31 GHz flux density appeared to be slightly
above the simple free-free extrapolation. The average dust emissivity,
relative to the 100 µm map, was found to be 3.3 ± 1.7 µK/(MJy/sr). This
is significantly lower than values observed at high latitudes and other
anomalous regions (Table 1). On the other hand, Scaife et al.
(2007)
observed a sample of 16 HII regions with the Arc Minute Imager (AMI)
at ≈ 15 GHz, and found no significant evidence for excess emission with
many regions emitting at < 1 µK/(MJy/sr).
4. New detection of spinning dust in RCW175
Preliminary results from the Very Small Array (VSA) Galactic
plane survey at 33 GHz (Todorovic et al, in prep.) indicated that the
RCW175 HII region was significantly brighter than expected from
lower frequency surveys. It was re-observed with higher resolution
with the CBI at 31 GHz and a consistent flux density was measured
(Dickinson et al. 2008). The integrated flux density spectrum of
RCW175 is shown in Fig. 2. Best-fitting models for free-free and
thermal dust emission are plotted along with a typical spinning dust
model. The upper limit at 94 GHz from WMAP data is very impor-
tant for the interpretation of the apparent excess since it effectively
rules out significant contributed from thermal dust or optically thick
free-free emission, perhaps from an ultracompact HII region. The
estimated column density from IRAS 100 µm maps gives a spinning
dust emissivity that is consistent with the Draine & Lazarian (1998)
models.
3
CMB Component Separation and the Physics of Foregrounds
14-18 July 2008, Pasadena, California
Figure 2: Integrated flux density spectrum for RCW175 (Dickinson et al. 2008).
Filled circles represent data fitted by a power-law with fixed spectral index
(dotted line) and a modified black-body with fixed emissivity (dashed line).
The Draine & Lazarian (1998b) CNM spinning dust spectrum (dot -dashed
line) has been fitted to the 31/33 GHz data.
5. Conclusion
Anomalous emission has been observed at both high latitudes
and from discrete Galactic objects. The best explanation is that it
is due to spinning dust grains. One of the best examples of spinning
dust is the HII region G159.6-18.5 in the Perseus molecular cloud.
There is also tentative evidence for excess emission from other HII
regions.
In particular, RCW175 has now been confirmed to have
excess emission at ∼ 30 GHz with spinning dust being a plausible
origin. HII regions are good places to search for spinning dust emis-
sion, but accurately calibrated data covering the frequency range
∼ 5 − 100 GHz is required.
References
Casassus, S., et al., 2006, ApJ, 639, 951
Davies, R.D., et al., 2006, MNRAS, 370, 1125
Dickinson, C., Davies, R.D., Davis, R.J., 2003, MNRAS, 341, 369
Dickinson, C., et al., 2006, ApJ, 643, L111
Dickinson, C., et al., 2007, MNRAS, 379, 297
Dickinson, C., et al., 2008, ApJL, submitted (arXiv:0807.3985)
Draine, B.T., Lazarian, A., 1998a, ApJ, 494, L19
Draine, B.T., Lazarian, A., 1998b, ApJ, 508, 157
Finkbeiner, D., et al., 2002, ApJ, 566, 898
Finkbeiner, D., Langston, G.I., Minter, A.H., 2004, ApJ, 617, 350
Scaife, A., et al., 2007, MNRAS, 385, 809
Watson, R.A., et al., 2005, ApJ, 624, L89
4
|
astro-ph/0302152 | 2 | 0302 | 2003-02-13T19:07:26 | Optical and Far-UV Spectroscopy of Knot D in the Vela Supernova Remnant | [
"astro-ph"
] | We present spectra of optical filaments associated with the X-ray knot D in the Vela supernova remnant. It has been suggested that Knot D is formed by a bullet of supernova ejecta, that it is a break-out of the shock front of the Vela SNR, and also that it is an outflow from the recently discovered remnant RXJ0852.0-4622. We find that Knot D is a bow shock propagating into an interstellar cloud with normal abundances and typical cloud densities (n_H ~ 4-11 cm^-3). Optical longslit spectra show that the [S II] 6716,6731 to Halpha line ratio is greater than unity, proving that the optical filaments are shock excited. The analysis of far-ultraviolet spectra obtained with the Hopkins Ultraviolet Telescope and with the Far Ultraviolet Spectroscopic Explorer (FUSE) LWRS aperture show that slower shocks (~100 km s^-1) produce most of the low ionization lines such as O III] 1662, while faster shocks (~180 km s^-1) produce the O VI 1032,1038 and other high ionization lines. C III and O VI lines are also detected in the FUSE MDRS aperture, which was located on an X-ray bright region away from the optical filaments. The lines have two velocity components consistent with ~150 km s^-1 shocks on the near and far sides of the knot. The driving pressure in the X-ray knot, P/k ~ 1.8E+7 cm^-3 K, is derived from the shock properties. This is over an order of magnitude larger than the characteristic X-ray pressure in the Vela SNR. The velocity distribution of the emission and the overpressure support the idea that Knot D is a bow shock around a bullet or cloud that originated near the center of the Vela remnant. | astro-ph | astro-ph |
Optical and Far-UV Spectroscopy of Knot D in the Vela
Accepted for publication in the ApJ
Supernova Remnant
Ravi Sankrit, William P. Blair
The Johns Hopkins University
Department of Physics and Astronomy, 3400 N. Charles St., Baltimore, MD 21218
[email protected], [email protected]
and
John C. Raymond
Harvard-Smithsonian Center for Astrophysics, 60, Garden St., Cambridge, MA 02138
[email protected]
ABSTRACT
We present spectra of optical filaments associated with the X-ray knot D in the
Vela supernova remnant. It has been suggested that Knot D is formed by a bullet
of supernova ejecta, that it is a break-out of the shock front of the Vela SNR, and
also that it is an outflow from the recently discovered remnant RXJ0852.0-4622.
We find that Knot D is a bow shock propagating into an interstellar cloud with
normal abundances and typical cloud densities (nH ∼ 4 -- 11 cm−3). Optical
longslit spectra show that the [S II] λλ6716,6731 to Hα line ratio is greater than
unity, proving that the optical filaments are shock excited. The analysis of far-
ultraviolet spectra obtained with the Hopkins Ultraviolet Telescope and with
the Far Ultraviolet Spectroscopic Explorer (FUSE) LWRS aperture show that
slower shocks (∼ 100 km s−1) produce most of the low ionization lines such as
O III] λ1662, while faster shocks (∼ 180 km s−1) produce the O VI λλ1032,1038
and other high ionization lines. C III and O VI lines are also detected in the
FUSE MDRS aperture, which was located on an X-ray bright region away from
the optical filaments. The lines have two velocity components consistent with
∼ 150 km s−1 shocks on the near and far sides of the knot. The driving pressure
in the X-ray knot, P/kB ∼ 1.8×107 cm−3 K, is derived from the shock properties.
-- 2 --
This is over an order of magnitude larger than the characteristic X-ray pressure
in the Vela SNR. The velocity distribution of the emission and the overpressure
support the idea that Knot D is a bow shock around a bullet or cloud that
originated near the center of the Vela remnant.
Subject headings: ISM: individual (Vela) -- ISM: supernova remnants -- shock
waves
1.
Introduction
In an X-ray image obtained with ROSAT, the Vela Supernova Remnant (SNR) is roughly
circular with a diameter ∼ 8◦ (Aschenbach, Egger & Trumper 1995). At a distance of 250 pc
(Cha, Sembach & Danks 1999) this corresponds to about 35 pc. Six bow shaped "knots"
of emission lie beyond the nominal circumference of the remnant. The shape and location
of these knots led Aschenbach et al. to suggest that they were due to ejecta "bullets" that
had overtaken the blast wave. Of the six features, the closest to the remnant and also
the brightest in X-ray is "Knot D". The optical nebula RCW 37 (Rodgers, Campbell &
Whiteoak 1960), lies along the outer edge of Knot D. This association with bright optical
filaments also distinguishes Knot D from the other five knots.
Although it is grouped with the ejecta bullets, there are two other explanations for the
origin of Knot D. One scenario, suggested by Plucinsky et al. (2002) is that Knot D is a
shock that has broken out from the Vela SNR due to inhomogeneities in the ambient medium.
They analyzed Chandra-ACIS spectra of Knot D and found that the oxygen abundance was
solar and that the neon abundance was modestly enhanced, and they found no evidence
for abundance differences at different locations in the knot. These results as well as the
morphology of the X-ray emission led Plucinsky et al.
(2002) to favor a shock break-out
origin for Knot D. The other suggestion (Redman et al. 2000, 2002) is that Knot D is
associated with RXJ0852.0-4622, a remnant that in projection is within the boundaries of
the Vela SNR. They analyzed optical echelle data of RCW 37 and found a velocity split near
the bright optical edge in both [S II] and [O III] emission, and an almost complete velocity
ellipse across the nebula in the latter. They inferred that the geometry of the optical nebula
is either an incomplete funnel or a wavy sheet. Based on the kinematic structure of emission
as well as the optical and X-ray morphologies of the nebula they suggested that an outflow
from RXJ0852.0-4622 impacted the pre-existing wall of the Vela SNR to produce Knot D.
As they point out, for their model to work both remnants have to be at about the same
distance (Redman et al. 2002).
-- 3 --
Regardless of its origin, it is known that Knot D is a source of shock excited emis-
sion. Blair, Vancura & Long (1995) presented far-ultraviolet observations of the region
obtained with the Voyager 2 Ultraviolet Spectrometer. Strong emission from C III λ977
and O VI λλ1032,1038 was detected in the spectrum. The authors noted the sheet-like mor-
phology of the filaments and the spatial coincidence of the X-ray and optical emission and
argued that these were characteristics of a shock-cloud interaction (such as had been seen
in the Cygnus Loop SNR) and that this interaction produced the ultraviolet lines as well as
the optical emission. They suggested that the large Voyager FOV included shocks with a
range of velocities, and that ∼ 120 km s−1 shocks were responsible for the C III emission
and 160 -- 300 km s−1 shocks produced the O VI emission.
In this paper we take a detailed look at the shock-cloud interaction and determine
the shock wave paramaters. Our results throw some light on the nature of Knot D, but
do not unambiguously determine its origin. We use ultraviolet spectra obtained with the
Hopkins Ultraviolet Telescope (HUT) and with the Far Ultraviolet Spectroscopic Explorer
(FUSE). These data have much higher spectral resolution than the Voyager spectrum and
the aperture FOVs are over a thousand times smaller. We also present optical longslit spectra
of the filaments. We confirm that the optical emission is shock excited. By comparing the
measured far-ultraviolet emission line strengths with shock model predictions, we find the
properties of the shocks around Knot D. The kinematic distribution of the emission around
the bow shock is clearly revealed in the high resolution FUSE data. The observations are
described in §2, and the results presented in §3. Then, in §4 we analyze the results and
present our interpretation. The last section, §5 summarizes the central results of this work
and presents some of the issues yet to be answered.
2. Observations
Images of the region of interest are shown in Figure 1. The top panel is a three color
image of the southern half of Knot D. Narrowband Hα and [O III] λ5007 images (first
presented by Blair, Vancura & Long (1995)) are shown in red and green, respectively. The
blue is the Chandra X-ray image, presented by Plucinsky et al. (2002) (and kindly provided
to us by the lead author). Overlaid on the image are the locations of the FUSE LWRS
and MDRS apertures. The bottom panel is a two color image of a region around the bright
filament. Only the optical emission is shown (Hα in red, [O III] in green) overlaid with the
HUT and FUSE LWRS aperture positions.
The HUT observations were made during the Astro-2 space shuttle mission on 1995
March 15. (The basic HUT design is described by Davidsen et al. (1992), and the improve-
-- 4 --
ments to the instrument for Astro-2 by Kruk et al.
(1995).) The wavelength coverage of
HUT was 820 -- 1840A with a resolution of ∼ 3A. The 56′′×10′′ aperture was centered at
α2000 = 09h00m24.s30, δ2000 = −45◦ 54′ 31.′′5, and placed at a position angle of 18◦ which
aligned the aperture with the optical filament. After approximately 1000 s of exposure time
at this position, the aperture was offset to a location ∼30′′ East, perpendicular to the aper-
ture's long dimension. We refer to these positions as P1 and P2, respectively (Figure 1b).
The data were obtained in a time-tag mode which allowed the separation of orbital day and
night photons. Since the daytime photons are contaminated by airglow emission, only the
nighttime data are used here. The effective exposure times are 932 s for P1 and 1162 s for
P2.
The HUT data were processed using an IRAF1 package originally developed to process
HUT data from the Astro-1 mission, and described by Kruk et al.
(1999). The flux cal-
ibration applied to these data is based on in-orbit observations of white dwarf stars fitted
with theoretical models. Other corrections, such as pulse persistence, dark count and Lyα
scattering have been characterized and included here (Kruk et al. 1995, 1999).
The FUSE observation (Program ID B1080201) was obtained on 2001 April 2 as part of
a Guest Investigator project to study the Vela-Puppis region. Eleven exposures with a total
integration time of 14,133 s were obtained with the low resolution (LWRS) 30′′×30′′ aperture
centered at α2000 = 09h00m24.s30, δ2000 = −45◦ 54′ 31.′′5. This position is the same as the
center of the HUT P1 observation, and lies on the bright optical filament (Figure 1b). Data
were obtained simultaneously through the medium resolution (MDRS) 20′′×4′′ aperture.
The MDRS aperture was located 208′′ away, 24◦ West of North from the LWRS aperture on
an X-ray bright region away from the bright optical emission (Figure 1a).
The wavelength range covered by FUSE is 905 -- 1187A. In this paper we present data
from segments SiC2A (∼905 -- 1000A), LiF1A (∼1000 -- 1100A) and LiF2A (∼1100 -- 1187A),
the segments with largest effective area in their wavelength ranges (Moos et al. 2000). The
data in other segments were used for comparison to distinguish between real features and
possible detector artifacts. The raw data from all the exposures were combined and the
pipeline CalFUSE version 1.8.7 was used to produce calibrated spectra. A shift was applied
to the LiF1A and SiC2A flux vectors to line up the geocoronal emission (Lyβ and Lyγ
respectively for the two channels) at the appropriate wavelength in the heliocentric frame
of reference. This was done for both LWRS and MDRS spectra. The source is an extended
1IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the As-
sociation of Universities for Research in Astronomy, Inc., under cooperative agreement with the National
Science Foundation.
-- 5 --
emission object and fills the apertures. For such a target, the effective spectral resolution
is a combination of the instrumental resolution (∼0.05A), the slit width, and a further
degradation caused by detector astigmatism (Sahnow et al. 2000). The width of the airglow
lines in a spectrum is a measure of the effective spectral resolution. In our data, the effective
spectral resolutions are ∼0.34A for the LWRS data and ∼0.09A for the MDRS data. (These
numbers correspond to about 100 km s−1 and 30 km s−1, respectively, at 1000A.)
The longslit optical spectra presented here were obtained in 1989 February at Las Cam-
panas Observatory. The observations used the du Pont 2.5 m telescope, the Modular Spec-
trograph and an 800 × 800 pixel TI CCD. A 600 line mm−1 5000A blaze grating was used
with an 85 mm camera, providing spectral coverage, 4750 -- 7150A at 3.0A per pixel. The
slit width was 2′′ and the effective spectral resolution was 8A. The spatial scale along the slit
was 0.′′8 pixel−1, with a usable slit length of 7′. The slit was placed along the East-West di-
rection, passing through the optical filaments ∼ 1′′ below the slit center of the HUT position
P1.
The data were reduced using standard IRAF procedures, including bias subtraction,
flat fielding, image rectification and extraction of one dimensional spectra. The background
was determined from a portion of the slit lying outside the remnant and subtracted from
the spectra. The flux calibration was done using observations of standard stars from the list
of Stone & Baldwin (1983). Fiducial stars along the slit, and the known spatial scale were
used to extract regions of the long slit data corresponding to the intersection of the slit with
the HUT apertures. The optical images (Figure 1b) show that the emission is fairly uniform
within the apertures, so the optical spectra should be representative of the overall emission
within the HUT apertures at P1 and P2.
3. Discussion
3.1. Optical Imagery and Spectra
The optical narrowband images show a striated emission structure. There are bands of
green (strong [O III]), yellow (strong [O III] and Hα) and red (strong Hα) running parallel
to the shock front (Figure 1b), indicating a systematic variation of the Hα to [O III] flux
ratio. The Hα emission is at its brightest in the region between about 50′′ and 70′′ behind
the leading [O III] edge. The [O III] emission reaches its peak at about the same location
as the Hα (∼ 50′′ behind the edge). In the green ([O III] dominated) band of emission, the
[O III] flux is about 40% of its peak value, while the Hα flux remains below 20% of its peak
value.
-- 6 --
The brightest optical filaments lie approximately parallel to the Southeast edge of the
X-ray emitting region (Figure 1a) and there is a gap of about 1′ between the X-ray edge and
the filaments. Observations of other shock-cloud interactions in the Cygnus Loop (Hester &
Cox 1986; Danforth et al. 2000) and in Vela (Raymond et al. 1997; Sankrit et al. 2001)
have shown that optical filaments are typically coincident with the edges of bright X-ray
features, and not separated as in Knot D. So though the association of X-ray and optical
regions is expected in SNR shocks, the existence of a gap between the two is surprising.
The optical spectra of the regions of bright and faint Hα (P1 and P2, respectively) are
shown in Figure 2. We measured the integrated line fluxes using the IRAF task, splot,
and calculated the surface brightnesses. We derive the color excess from the observed ratio
between between the Hα and Hβ fluxes. The intrinsic ratio between the two lines for shock
excited emission is 3.0 (Raymond 1979). The observed ratio at P1, IHα/IHβ ≃ 3.4, implies
that EB−V = 0.1, a value consistent with those obtained for other regions in Vela (Wallerstein
& Balick 1990). At position P2, we only have an upper limit for the Hβ flux and cannot
estimate the color excess. Therefore we assume that the P1 value applies to the whole region.
We used the extinction curve suggested by Fitzpatrick (1999), and total to visual selective
extinction R = 3.1 to calculate correction factors for all the lines. The surface brightnesses
of the optical lines, corrected for interstellar extinction are presented in Table 1.
The brightness ratio, I[S II]/IHα is greater than 1 at both positions P1 and P2. The high
ratios imply that the gas is collisionally excited and confirm that the optical filaments are
due to shocks. As expected from the narrowband images, the optical spectra at the two
positions differ significantly in the strength of [O III] emission relative to emission from the
lower ionization lines. The [O III] λλ4959,5007 flux is about 3 times weaker at position P2
than at P1, whereas all the other lines are about an order of magnitude weaker at P2. Some
lines, including [O I] λλ6300,6364 are not detected at the latter position. This difference
suggests that at P2, the shock is "incomplete" - it has not yet swept up enough material
for the post-shock gas to have recombined and produced strong Hα, [S II] and other low
ionization lines.
3.2. HUT Spectra
The HUT spectra of the two positions in Knot D contain numerous emission lines
expected from SNR shocks (Figure 3). As expected from the Voyager UVS spectrum (Blair,
Vancura & Long 1995), C III λ977 and O VI λλ1032,1038 are the strongest lines below
1200A. Other lines such as N III λ991, which were lost in the low resolution Voyager
spectrum, are detected in these HUT data. Longward of 1200A, the strongest lines are
-- 7 --
C IV λ1549, O III] λ1663 and N III] λ1751. Many (but not all) of the detected lines
are marked in Figure 3. One surprise is that N V λ1240 is relatively weak:
in spectra of
radiative shocks in the Cygnus Loop the line is comparable in strength to the C IV line
(Blair et al. 1991).
The integrated line fluxes in the spectra were measured using the IRAF tasks splot
and the more sophisticated specfit (Kriss 1994), the latter for blended lines. The pair of
lines, Si IV λ1402 and O IV] λ1403 could not be deblended automatically. Therefore, it was
assumed that the Si IV λ1402 flux was exactly half the Si IV λ1394 flux (valid for optically
thin emission). This amount was subtracted from the measured flux of the 1403A feature
and added to the 1394A line. The surface brightnesses of the lines, corrected for interstellar
extinction, are listed in Table 2 along with the correction factors. It should be noted that the
interstellar extinction curve at wavelengths below 1000A is poorly known and is the main
source of uncertainties in the intrinsic brightness of the shortest wavelength lines.
The line strengths at P1 are about twice those at P2, but apart from that overall
factor the spectra at the two positions are similar. The dramatic differences in the relative
intensities of optical lines are not present for the ultraviolet lines. This implies that the
shock conditions are similar at the two positions. For instance IO VI1032,1038/IO III]1664, a ratio
that is very sensitive to the maximum temperature reached behind the shock, is ∼ 1.3 at P1
and ∼ 1.2 at P2. Thus the far-ultraviolet spectra support the idea that the difference in the
optical emission between P1 and P2 is due to different levels of shock completeness, rather
than differences in shock velocity. The main difference between the HUT P1 and P2 spectra
is the ratio between the two O VI doublet lines: I1038/I1032 ≃ 0.76 at P1 and ≃ 0.61 at
P2. The ratio approaches 0.5 when the line optical depth tends to zero, and approaches 1.0
when the line is optically thick. The measured ratios show that the line fluxes are affected
by resonance line scattering and that the effect is stronger at P1 than at P2.
The ratio between the O VI line fluxes can be affected by absorption of the 1038A line
by molecular hydrogen (the Lyman band transitions, R(1)5−0 and P(1)5−0, at 1037.15A and
1038.16A, respectively) and by C II* (1037.02A). The ratio can also be affected by self-
absorption of the lines by O VI in the ISM. Jenkins, Silk & Wallerstein (1976) find the
column of low velocity H2 in the J=1 state NJ=1 ≃ 1014.5, 1015.5, and 1017.8 towards three
stars behind Vela. Even the highest of these would absorb only about 9% of the flux from
a 150 km s−1 width line centered at zero velocity. The more typical lower values for the
column would have negligible effect on the O VI λ1038 flux. The C II* line is stronger, but
is at −173 km s−1 relative to the O VI λ1038 rest wavelength. The velocity profile of the
O VI lines in the FUSE LWRS spectrum (see below) shows that C II* absorption would not
reduce the O VI flux significantly. The O VI column out to the distance of Vela (250 pc)
-- 8 --
is ∼ 1013 cm−2 (Jenkins et al. 2001), which would have negligible affect on the line fluxes.
Also, Knot D lies beyond the remnant boundary, and it is unlikely that other shocked clouds
associated with Vela lie in front of it. Therefore, the ratios measured in the HUT spectra
should not be modified significantly by foreground absorption.
3.3. FUSE Spectra
3.3.1. LWRS Spectrum
The FUSE LWRS spectrum is shown in Figure 4. The strongest SNR lines in the FUSE
bandpass are C III λ977 and O VI λλ1032,1038. These lines are allowed to go off-scale in
the plot in order to show a number of the weaker lines present in the spectrum. The weak
lines are well separated from each other and are easily identified. Furthermore, SNR lines
can be distinguished from airglow lines by comparing the total spectrum with a spectrum
screened to include only orbital night data. Many of these weaker lines have been detected
in SNR spectra for the first time by FUSE. For example, this paper reports the first such
detection of S III λ1077. Note that the region around He II+[N II] λ1085 is not included in
the plots. This range lies at the edge of only one detector segment (Sahnow et al. 2000),
and the telescope effective area is very small in this region.
The total line fluxes in the LWRS spectrum were obtained by trapezoidal integration
of the flux vector with suitable background subtraction. The strong lines contain several
thousand counts, so the random errors in their measured fluxes are low (∼ 1%). The errors
are about 6% for some of the weaker lines, such as S III λ1015 and S IV λ1063. For the
weakest line fluxes, particularly those detected in the SiC 2A segment (e.g. S VI λλ933, 944),
uncertainties in the background placement contribute to the total error, which is about 20%.
The uncertainty in the absolute flux calibration is about 10% (Sahnow et al. 2000), which
dominates the statistical errors in the measured fluxes of the strong lines and is comparable
to those of the weak lines. The surface brightnesses of the lines, corrected for interstellar
extinction, are presented in Table 3.
The fluxes of the strong lines (C III and O VI) plotted on a velocity scale are shown in
Figure 5. (The velocity is relative to the Local Standard of Rest, in which frame the sun is
moving away from Vela at about 12.5 km s−1.) The top panel is an overlay of the two O VI
lines. The lines are centered at ∼ +18 km s−1 and the FWHM of each line is ∼ 130 km s−1.
If the emission fills the slit uniformly, then the observed line profile is the intrinsic profile
convolved with the line spread function, which, to very good approximation, is a 106 km s−1
wide "tophat" function. Under this assumption the O VI lines have intrinsic FWHMs of
-- 9 --
∼ 100 km s−1. The surface brightnesses of the O VI in the FUSE LWRS spectrum lies
between the HUT P1 and P2 values (Tables 2, 3). However, the ratio I1038/I1032 ≃ 0.58 in
the FUSE LWRS spectrum, which is lower than the value at either HUT position. Thus,
when detected over the more spatially extended LWRS aperture, resonance scattering has a
smaller effect on the O VI emission than when detected over the HUT aperture at P1 or P2.
The feature on the blue wing of O VI λ1038 is identified as C II λ1037.02, a line also
detected in a FUSE spectrum of an X-ray bright knot in the center of Vela (Sankrit et
al. 2001). The transition leading to the 1037.02A line has a lower state 63 cm−1 above the
ground state. The companion line, C II λ1036.34 is a strong ground state transition (Morton
1991) and is not detected because it is absorbed by the interstellar gas between us and the
SNR.
The bottom panel of Figure 5 shows an overlay of the C III and the O VI λ1032 lines.
The C III line is significantly broader than the O VI line, and it has a central reversal. By
fitting the wings of the C III profile with gaussians we find that the observed peaks are
centered at ∼ −60 km s−1 and ∼ +80 km s−1, and that the components have FWHMs
∼ 100 km s−1. C III is a strong resonance line, and while the observed profile is consistent
with there being two kinematic components, we expect optical depth effects to influence the
line profile. Some idea about the intrinsic velocity distribution of low ionization emission is
gleaned from the [S II] λ6716 position-velocity data presented by Redman et al. (2000). In an
echelle spectrum with the slit cutting across the optical filament about 2′ south of the FUSE
LWRS position, Redman et al. (2000) detected bright [S II] emission around zero-velocity.
In addition to this central component, they found fainter red-shifted and blue-shifted peaks
(see their Figure 4). The intrinsic velocity structure of the C III emission is likely to be as
complex, in which case the reversal in the observed profile is due to zero-velocity emission
being self-absorbed by both filament material and the ISM along the sightline.
3.3.2. MDRS Spectrum
C III λ977 and the O VI doublet are detected in the FUSE MDRS spectrum, obtained at
a position several arcminutes behind the bright optical filaments. The line fluxes are plotted
against LSR velocity in Figure 6. The two O VI lines track each other closely (Figure 6 top
panel). Each O VI line profile has two peaks, centered at ∼ −80 km s−1 and ∼ +80 km s−1.
The width of the gap between the two components is about 90 km s−1, which is much
greater than the typical velocity width ∼ 40 km s−1 for absorption by interstellar O VI
(Jenkins et al. 2001). We conclude that we are seeing two distinct emitting components
(rather than a single, self-absorbed component) along the line of sight at this location. The
-- 10 --
kinematic structure of the C III emission follows that of the O VI emission (Figure 6 bottom
panel). The C III line has two peaks, centered at the same velocities as the O VI lines. The
separation between the two C III velocity components is larger in the the MDRS spectrum
than in the LWRS spectrum. Because of its bow shape, and because of the lack of complex
optical morphology at the location of the MDRS aperture (Figure 1), we conclude that the
two emission components are due to shocks on the front and back sides of Knot D. The
blue-shifted O VI lines have FWHMs of about 70 km s−1, and are broader than their red-
shifted counterparts, which are about 50 km s−1 wide (FWHM). Both C III components
have FWHMs of about 50 km s−1.
The surface brightnesses, corrected for interstellar extinction, of the blue and red shifted
components of the three lines are presented in Table 4. The blue shifted component of
C III λ977 is about 1.3 times as bright as the red shifted component. In contrast, the total
O VI flux of the blue shifted component is just half that of the red shifted component.
IO VI/IC III is about 0.9 for the red shifted component and about 0.4 for the blue shifted
component. The ratios between the doublet lines, I1038/I1032 are 0.54 for the blue shifted
component and 0.56 for the red shifted component. These ratios are closer to the optically
thin limit compared with the values obtained for the HUT and FUSE LWRS spectra, which
is consistent with the less edge-on viewing geometry expected at this position.
4. Analysis and Interpretation
4.1. Shock Models
To facilitate interpretation of the ultraviolet spectra, we compare the data with shock
model calculations. We focus on the bright filament covered by the HUT P1 and FUSE
LWRS apertures, and generate a list of line strengths relative to IO III] = 100. The two
apertures cover different regions on the sky, and the O VI surface brightness is lower in the
FUSE LWRS spectrum. We use the O VI to O III] line ratio measured in the HUT P1
spectrum. The line strengths of other high ionization lines below 1200A, S VI, Ne VI] and
Ne V], are scaled relative to IO VI based on the FUSE LWRS measurement, and then rescaled
relative to IO III] based on the HUT P1 O VI to O III] line ratio. The FUSE LWRS surface
brightnesses are used for all other lines in the FUSE bandpass. These data are shown in
column 3 of Table 5.
Shock models were calculated using an updated version of the code presented by Ray-
mond (1979). The updates are discussed by Raymond et al. (1997), and some atomic rates
for Si III, Si IV and S IV lines have been updated based on the CHIANTI database (Dere et
-- 11 --
al. 2001). The model follows the emission and cooling behind a steady shock moving into
a constant density medium. The main inputs are the shock velocity, the pre-shock density
and the elemental abundances. Models are presented for a range of shock velocities.
In
each case the pre-shock hydrogen number density, n0 is 1 cm−3. (The ultraviolet line fluxes
scale linearly with pre-shock density so model line ratios do not depend on n0.) Elemental
abundances, H : He : C : N : O : Ne : Mg : Si : S : Ar : Ca : Fe : Ni = 12.00 : 11.00
: 8.55 : 7.97 : 8.79 : 8.07 : 7.58 : 7.55 : 7.21 : 6.60 : 6.36 : 7.51 : 6.25 on a logarithmic
scale, are used in the models. Except for the Oxygen abundance (discussed below) these are
solar abundances based on Grevesse & Anders
(1989) and tabulated by Ferland (1997).
In all the models the pre-shock magnetic field is 1µG, the temperature of the pre-shock gas
10,000 K, and the gas is fully ionized. These parameters do not affect the ultraviolet line
strengths significantly. For each model, the calculation is followed until the post-shock gas
reaches about 1000 K, by which point the recombination zone is complete.
The ratio of N III] λ1750 to O III] λ1664 is relatively insensitive to the shock velocity.
For the range of shock velocities between about 80 km s−1 and 120 km s−1, the ratio is
a measure of the relative abundances of Nitrogen and Oxygen. There is some amount of
uncertainty in the value for solar oxygen abundance. Ferland (1997), from which we have
taken the other abundances, lists [O] = 8.87. Recently, it has been suggested that the
actual value is lower than normally assumed: Holweger (2001) gives [O] = 8.73, and Allende
Prieto, Lambert, & Asplund (2001) give [O] = 8.69. We ran a set of 100 km s−1 shock models
varying the oxygen abundance between 8.69 and 8.87 (and keeping other abundances fixed).
The observed N III] to O III] line ratio is obtained for Oxygen abundance, [O] = 8.79, and
we used this value in the models presented in columns 4 -- 9 of Table 5.
The optical depths of the strongest resonance lines observed are of order unity in the
direction normal to the shock front. When viewed close to edge-on (as expected at and near
P1), the optical depths are higher. A significant fraction of the line photons emitted by
the shocked gas are scattered out of the line of sight, reducing the observed fluxes. While
comparing spectra with shock model predictions, it is necessary to take into consideration
this decrease in fluxes due to resonance line scattering.
In the case of the O VI lines, the ratio between the intensities of the two lines of the
doublet is a measure of the line optical depth. If we assume that the optical depth along
the line of sight is much higher than the transverse optical depth then, as described by
Long et al. (1992), we can calculate the intensity correction factor. For an observed ratio
I1038/I1032 ≃ 0.76, the intensity correction factor is ∼ 2.14 (for the sum of both lines). The
carbon lines, C III λ977, C II λ1335 and C IV λ1549 are all significantly affected by resonance
scattering. The flux ratio between C III λ977 and C III λ1176 provides an estimate of the
-- 12 --
effects of resonance scattering on the 977A line. The models predict a ratio of about 80, while
the observed ratio is about 20, indicating a correction factor of about 4. This conclusion
is somewhat compromised because of uncertainties in the model predictions of the 1176A
line. We expect that N V λ1240 flux is less affected by resonance scattering than any of
these other lines because nitrogen is less abundant. Specifically, we may assume that the
correction factor for the 1240A line (actually an unresolved doublet) is not more than the
correction factor for the O VI doublet derived above.
4.2. Shock Properties
The ultraviolet spectrum provides several diagnostic line ratios that can be used to
estimate the shock velocities. The ratios N IV] λ1490 to N III] λ1750 and O IV] λ1403
to O III] λ1604 are particularly useful: the lines are intercombination lines and so are not
subject to optical depth effects, and the ratios do not depend on abundances. At the HUT P1
position, IN IV]/IN III] ∼ 0.32 and IO IV]/IO III] ∼ 0.44. A shock velocity vs . 100 km s−1 is
required to reproduce these observed values. It is clear from Table 5 that such a low velocity
shock cannot produce the observed flux of higher ionization lines such as N V λ1240 and
the O VI doublet. If we assume that both N V and O VI are equally affected by resonance
scattering then IO VI/IN V ∼ 7.5. To produce this ratio requires a shock velocity of about
170 km s−1. If N V is less affected by resonance scattering, then the ratio is higher and
so is the required shock velocity. The ratio between Ne VI] λ1006 and Ne V] λ1146 is less
accurate because the lines are weak, but these lines are not affected by resonance scattering,
and the observed ratio implies shock velocities in excess of 180 km s−1. Finally, we note that
the ratio of optical lines I[O III]/I[S II] ∼ 3.1 (Table 1), which requires shock velocities of at
least 80 km s−1. (The forbidden line strengths are predicted by shock models but are not
presented in Table 5.)
From these comparisons we find that the lower ionization lines, including O III], O IV],
N III] and N IV], come predominantly from shocks with velocities . 100 km s−1 and the
higher ionization lines such as O VI, N V, S VI, Ne V] and Ne VI] come from faster shocks,
with velocities about 180 km s−1. The latter velocity falls within the range given by Blair,
Vancura & Long (1995) for O VI producing shocks, but our value for the velocity of the
slower shock is less than their estimate for the C III producing shocks (see §1). Intermediate
velocity shocks cannot contribute a significant fraction of the emission since that would affect
all the line ratios in the wrong way. For example, if there were a substantial contribution from
say a 140 km s−1 shock, then IN IV]/IN III] would be higher than observed while IO VI/IN V
would be lower than observed. The lack of correlation between the C III and O VI velocity
-- 13 --
distributions in the FUSE LWRS spectrum (Figure 5) provides corroborating evidence that
the two lines arise in different shocks. Since the ultraviolet line ratios are similar in both
HUT spectra (§3.2), the arguments and conclusions presented above for P1 also hold for P2.
The observed O III] λ1664 surface brightness, Iobs = 3.1 × 10−15 erg s−1 cm−2 arcsec−2
(Table 2). To proceed with our analysis, we make the simplifying assumption that a
100 km s−1 shock is responsible for all the O III] emission. A 100 km s−1 shock model
using a pre-shock hydrogen number density nmod = 1 cm−3 predicts an O III] line in-
tensity 4.5 × 10−6/2π erg s−1 cm−2 sr−1 (Table 5). This is equivalent to Imod = 1.7 ×
10−17 erg s−1 cm−2 arcsec−2. The model prediction scales with the pre-shock density. Also
the model predicts the intensity emerging perpendicular to the shock front. Since the shock
front is viewed close to edge-on, the effective area of the shock observed is larger than the
aperture area. The 56′′ × 10′′ HUT aperture is placed parallel to the shock front. If the path
length through the O III] emitting gas is l(′′), then the ratio of the shock area to aperture
area is l/10. Thus, Iobs = n100 × Imod × l/10, where n100 is the pre-shock density for
the 100 km s−1 shock component. Substituting the values for observed and model surface
brightnesses, we obtain: n100 × l ≃ 1800.
Below we find that the dynamic pressure of the shocks observed at the FUSE MDRS
shock ≃ 2.6 × 10−9 dyne cm−2. Assuming that this is the pressure driving the
position, ρ0v2
shocks at P1, for vshock = 100 km s−1 and [He] = 11.0, we obtain n100 ≃ 11 cm−3. This, in
turn, implies that the path length through the emitting gas is about 164′′, which is about
0.2 pc at the distance of Vela. The length of the optical filament in the plane of the sky
is about 430′′, which is significantly higher than the derived path length through the O III]
emitting gas. The difference is probably due to the curvature of the shock front into the
plane of the sky, but could also be because the cloud that the shock is running into is
approximately cylindrical with its long direction oriented in the plane of the sky. As we
showed above, the high ionization lines are produced by faster shocks. For a 180 km s−1
shock, the isobaric condition yields a pre-shock density n100 × (100/180)2 ∼ 3.5 cm−3.
At the FUSE MDRS position, the observed O VI to C III ratios are 0.4 for the blue-
shifted component and 0.9 for the red-shifted component (Table 4). (Note: no correction for
resonance scattering was made in this case.) We ran a set of shock models with velocities
in the range 140 -- 160 km s−1 spaced by 5 km s−1. By interpolating from the predicted line
ratios we found that the observed values for the blue-shifted and red-shifted components are
satisfied for shock velocities of ∼ 151 km s−1 and ∼ 157 km s−1, respectively. We note for
shock velocities from 140 to 160 km s−1 that IO VI increases rapidly while IC III stays fairly
constant (Table 5).
If our line of sight intersects with a shock front once, then the following holds (Raymond
-- 14 --
et al. 1997): I0 = Iobsvobs/vshock, where I0 is the intensity of the shock viewed face-on,
vobs is the observed radial velocity (absolute value) and vshock is the shock velocity. We
apply this to the C III emission. The observed surface brightnesses are 21.8 × 10−16 and
16.8 × 10−16 erg s−1 cm−2 arcsec−2 for the blue- and red-shifted components (Table 4). The
central velocities of the components are ∼ −80 km s−1 and ∼ +80 km s−1. Using these
values in the equation above yields I0(blue) ≈ I0(red) ≃ 1 × 10−15 erg s−1 cm−2 arcsec−2.
The face-on C III intensity predicted by a 150 km s−1 shock model for pre-shock density
1.0 cm−3, Imod(C III) ≃ 2 × 10−16 erg s−1 cm−2 arcsec−2. The line intensity scales with pre-
shock density, so the pre-shock density required to produce the observed C III brightness is
I0/Imod(C III) ∼ 5 cm−3. Using the values for shock velocity and pre-shock density derived
above, we find that the pressure in the X-ray knot driving the shock (P = ρ0v2
shock) is about
2.6 × 10−9 dyne cm−2 (P/kB ∼ 1.8 × 107 cm−3 K).
The dynamical pressure of the shock derived above is about twice the value derived
from the HUT spectrum of a face-on shock near the center of the remnant (Raymond et
al. 1997), and about seven times the value in the X-ray region associated with the face-
It is also about 25 times the characteristic pressure of
on shock (Sankrit et al. 2001).
the X-ray emitting gas derived by Kahn et al.
(1985), corrected for the revised distance
of 250 pc to Vela. The high pressure is consistent with Knot D originating in a bullet
near the center of the remnant in that the bullets have to have ram pressures substantially
larger than the average to punch out through the shell. The enhanced pressure would not be
consistent with the knot being a blister on the surface, such as formed in the model suggested
by Meaburn, Hartquist & Dyson (1988) for expanding SNR shells. It is interesting that
Jenkins & Wallerstein (1995) derived a pressure slightly higher than our value in the region
around the line of sight towards HD 72089, which is also near some arcuate optical filaments.
They also found several high velocity components in the absorption spectrum of the star.
The enhanced pressure suggests that the high velocity emission and optical filaments at the
position observed by Jenkins & Wallerstein (1995) may be related to a bullet similar to
Knot D but traveling towards us.
We have used steady flow shock models to explain the observed emission lines. At the
MDRS location, these lines arise in a shell within the surface of an evolving bow shock.
Hartigan, Raymond & Hartmann (1987) approximate a Herbig-Haro bow shock by a series
of plane parallel, steady flow oblique shocks; an approximation that they find valid if the
cooling time is short compared to dynamical times. Knot D is about 30′ (6.7 × 1018 cm)
beyond the main blast wave. Assuming that it is moving at ∼ 500 km s−1 (Aschenbach,
Egger & Trumper 1995), it has taken about 4000 years for it to get to its current position
since overtaking the blast wave. The cooling time for a 150 km s−1 shock running into 5 cm−3
material is about 300 years. The front of the bow shock is running into a denser part of the
-- 15 --
cloud. This encounter results in a slower shock that gives rise to the optical filaments as also
the low ionization ultraviolet lines.
The use of steady flow models has another limitation: shocks of 180 km s−1 are subject
to thermal instabilities (e.g. Chevalier & Imamura 1982). The structure and spectrum of
such a shock fluctuate in time (Innes, Giddings & Falle 1987; Gaetz, Edgar & Chevalier
1988), but the spectrum averaged over time is not greatly affected (Innes 1992). The spec-
trograph apertures include emission from sufficiently complex filaments that the fluctuations
are probably smoothed out.
5. Concluding Remarks
We have analyzed far-ultraviolet and optical observations of Knot D in the Vela SNR,
and shown that the emission is due to shocks driven into a medium with typical interstellar
cloud densities and normal elemental abundances. We have also found that the dynamical
pressure of the shocks is higher than the characteristic value for the remnant. We have
estimated that the path length through the optical filaments is shorter than their extent
in the plane of the sky implying a curved shock front. The velocity profiles of the lines
revealed by high resolution FUSE MDRS is consistent with emission from a bow shock. The
over-pressure in Knot D supports the idea that it is a bow shock around a bullet or cloud
that originated near the center of the remnant and has punched through the shell. However,
there is no evidence for enhanced abundances. If the driver of the bow shock is a shrapnel
of ejecta (Aschenbach, Egger & Trumper 1995), then it remains undetected. An alternative
candidate for the driver is an accelerated cloud that originated near the center. The existence
of such clouds was suggested by McKee, Cowie & Ostriker
(1978) to explain some of the
high velocity emission observed in SNRs.
Knot D, in spite of its unique properties, may be one of many such features including
the ones observed in the plane of the sky around Vela, as well as others moving radially
towards or away from us. In any case, our study provides an estimate of the properties of
Knot D that can be usefully compared with results derived from X-ray data. The properties
of Knot D can also be compared with those of other regions to find out the extent of their
similarities.
The suggestion that Knot D is associated with RXJ0852.0-4622 (Redman et al. 2002)
has been made mainly on morphological grounds. The connection also depends on the
smaller remnant being at about the same distance as the Vela SNR. The distance and
also the age of RXJ0852.0-4622 are highly uncertain (Duncan & Green 2000) and provide
-- 16 --
very few constraints on the properties of putative blow-out regions. We have found that the
interstellar extinction towards the optical filaments is typical for the Vela SNR. Furthermore
our ultraviolet data do not show any peculiarities (e.g. in the velocity distribution) that
suggest a connection between the knot and the newly discovered remnant. Therefore, we
believe that Knot D is associated with the Vela SNR. It will require additional studies to
prove or disprove conclusively the association between RXJ0852.0-4622 and Knot D.
One major issue that remains unanswered is the gap of about 1′ between the optical
filaments and the edge of the X-ray knot (Figure 1). This is not seen in typical shock-
cloud interactions (see references in §3.1) where the X-ray edge and optical emission are
co-incident.
In those cases, the relationship of the two components is well explained by
a model in which the blast wave hits the cloud and a reverse shock is propagated back
into the remnant (Graham et al. 1995).
In Knot D, the gap must be a consequence of
the dynamics of the bow shock. The gap region could contain gas that is hot but with a
temperature not high enough to make X-rays. The region may then be a source of bright
O VI emission. It could also be a transient region of low pressure as predicted by models
of thermal instability (Innes 1992). A third possibility is that it is a region of cool gas
supported by magnetic pressure. Each of these possibilities has interesting implications for
the structure and evolution of astrophysical bow shocks, but the existing data do not allow
us to distinguish among them. However, it is worth noting that the spatial extent of the gap
may be related to the cooling length of the shocked gas. At the distance of Vela, 1′ is about
2 × 1017 cm, which is roughly the cooling length scale of a 240 km s−1 shock in a medium
with density 3 cm−3. These conditions may have obtained in the course of the evolution of
the bow shock. Detailed models and direct observations are needed for further elucidation.
We thank Paul Plucinsky for useful discussions, and for providing us with the Chandra
image presented in Figure 1. We also thank Matt Redman for giving us access to his [S II]
position-velocity data. H2ools, a software package written by Steve McCandliss was used
in estimating the effects of molecular hydrogen absorption. The referee's comments helped
improve the flow and focus of the paper. This work has been supported by NASA grant
NAG5-10248 and NASA contract NAS5-32985, both to the Johns Hopkins University.
-- 17 --
REFERENCES
Allende Prieto, C., Lambert, D. L., & Asplund, M 2001, ApJ, 556, L63
Aschenbach, B. 1998, Nature, 396, 141
Aschenbach, B., Egger, R., & Trumper, J. 1995, Nature, 373, 587
Blair, W. P. et al. 1991, ApJ, 379, L33
Blair, W. P., Vancura, O, & Long, K. S. 1995, AJ, 110, 312
Cha, A. N, Sembach, K. R., & Danks, A. C. 1999, ApJ, 515, L25
Chevalier, R. A., & Imamura, J. N. 1982, ApJ, 261, 543
Danforth, C. W., Cornett, R. H., Levenson, N. A., Blair, W. P., & Stecher, T. P. 2000, AJ,
119, 2319
Davidsen, A. F. et al. 1992, ApJ, 392, 264
Dere, K. P., Landi, E., Young, P. R., & Del Zanna, G. 2001, ApJS, 134, 331
Dubner, G. M., Green, A. J., Goss, W. M., Bock, D. C.-J., & Giacani, E. 1998, AJ, 116, 813
Duncan, A. R., & Green, D. A. 2000, A&A, 364, 732
Ferland, G. J. 1997, Hazy I, a brief introduction to Cloudy 90, pg. 50
Fitzpatrick, E. L. 1999, PASP, 111, 63
Gaetz, T. J., Edgar, R. J, & Chevalier, R. A. 1988, ApJ, 329, 927
Graham, J. R., Levenson, N. A., Hester, J. J., Raymond, J. C., & Petre, R. 1995, ApJ, 444,
787
Grevesse, N., & Anders, E. 1989, Cosmic Abundances of Matter, AIP Conf. Proc. 183, 1 Ed.
C. J. Waddington (New York : AIP)
Hartigan, P., Raymond, J., & Hartmann, L. 1987, ApJ, 316, 323
Hester, J. J., & Cox, D. P. 1986, ApJ, 300, 675
Holweger, H. 2001, Solar and Galactic Composition, Ed. R. F. Wimmer-Schweingruber
(Melville, NY: AIP), p. 23
-- 18 --
Innes, D. E. 1992, A&A, 256, 660
Innes, D. E., Giddings, J. R., & Falle, S. A. E. G. 1987, MNRAS, 226, 67
Jenkins, E. B., Bowen, D. V., Sembach, K. R., & the FUSE Science Team 2001, astro-ph,
0109363
Jenkins, E. B., Silk, J., & Wallerstein, G. 1976, ApJS, 32, 681
Jenkins, E. B., & Wallerstein, G. 1995, ApJ, 440, 227
Kahn, S. M., Gorenstein, P., Harnden, F. R., & Seward, F. D. 1985, ApJ, 299, 821
Kriss, G. A. 1994, in ADASS III, ASP Conf. Proc. 61, ed. D. R. Crabtree, R. J. Hanisch, &
J. Barnes (San Francisco: ASP), 437
Kruk, J. W., Durrance, S. T., Kriss, G. A., Davidsen, A. F., Blair, W. P., Espey, B. R., &
Finley, D. S. 1995, ApJ, 454, L1
Kruk, J. W., Brown, T. A., Davidsen, A. F., Espey, B. R., Finley, D. S., & Kriss, G. A.
1999, ApJS, 122, 299
Long, K. S., Blair, W. P., Vancura, O., Bowers, C. W., Davidsen, A. F., & Raymond, J. C.
1992, ApJ, 400, 214
McKee, C. F., Cowie, L. L., & Ostriker, J. P. 1978, ApJ, 219, L23
Meaburn, J., Hartquist, T. W., & Dyson, J. E. 1988, MNRAS, 230, 243
Moos, H. W., et al. 2000, ApJ, 538, L1
Moriguchi, Y., Yamaguchi, N., Onishi, T., Mizuno, A., & Fukui, Y. 2001, PASJ, 53, 1025
Morton, D. C. 1991, ApJS, 77, 119
Plucinsky, P. P., Smith, R. K., Edgar, R. J., Gaetz, T. J., Slane, P. O., Blair, W. P.,
Townsley, L. K., & Broos, P. S. 2002, ASP Conf. Ser. 271: Neutron Stars in Supernova
Remnants, 407
Raymond, J. C. 1979, ApJS, 39, 1
Raymond, J. C., Blair, W. P., Long, K. S., Vancura, O., Edgar, R. J., Morse, J., Hartigan,
P., & Sanders, W. T. 1997, ApJ, 482, 881
-- 19 --
Redman, M. P., Meaburn, J., O'Connor, J. A., Holloway, A. J., & Bryce, M. 2000, ApJ, 543,
L153
Redman, M. P., Meaburn, J., Bryce, M., Harman, D. J., & O'Brien, T. J. 2002, MNRAS, in
press
Rodgers, A. W., Campbell, C. T., & Whiteoak, J. B. 1960, MNRAS, 230, 243
Sahnow, D. J., et al. 2000, ApJ, 538, L7
Sankrit, R., Shelton, R. L., Blair, W. P., Sembach, K. R., & Jenkins, E. B. 2001, ApJ, 549,
416
Stone, R. P. S., & Baldwin, J. A. 1983, MNRAS, 204, 347
Wallerstein, G., & Balick, B. 1990, MNRAS, 245, 701
This preprint was prepared with the AAS LATEX macros v5.0.
-- 20 --
Fig. 1. -- (a) Three color image of the southern part of Knot D. Narrowband Hα and [O III]
are shown in red and green, respectively. A Chandra ACIS-1 image (courtesy P. Plucinsky)
is shown in blue. The FUSE LWRS (30′′ × 30′′) aperture lies on the optical filament while
the MDRS (4′′ × 20′′) aperture lies on an X-ray bright region ∼ 3.′5 away. (b) Blow-up of the
Hα and [O III] images showing the bright optical filaments. The HUT aperture positions
and the FUSE LWRS position are shown as white boxes. North is up and East to the left
in both images. The LWRS aperture is 30′′ × 30′′.
Fig. 2. -- Optical spectra of the filaments at positions P1 and P2. Spatial extractions from
a longslit spectrum of regions overlapping the HUT aperture positions were used to obtain
these spectra (see text for details). There are no lines observed between 5100A and 6200A,
so that region is not shown in these plots.
Fig. 3. -- HUT spectra of the filaments at position 1 (top panel) and position 2 (bottom
panel). Note the difference in the range of the y-axis scale between the two plots. The
aperture positions are shown in Figure 1.
Fig. 4. -- FUSE LWRS spectrum of the filament: the top, middle and bottom panels show
data from the SiC 2A, LiF 1A and LiF 2A channels, respectively. The flux range has been
chosen to show the weaker lines, and all data have been binned by 12 pixels (∼ 0.07A) along
the wavelength axis.
Fig. 5. -- Fluxes of strong lines observed in the FUSE LWRS spectrum, plotted against
velocity. Top panel: overlay of the two O VI lines. The excess emission seen on the blue
wing of the 1037A line is identified as C II λ1037 emission (see text for details). Bottom
panel: overlay of O VI λ1032 and C III λ977.
Fig. 6. -- Fluxes of lines observed in the FUSE MDRS spectrum, plotted against velocity.
Top panel: overlay of the two O VI lines. Bottom panel: overlay of O VI λ1032 and
C III λ977.
-- 21 --
Table 1. Optical Surface Brightnesses at HUT P1 and P2
Line ID
λ(A)
SB(P1)
SB(P2) Red. Corr.
Hβ
[O III]
[Fe III]
[O III]
[Fe II]+Fe[III]
[N II]
He I
[O I]
[O I]
[N II]
Hα
[N II]
[S II]
[S II]
4861
4959
4987
5007
5270
5754
5876
6300
6364
6548
6563
6583
6716
6731
9.1
25.9
2.4
80.7
0.9
1.2
0.7
2.6
0.9
8.4
27.5
26.5
19.7
14.6
<1.0
9.2
· · ·
28.7
· · ·
· · ·
· · ·
· · ·
· · ·
0.7
2.7
2.3
2.5
1.5
1.40
1.38
1.38
1.38
1.35
1.30
1.29
1.26
1.26
1.24
1.24
1.24
1.23
1.23
Note. -- Surface brightness units: 10−16 erg s−1 cm−2
arcsec−2. Measured fluxes were divided by the area of over-
lap between the longslit and the HUT aperture positions
(2′′ × 10′′) and corrected for interstellar extinction. The cor-
rection factors are presented in the last column of the table.
-- 22 --
Table 2. Ultraviolet Surface Brightnesses measured by HUT
Line ID
λ(A)
SB(P1)
SB(P2) Red. Corr.
S VI
S VI
C III
N III
O VI
O VI
S IV
S IV
He II+[N II]
[Ne V]
C III
N V
C II
O V
Si IV
O IV]
N IV]+[N IV]
C IV
[Ne IV]
He II
O III]
N III]
933
944
977
991
1032
1038
1064
1074
1085
1146
1176
1240
1335
1371
1393
1403
1485
1549
1602
1640
1664
1750
3.6
2.9
42.4
15.6
22.5
17.0
1.8
2.6
4.5
1.3
2.6
5.2
4.5
<0.7
9.3
13.7
3.9
26.4
0.8
11.4
31.1
12.8
3.1
2.5
19.7
7.8
10.0
6.1
0.7
1.6
1.6
· · ·
1.4
3.4
2.2
<0.6
4.7
9.1
3.0
15.2
1.2
3.7
13.8
5.5
5.38
5.12
4.47
4.25
3.73
3.66
3.42
3.34
3.25
2.88
2.74
2.52
2.30
2.24
2.21
2.20
2.11
2.07
2.04
2.03
2.03
2.02
Note. -- Surface brightness units: 10−16 erg s−1 cm−2
arcsec−2. Measured fluxes were divided by the HUT aper-
ture area (10′′ × 56′′) and corrected for interstellar extinc-
tion. The correction factors are presented in the last column
of the table.
-- 23 --
Table 3. Surface Brightnesses in the FUSE LWRS Spectrum
Line ID λ(A)
SB
S VI
S VI
C III
N III
Ne VI
S III
S III
O VI
O VI
S IV
S IV
S III
Si III
Si III
Si IV
Si IV
Ne V
Ne V
C III
933
944
977
991
1006
1015
1021
1032
1038
1063
1074
1077
1110
1113
1123
1128
1137
1146
1176
1.6
1.0
40.6
11.1
0.8
0.7
0.9
17.6
10.3
2.2
1.9
0.6
0.3
0.4
0.3
0.6
0.2
0.5
1.8
Note.
-- Surface
10−16
Brightness units:
erg s−1 cm−2 arcsec−2.
Measured fluxes were di-
vided by the LWRS aper-
ture area (30′′ × 30′′) and
corrected for interstellar
extinction. The correc-
tion factors are presented
in Table 2.
-- 24 --
Table 4. Surface Brightnesses in the FUSE MDRS Spectrum
Line ID λ(A)
SB (blue)
SB (red)
C III
O VI
O VI
977
1032
1038
21.8
5.1
2.8
16.8
9.8
5.5
Note. -- Surface brightness units: 10−16
erg s−1 cm−2 arcsec−2. Measured fluxes
were divided by the MDRS aperture area
(4′′ × 20′′) and corrected for interstellar ex-
tinction. The correction factors are pre-
sented in Table 2.
-- 25 --
Table 5. UV Spectra Compared with Shock Models
Line ID λ(A)
I (P1) M80 M100 M120 M140 M160 M180
S VI
C III
N III
Ne VI]
S III
O VI
S IV
S III
Si III
Si IV
Ne V]a
C III
N V
C II
O V
Si IV
O IV]
N IV]b
C IV
He II
O III]
N III]
937
977
991
1006
1015
1034
1070
1077
1112
1125
1146
1176
1240
1335
1371
1396
1403
1490
1549
1640
1664
1750
12
131
36
4
5
127
13
2
2
3
3
6
17
14
<2
30
44
13
85
37
100
41
0
1073
86
0
5
0
5
2
14
3
0
12
0
263
0
129
9
6
83
4
100
42
0
1154
88
0
4
0
7
2
10
4
0
14
1
243
0
113
49
20
398
19
100
41
2
1070
93
0
5
0
9
2
4
2
1
13
14
210
1
47
85
30
711
44
100
42
11
528
101
0
4
12
12
2
3
1
4
7
81
96
4
23
131
46
638
38
100
39
29
351
67
2
3
447
7
1
3
1
20
4
229
80
20
22
183
34
317
33
100
27
30
418
67
15
3
2221
8
1
3
1
45
5
184
89
32
26
191
30
360
36
100
27
I(O III])c 1664
· · ·
0.33
0.45
0.61
1.21
1.53
1.56
aSum of 1137A and 1146A lines.
bIncludes decays from both 3P1 and 3P2 states; the latter is a forbidden
transition.
cFlux emerging from the shock front 10−5 erg s−1 cm−2, emitted into 2π
steradians.
P1
P2
(a)
(b)
P1
P1
P2
P2
P1
P1
P2
P2
)
1
-
Å
2
-
m
c
s
1
-
g
r
e
3
1
-
0
1
(
λ
F
S VI
S VI
C III
N III
1.2
0.9
0.6
0.3
0.0
920
940
960
980
1000
1.8
1.2
0.6
0.0
1000
0.6
0.4
0.2
0.0
1100
S IV
O VI
S IV
S III
S III
Ne VI]
1020
1040
1060
1080
C III
Si IV
Ne V]
Si III
Si IV
Ne V]
1120
1140
1160
1180
Wavelength (Å)
10
8
6
4
2
3
1
0
1
x
λ
F
0
-400
14
12
10
8
6
4
3
1
0
1
x
λ
F
2
0
-400
OVI 1032
OVI 1038
LWRS
C II 1037
-200
0
Velocity (km/s)
200
400
OVI 1032
CIII 977
LWRS
-200
0
Velocity (km/s)
200
400
3
1
0
1
x
λ
F
3
1
0
1
x
λ
F
OVI 1032
OVI 1038
MDRS
1.5
1.0
0.5
0.0
-400
-200
0
Velocity (km/s)
200
400
OVI 1032
CIII 977
MDRS
3.0
2.5
2.0
1.5
1.0
0.5
0.0
-400
-200
0
Velocity (km/s)
200
400
|
astro-ph/0307197 | 1 | 0307 | 2003-07-10T14:43:21 | Observational constraints for Lithium depletion before the RGB | [
"astro-ph"
] | Precise Li abundances are determined for 54 giant stars mostly evolving across the Hertzsprung gap. We combine these data with rotational velocity and with information related to the deepening of the convective zone of the stars to analyse their link to Li dilution in the referred spectral region. A sudden decline in Li abundance paralleling the one already established in rotation is quite clear. Following similar results for other stellar luminosity classes and spectral regions, there is no linear relation between Li abundance and rotation, in spite of the fact that most of the fast rotators present high Li content. The effects of convection in driving the Li dilution is also quite clear. Stars with high Li content are mostly those with an undeveloped convective zone, whereas stars with a developed convective zone present clear sign of Li dilution. | astro-ph | astro-ph |
Astronomy & Astrophysics manuscript no. 3565delaverny
(DOI: will be inserted by hand later)
November 2, 2018
Observational constraints for Lithium depletion before the RGB.⋆
P. de Laverny1, J. D. do Nascimento Jr2, A. L`ebre3 and J. R. De Medeiros2
1 Observatoire de la Cote d'Azur, D´epartement Fresnel, UMR 6528 CNRS, BP 4229, 06304 Nice, France
2 Departamento de F´ısica, Universidade Federal do Rio Grande do Norte, 59072-970 Natal, RN., Brazil
3 Groupe d'Astrophysique, UMR 5024/CNRS, U. de Montpellier, Place Bataillon, 34095 Montpellier, France
Received 4 February 2003/ Accepted 18 April 2003
Abstract. Precise Li abundances are determined for 54 giant stars mostly evolving across the Hertzsprung gap.
We combine these data with rotational velocity and with information related to the deepening of the convective
zone of the stars to analyse their link to Li dilution in the referred spectral region. A sudden decline in Li
abundance paralleling the one already established in rotation is quite clear. Following similar results for other
stellar luminosity classes and spectral regions, there is no linear relation between Li abundance and rotation, in
spite of the fact that most of the fast rotators present high Li content. The effects of convection in driving the
Li dilution is also quite clear. Stars with high Li content are mostly those with an undeveloped convective zone,
whereas stars with a developed convective zone present clear sign of Li dilution.
Key words. stars: abundances -- stars: evolution -- stars: interiors -- stars: late-type -- stars: rotation
1. Introduction
Despite the important advances made in the past decade
in the study of the stellar lithium behavior, a large num-
ber of questions are not yet answered. We do not com-
pletely understand the processes controlling lithium pro-
duction, nor how and when lithium is depleted. Related to
these questions are the physical bases of the mixing mech-
anisms in stellar interiors. Following the initial study by
Bonsack (1959) different works have attempted to estab-
lish the behavior of lithium abundance along the giant
branch (e.g.: Alschuler 1975; Wallerstein 1966; Brown et
al. 1989) and its link to rotation (Wallerstein et al. 1994;
De Medeiros et al. 2000). These works have shown a steady
decline in lithium abundances from spectral types F5III to
F8III, a wide spread around G0III and a gradual decline
with temperature for stars redward of such spectral type.
Wallerstein et al. (1994) have found that giant stars with
v sin i > 50 km s−1 , located in the (B−V) color interval
from 0.40 to 0.70, present lithium abundance close to the
presumed primordial value, whereas slower rotators show
reduced lithium abundances in spite of their earlier spec-
tral types. De Medeiros et al. (2000) have found a trend of
discontinuity in the distribution of the lithium abundances
around the spectral type G0III, paralleling the sudden de-
offprint
Send
[email protected]
requests
to:
J. R. De Medeiros:
⋆ Based on observations collected at ESO, La Silla, Chile,
and at the Observatoire de Haute Provence, France, operated
by the Centre National de la Recherche Scientifique (CNRS)
cline observed in rotational velocity. The origin of such a
discontinuity is not yet well established but it seems to
strongly depend on stellar mass (De Medeiros and Mayor
1990). Nevertheless, the rotation-lithium connection in gi-
ant stars appears to be a more complicated problem. In
spite of the fact that high lithium content is associated
with fast rotation, slow rotators present a large spread in
the values of lithium abundance (De Medeiros et al., 2000).
In addition, this connection seems to depend on the stel-
lar mass, metallicity and age. Another important question
concerns the level of dilution of lithium along the giant
branch. While standard theory predicts a factor of dilution
of about 40 to 60 for 1 M⊙ and 2 M⊙ respectively (Iben,
1965a, 1965b), the observations for these stellar masses
show that the factor of dilution of lithium is far higher,
reaching values as large as 400 to 1000. A more solid dis-
cussion about this factor of dilution requires more mea-
surements of lithium abundance for stars located near and
at the Hertzsprung gap, in particular for stars blueward
of G0III, namely stars at the blue side of the gap. The dis-
tribution of lithium abundances before stars evolve along
the giant branch may help us to understand the nature
of the apparent discontinuity in lithium content around
G0III and to establish on a more solid basis the dilution
factor of stars evolving along the giant branch.
In this work we present new measurements of lithium
abundance for giant stars in the spectral range F5III
to G5III, typically stars located near and along the
Hertzsprung gap. By combining these data with precise
rotational velocities we analyse the rotation-lithium con-
2
de Laverny, do Nascimento, L`ebre, De Medeiros
nection and the nature of the discontinuities in the distri-
bution of these two stellar parameters.
of observation. Flat-field corrections and wavelength cali-
brations were performed using the MIDAS package.
2. Observations
For this study we have selected 54 luminosity class III
giants with spectral types ranging from F5III to G5III.
All these stars have been previously observed with the
CORAVEL spectrograph (Baranne et al., 1979) for rota-
tional velocity measurements and binarity signatures (De
Medeiros and Mayor, 1999).
2.1. Rotational velocity
For most of the stars of the present sample, rotational
velocity measurements were taken from The Catalog of
Rotational and Radial Velocities for Evolved Stars by De
Medeiros and Mayor (1999). In this work, the authors
present precise rotational velocity, v sin i
, for evolved
stars, obtained on the basis of observations acquired
with the CORAVEL spectrometers (Baranne et al. 1979).
For giant stars of luminosity class III in particular, the
measured rotational velocities have an uncertainty of
about 1.0 km s−1 for stars with v sin i lower than about
30.0 km s−1 . For faster rotators, the estimations indicate
an uncertainty of about 10% on the measurements of v sin i
. For a few stars of the sample, most of them presenting
evidence of high rotation, the v sin i value was determined
on the basis of the spectral synthesis carried out through
this work.
2.2. Spectroscopic observations
The spectral region around the lithium line at 6707.81 A
was observed with two different telescopes. For northern
stars, high-resolution spectra of the lithium region were
acquired with the AURELIE spectrograph (Gillet at al.,
1994) mounted at the 1.52 m telescope of the Observatoire
de Haute Provence (France). The spectrograph used a
cooled 2048-photodiode detector forming a 13 µm pixel
linear array. A grating with 1800 lines/mm was used, giv-
ing a mean dispersion of 4.7 A/mm and a resolving power
around 45 000 (at 6707 A). For this instrumentation and
the selected set-up, the spectral coverage was about 120 A.
The signal to noise ratio was always better than 50. For
southern stars, high-resolution spectra were acquired with
the Coud´e Echelle Spectrometer (CES) in the long camera
mode (Kapper & Pasquini, 1996), mounted at the 1.44 m
CAT telescope, at La Silla, ESO. A RCA high resolution
CCD with 640 x 1024 pixels was used as detector, with
a pixel size of 15 x 15 µm. The dispersion was around
1.9 A/mm and the resolving power was about 95 000 (at
6707 A). The spectral coverage for this instrument was
about 70 A and the signal to noise ratio was always better
than 80. For both observing runs, thorium lamps were ob-
served before and after each stellar observation for wave-
length calibration, whereas 3 series of flat-field using an in-
ternal lamp of tungstene were obtained during each night
3. Lithium abundances and spectrum synthesis
In the present study, we adopted the spectral analysis
method used by L`ebre at al. (1999) and Jasniewicz et al.
(1999), in order to derive the lithium abundances from the
resonance LiI line (6707.81 A). We refer to these authors
for a description of our abundance analysis assuming LTE.
Nevertheless a few important points are described below.
To compute the more appropriated synthetic spectra
to fit the observations, a first guess for the stellar param-
eters of our list of giants has been estimated. Thus we
have looked into the literature and found effective tem-
peratures (Teff ), gravities (log g ) and often metallicites
([Fe/H]) for 37 stars (over 54), mainly in Cayrel de Strobel
et al. (2001), in Allende Prieto and Lambert (1999), and
in Alonso et al. (1999). For the rest of our sample (17 stars
without any published stellar parameters determination),
we first estimated Teff from (B-V) colour index (Flower,
1996), and we have set log g = 3.0, a microtuburlence ve-
locity of 2 km s−1 and [Fe/H] = 0.0 which are values com-
monly adopted for Pop. I giant stars.
Synthetic spectra were then computed from these stel-
lar parameters and from the same line list and model at-
mospheres used in L`ebre at al. (1999) and Jasniewicz et
al. (1999). The stellar parameters of all the giants ( mainly
Teff , [Fe/H] and v sin i ) were then corrected, when nec-
essary, in order to improve the fit quality of the several
Fe I and other metallic lines found in the observed spec-
tral range. The final adopted stellar parameters and the
derived Li abundances are presented in Table 1 for all
the stars of the sample, except HD 156015 which was too
cool to accurately derive its effective temperature with the
adopted procedure (however, the absence of Li signature
is clear for this star). As already discussed by L`ebre et
al. (1999), the major source of uncertainty for this abun-
dance analysis is due to errors in the determination of the
Teff . We estimated that effective temperature were derived
with an uncertainty smaller than ± 200 K. This leads to
an error of less than 0.2 dex on the derived metallicities
and lithium abundances. We also checked with the work
of Carlsson et al. (1994) that non-LTE effects can be ne-
glected for the studied stars since these effects are always
much smaller than 0.1 dex.
4. Results and Discussion
4.1. Rotation and ALi in the Hertzsprung gap
The first step for the present analysis was the construc-
tion of the HR diagram to locate the evolutionary stage
of the stars composing the working sample. In fact, such
a procedure is important because in previous studies the
only criterion for giant star classification was the spec-
tral type. For this purpose, we have used HIPPARCOS
(ESA 1997) trigonometric parallax measurements and V
Constraints for Lithium depletion before the RGB
3
vsini < 5
A(Li) < 1
Fig. 1. Distribution of the rotational velocity of the stars in the
HR diagram. Single and binary stars are identified with open
and filled circles respectively. The size of the circles is pro-
portional to the v sin i (in km s−1 ). Luminosities have been
derived from the Hipparcos parallaxes. Evolutionary tracks at
[Fe/H]=0 are shown for stellar masses between 1 and 4 M⊙.
The turnoff and the beginning of the ascent on the red giant
branch are indicated by the dashed and dotted lines respec-
tively.
magnitudes to compute absolute magnitudes and lumi-
nosities. Bolometric corrections BC were determined from
Flower (1996). Evolutionary tracks were computed from
the Toulouse-Geneva code for stellar masses between 1 and
4 M⊙ and solar metallicity (see also do Nascimento et al.
2000 for complementary informations). The HR diagrams
with the referred evolutionary tracks are displayed in Figs.
1 and 2. In addition, these figures show the behaviors of
the lithium abundance and of the rotational velocity v sin i
, respectively. In these diagrams the dashed line indicates
the evolutionary region where the subgiant branch starts,
corresponding to the hydrogen exhaustion in stellar cen-
tral regions, whereas the dotted line represents the begin-
ning of the ascent along the red giant branch. Except for a
few stars located along the subgiant branch, the large ma-
jority of stars in the present sample are effectively giants
evolving prior to the ascent of the red giant branch.
Figure 1 shows the well established rotational dis-
continuity for giants of
luminosity class III (e.g. De
Medeiros and Mayor 1990). As already shown by these au-
thors, giants blueward of the spectral G0III (correspond-
ing to the location of the rotational discontinuity, around
log(Teff ) ≈ 3.75), show a wide range of rotational velocity
from a few km s−1 to about one hundred times the solar
rotation. Giants redward of such spectral type (G0III) are
essentially slow rotators, except for the synchronized bi-
Fig. 2. Distribution of Li abundances in the HR diagram. Same
as Fig. 1 except that the symbol size is proportional to the Li
abundances.
nary systems and a dozen of late -- G and K single giants
for which the origin of their high rotation is not yet fully
understood. At least for stars with turnoff mass greater
than 1.5 M⊙ , the sudden decline in rotation appears sim-
ply to reflect the rapid increase of the moment of inertia
as the star evolves across this region of the HR diagram.
Figure 2 presents the behavior of lithium abundance,
with a sudden decline in ALi
near the same spec-
tral region where the discontinuity in rotation is ob-
served. Nevertheless, it is quite clear from Figs. 1 and 2
that the behavior of rotation and lithium content in the
Hertzsprung gap, as well as the location of both discon-
tinuities, strongly depends on stellar mass. Such a fact
shows that, for giant stars, it is not correct to define the
location of the discontinuities in rotation and in ALi at
the same spectral type, as suggested by previous works
based only on the analysis of the distributions of ALi and
v sin i versus spectral type or (B−V) color index. Single
stars located on the blue side of the Hertzsprung gap show
a large spread in the lithium abundance, with values of
ALi
ranging from less than 0.5 to 3.0 dex. In contrast,
stars located at the red side show essentially low values
of ALi, reflecting the dilution effects along this spectral
region.
Clearly, one observes from Fig. 2 that stars with 2 to
3 M⊙ located redward of G0III (log(Teff ) ≈ 3.75) show a
factor of dilution of at least 600, which is far in excess
from the theoretical predictions. Such a fact points to an
extra-mixing mechanism occuring before the beginning of
the ascent of the red giant branch. The less massive gi-
ants show a similar disagreement between predicted and
observed abundances. Brown et al. (1989) have measured
4
de Laverny, do Nascimento, L`ebre, De Medeiros
show a trend also observed for other luminosity classes, as
well as for other spectral regions of giants. For slow rota-
tors - in the present situation, for stars with v sin i lower
than about 20 km s−1 - there is a large spread of the values
of ALi, with a dispersion of at least 3 magnitudes. Such
a feature indicates that, also in the Hertzsprung gap, ro-
tation is not the unique parameter driving Li abundance.
Let us recall that, in spite of the limited number of binary
systems with high v sin i value, the dependence of lithium
on rotation seems to follow the same trend observed for
the single stars. We have also analyzed the relation metal-
licity versus v sin i effects, looking for possible effects of
metallicity on the spread observed in the ALiversus rota-
tion relation, by using data presented in Table 1. No clear
trend arises from such an analysis.
4.3. The connection lithium - deepening of the
convective envelope
Because the level of dilution of lithium depends on the
level of convection in the stars, we analyse here the behav-
ior of lithium abundance as a function of the deepening
of the convective zone for our working sample. Such an
analysis sounds interesting because most of the stars are
crossing the Hertzsprung gap, where the convective zone
is predicted to reach its most important development. For
our purposes, we have first estimated the mass of each
star M∗ from the HR diagram constructed for Figs. 1 and
2. Then, we have estimated the deepening in mass of the
convective zone MCZ , according to the recipe applied
by do Nascimento et al. (2000) and do Nascimento et al.
(2003). In short, with the mass of each star in hand, we
have placed such mass in the convective zone mass deep-
ening Mcz/M∗ versus Teff diagram from do Nascimento et
al. (2000) and estimated the parameter Mcz/M∗ for each
star. The individual values of M∗ and Mcz/M∗ are listed
in Table 1. The star HD 224342 is outside this analysis be-
cause of its very large and uncertain luminosity and mass.
Figure 4 displays the deepening (in mass) of the convec-
tive envelope versus effective temperature, for single and
binary stars. In this figure, the presence of the binary stars
should be analysed with caution due to the large uncer-
tainties in the estimation of mass for stars in binary sytems
using evolutionary tracks. However it is clear from this fig-
ure that the sudden decline in ALi in the Hertzsprung gap
is directly associated with the rapid increase of the con-
vective envelope in such region. Most of the stars with
high Li content present an undeveloped convective zone,
whereas stars with low Li content show a developed con-
vective zone.
5. Conclusions
The present study brings precise abundances of Li for
giant stars of luminosity class III evolving across the
Hertzsprung gap. By combining these data with rota-
tional velocity and other stellar parameters such as lu-
minosity and effective temperature, we show a trend for
Fig. 3. Lithium abundance as a function of rotational velocity
for giant stars in the Hertzsprung gap. Single and binary stars
are identified by open and filled circles respectively. Vertical
arrows indicate ALi upper limits
lithium abundances for a large sample of late-G to K gi-
ants, with mean mass around 1.5 M⊙ . For such stars, they
found the mean lithium abundance lower than 0.0 dex,
that is at least ∼ 1.5 dex below the value predicted by
standard theory. In Fig. 2 of the present paper, by taking
into consideration the abundance of lithium for stars with
masses between 1.5 M⊙ and 3.0 M⊙ , located at the blue
side of the Hertzsprung gap, and referring to a mean value
of ALi
lower than 0.0 dex as estimated by Brown et al.
(1989), we observe that the factor of dilution for this in-
terval of mass is even more important than that observed
for stars with 1.5 M⊙ .
It is interesting to point out that the binary systems
seem to follow the same tendency presented by single
stars. Redward of the spectral type G0III, the binary sys-
tems show essentially low values of ALi, reflecting also the
effects of dilution along the Hertzsprung gap.
4.2. The connection lithium - rotation in the
Hertzsprung gap
The dependence of lithium content on rotation along the
Hertzsprung gap has been already reported in different
works (Wallerstein et al., 1994; Alschuler 1975). For the
present sample of stars, this trend is also clearly observed
from Fig. 3, which displays lithium abundance as a func-
tion of rotational velocity. Stars with high rotation also
present high values of ALi. In addition, the present data
Constraints for Lithium depletion before the RGB
5
gratefully acknowledge C. Ferrari for her wonderful software
tiramisu.
References
Allende Prieto, C., Lambert, D. L. 1999, A&A 352, 555
Alonso, A., Arribas, S., Martinez -- Roger, C. 1999, A&AS 140,
261
Alschuler, W. R. 1975, ApJ 195, 649
Baranne, A., Mayor, M., Poncet, J. L. 1979, Vistas Astron. 23,
279
Bonsack, W. K. 1959, ApJ 130, 843
Brown, J. A., Sneden, C., Lambert, D. L., Dutchover, E. 1989,
ApJSS 71, 293
Carlsson M., Rutten R.J., Bruls J.H.M.L., Shchukina N.G.
1994, A&A 288, 360
Cayrel de Strobel, G., Soubiran, C., Ralite, N. 2001, A&A 373,
159
De Medeiros, J. R. and Mayor, M. 1990,
in Cool Stars,
Stellar Systems and the Sun, Ed. G. Wallerstein (ASP, San
Francisco), p. 404
De Medeiros, J. R., Mayor, M. 1999, A&AS 139, 433
De Medeiros, J. R., do Nascimento, J.D.Jr., Sankarankutty, S.,
Costa, J. M., Maia, M. R. G., 2000, A&A 363, 239
do Nascimento, J.D.Jr., Charbonnel, C., L`ebre , A., de Laverny,
P., De Medeiros, J. R., 2000, A&A 357, 931
do Nascimento J.D.Jr., Melo C.H.F., Canto Martins B.L., De
Medeiros J.R. 2003, A&A, 405, 723
ESA 1997, The Hipparcos and Tycho Catalogues, ESA SP-1200
Flower P.J. 1996, ApJ 469, 355
Gillet D., Burnage R., Kohler D., Lacroix D., Adrianzyk G.,
Baietto J. C., Berger J. P., Goillandeau M., Guillaume C.,
Joly C., Meunier J. P., Rimbaud G., & Vin A. 1994, A&AS,
108, 181
Iben I. 1965a, ApJ 142, 1447
Iben I. 1965b, ApJ 141, 993
Kaper L., Pasquini L., 1996, CAT + CES Operating Manual.
ESO Report, 3P6CAT-MAN-0633-0001
L`ebre A., de Laverny P., De Medeiros J.R., Charbonnel C., da
Silva L. 1999, A&A 345, 936
Jasniewicz G., Parthasarathy M., de Laverny P., Th´evenin F.
1999, A&A 342, 831
Rutten R.G.M. 1987, A&A 177, 131
Wallerstein G. 1966, ApJ 143, 823
Wallerstein G., Bohm-Vitense E., Vanture A.D., Gonzalez G.
1994, AJ 107, 2211
Fig. 4. The deepening (in mass) of the convective envelope as a
function of the effective temperature for the stars in the present
sample. Single and binary stars are represented by open and
filled circles respectively. The symbol size is proportional to
the Li abundances quoted in the figure.
the existence of a sudden decline in ALi located around
log(Teff ) ≈ 3.75, and corresponding to the spectral type
G0III, where the well established rotational discontinu-
ity for giants is defined. Blueward of this region, most of
the stars with mass greater than about 1.5 M⊙ present a
trend for high Li content, whereas for stars redward of
this region Li is essentially diluted. The scenario of Li
content in single stars seems to be followed by binaries.
The dependence of Li content on rotation, in the sense
that fast rotators show high ALi, is also confirmed by
the present set of data. In addition, the large spread in
ALi for slow rotators observed for other luminosity classes
and other spectral regions occurs also in the Hertzsprung
gap. Metallicity seems to have no effect on such a spread.
Finally, we have combined the ALi values with theoretical
predictions on the convective zone deepening to analyse
the extent of the level of convection on the Li dilution. It
is clear from this analysis that stars with high Li content
have mostly undeveloped convective zones, whereas stars
with developed convective zone present low Li content.
Acknowledgements. JRM warmly acknowledges colleagues of
the Cote d'Azur Observatory at Nice, where a large part of this
work was prepared during his stay in France. This work has
been supported by continuous grants from the CNPq Brazilian
Agency and by funds from the french Programme National
de Physique Stellaire (INSU/CNRS). J.D.N.Jr. acknowledges
the CNPq for the fellowship PROFIX 540461/01-6. We also
6
de Laverny, do Nascimento, L`ebre, De Medeiros
Table 1. The stars of the present working sample with their physical parameters
HD (B-V)
0.901
0.68
0.442
0.697
0.758
0.443
0.397
0.571
0.58
0.856
0.681
0.935
0.685
0.89
0.774
0.781
0.895
0.324
0.401
0.515
0.445
0.681
0.857
0.845
0.54
0.873
0.877
0.921
0.897
0.504
1.164
0.724
0.51
0.567
0.683
0.487
0.839
0.782
0.456
0.777
0.946
0.994
0.549
1.005
0.474
0.786
0.463
0.471
0.771
0.502
0.802
0.617
0.461
0.712
87
895
1671
6903
17878
48737
55052
63208
65448
71369
72779
74485
74874
78715
81025
82210
85945
92787
101133
107700
108722
111812
119458
121107
136202
140438
151627
152863
153751
155646
156015
157358
159026
160365
161239
169985
173920
175492
182900
185758
200039
200253
202447
203574
203842
208110
209149
210459
214558
215648
218658
220657
223346
224342
Teff
5030
5100
6470
5700
5300
6600
6820
5100
5250
5220
5790
4960
5400
5050
5200
5250
5200
7250
6800
5500
6600
5600
5100
5100
6040
5110
5090
4980
5040
6160
<4000
5300
6300
6150
5900
5400
5350
5100
6400
5440
4990
5100
5300
4860
6420
5260
6500
6500
5200
5950
5200
6000
6600
5520
log g
3.0
3.0
3.8
3.0
2.7
3.9
3.5
3.0
3.0
2.6
3.0
2.1
3.0
3.0
3.0
3.4
3.0
4.2
3.4
3.1
3.6
3.0
3.0
3.2
4.0
3.0
3.0
3.0
2.0
3.9
3.0
3.0
3.4
3.8
3.2
3.0
3.0
3.9
3.1
3.0
3.0
3.2
3.0
3.0
3.0
3.8
3.0
3.0
4.0
2.9
3.0
4.1
2.0
[Fe/H]
-0.3
-0.3
-0.1
-0.2
0.1
0.0
0.1
0.0
0.0
0.0
0.0
-0.4
0.0
-1.0
-1.0
-0.3
0.0
-0.3
0.0
-0.2
0.0
-0.3
-0.2
0.0
0.0
0.1
0.0
0.1
-0.3
0.0
0.0
0.0
0.0
0.4
0.1
0.0
0.0
0.0
-0.1
-0.4
0.2
-0.4
-0.4
-0.3
-1.0
0.0
0.0
-0.4
-0.4
-0.2
0.0
0.0
-1.0
v sin i M/M⊙ Mcz/M∗
3.8
2.5
46.5
90.0
2.6
70.0
75.
5.7
2.5
4.3
90.0
6.6
4.0
2.0
5.0
5.5
6.2
45.0
33.5
3.9
100.0
66.5
4.0
14.5
4.8
5.8
4.1
2.9
23.
6.9
15.6
1.0
139.0
100.0
5.9
10.0
8.0
4.2
26.7
7.1
1.0
8.0
1.3
1.0
90.0
3.3
50.0
120.0
1.4
7.9
5.5
80.0
18.5
12.8
3.3
2.8
1.8
2.9
3.8
1.6
2.2
4.0
3.3
3.8
2.7
3.3
3.0
3.2
2.9
2.0
3.4
1.6
2.1
3.0
1.9
3.0
3.5
4.0
1.2
3.5
3.8
3.2
4.0
1.5
4.0
3.1
3.7
2.0
1.4
3.8
4.0
4.0
1.6
4.0
3.1
3.7
3.1
3.0
2.0
3.0
1.5
2.8
3.1
1.2
3.3
2.4
1.5
4.5
0.27
0.12
0.00
0.00
0.02
0.00
0.00
0.02
0.02
0.01
0.00
0.25
0.00
0.25
0.02
0.26
0.02
0.00
0.00
0.00
0.00
0.00
0.05
0.02
0.00
0.05
0.04
0.25
0.05
0.00
0.83
0.05
0.00
0.00
0.00
0.00
0.01
0.01
0.00
0.00
0.25
0.04
0.03
0.35
0.00
0.05
0.00
0.00
0.03
0.00
0.02
0.00
0.00
−−
Rem
SB?
SB
(1)
(1)
SB
(1)
SBO
SBO
SB
(1,2),SB
SBO
(1)
SBO
SBO
SB
SBO
SBO
(1),SB
(1),SB
SBO
(2)
SB?
(1)
(2),SB
(1)
SB
(1)
(1)
SB
SBO
(1)
ALi
<0.5
<0.5
2.8
2.7
<0.5
3.3
<2.0
1.0
1.6
<0.5
3.3
1.0
1.4
<0.5
<0.5
0.9
1.0
2.5
2.5
1.6
3.3
2.7
1.2
<0.5
<1.0
<0.5
<0.5
<0.5
1.2
3.1
<0.5
1.1
<2.0
3.3
<0.5
1.4
<0.5
1.0
2.8
<0.5
<0.5
1.5
<0.5
<0.5
3.1
<0.5
3.0
3.2
<0.5
2.1
<0.5
2.5
<0.5
<0.5
(1) v sin i determined from the spectral synthesis. (2) Uncertain Teff or suspected binary
|
astro-ph/0002057 | 1 | 0002 | 2000-02-02T17:43:31 | The Warp of the Galaxy and the Orientation of the LMC Orbit | [
"astro-ph"
] | After studying the orientation of a warp generated by a companion satellite, we show that the Galactic Warp would be oriented in a different way if the Magellanic Clouds were its cause. We have treated the problem analytically, and complemented it with numerical N-Body simulations. | astro-ph | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 24 October 2018
(MN LATEX style file v1.4)
The Warp of the Galaxy and the Orientation of the LMC
Orbit
I. Garc´ıa-Ruiz1, K. Kuijken1, J. Dubinski2
1Kapteyn Astronomical Institute, Postbus 800, 9700 AV Groningen, The Netherlands
2Department of Astronomy and CITA, University of Toronto, 60 St. George Street, Toronto, Ontario M5S 3H8, Canada
24 October 2018
ABSTRACT
After studying the orientation of a warp generated by a companion satellite, we show
that the Galactic Warp would be oriented in a different way if the Magellanic Clouds
were its cause. We have treated the problem analytically, and complemented it with
numerical N-Body simulations.
Key words:
Clouds
galaxies:kinematics and dynamics -- Galaxy:structure -- Magellanic
1
INTRODUCTION
The disk of the Milky Way is remarkably flat out to 10
kpc, where it starts to bends in opposite directions in the
southern and northern parts. The cause of it is still a puzzle:
for a review, see Binney (1992).
One possibility is that the Magellanic Clouds distort the
Galactic disk in the observed way. The fact that the direction
of the maximum warping lies very close to the galactocentric
longitude of the Clouds makes this hypothesis tempting.
The problem with this scenario is that the Clouds are
not massive enough to generate the warp amplitudes that we
observe at their present distance. This was noticed by the
first researches in this field (Burke 1957; Kerr 1957), and
later by Hunter & Toomre (1969). A remedy which might
allow this scenario to work was to suppose that the Clouds
are currently not at the pericenter of their orbit, so that
they have been much closer to the Galactic disk in a pre-
vious epoch (≃ 20 kpc is what was needed). However, later
work by Murai & Fujimoto (1986) determined the orbit of
the Clouds, and proved that the Clouds are actually at their
pericenter, with an apocenter close to 100 kpc, so the prob-
lem of the small amplitude still remains if the Clouds are to
be blamed as the cause of the Galactic warp.
Recently a mechanism for amplifying the effect of a
satellite has been proposed by Weinberg (1998). He describes
a calculation in which a disk galaxy surrounded by a dark
halo is perturbed by a massive satellite, similar to the LMC.
By means of a linear perturbation analysis, he follows the
perturbation (wake) created by the satellite in the halo, in-
cluding its self-gravity. He finds that the torque exerted on
the disk is several times larger than that due directly to
the satellite: the latter is amplified because (i) the satellite-
induced wake in the halo itself exerts a torque, roughly in
phase with that from the satellite; and (ii) the wake itself
further perturbs the halo, resulting in a torque that is larger
c(cid:13) 0000 RAS
again. Under circumstances in which the satellite orbital fre-
quency is close to the natural precession frequency of the
disk (i.e., the warp mode frequency of Sparke & Casertano
(1988)), a significant amplitude can result. A calculation by
Lynden-Bell (1985) of a similar scenario gives comparable
results, as does a simple model described by Kuijken (1997).
In this paper, we focus on the orientation of a warp
generated by a massive orbiting satellite with less emphasis
on the amplitude of the warp. In §2 we discuss a simple
analytic model in which the disk and halo are rigid: this
establishes the baseline response of a disk to satellite tides,
and its orientation with respect to the satellite orbital plane.
As we show, this orientation is different from that of the
Galactic warp to the LMC orbit. §3 contains a description
of the N-body code used, and §4 to §6 the results of our
N-body simulations, showing that the orientation problem
remains. In §7 we give our conclusions.
2 ANALYTIC RESULTS WITH A SIMPLIFIED
MODEL
A simple model serves as a reference for the warp response
of the disk to an orbiting satellite. Consider a rigid disk,
embedded in a rigid halo potential, and subjected to the
potential of an orbiting satellite. The evolution of the disk is
governed by the combined torque from halo and satellite. A
stellar or gaseous disk is floppy, and so will warp when tilted,
since it is not able to generate the stresses that would be
required to keep it flat; however the overall re-alignment of
the disk angular momentum should be comparable between
the rigid and floppy cases.
The angles used in this paper related to the satellite,
and the definition of our coordinate system are illustrated
in Figure 1. The tilting of the disk is measured by the angle
2 Garc´ıa-Ruiz et al.
SGP
z
sθ
l=180
l=90
LMC
NGP
Figure 1. Definition of the coordinate system, satellite angles,
and orientation of the disk of the Galaxy. The disk lies on the
z = 0 plane. The sun is on the left, and the galactic plane is
viewed from the South Galactic Pole so that the disk rotates
counterclockwise (indicated by an arrow).
between the z axis and the angular momentum of the disk.
The longitude of this vector is the same as the longitude of
the maximum of the warp when looking at the disk from the
North Galactic Pole.
The Lagrangian for a rigidly spinning, axisymmetric ob-
ject is
L = 1
2 I1( θ2 + φ2 sin2 θ) + 1
2 I3( φ cos θ + ψ)2 − V (θ, φ)
(1)
where (θ, φ, ψ) are the Euler angles, and I3 and I1 are the
moments of inertia of the object about its symmetry axis and
about orthogonal directions. V is the potential energy of the
body in the halo plus satellite potential. The ψ-equation of
motion leads to the conserved quantity S = I3( φ cos θ + ψ),
the spin, and the other two equations of motion then become
I1 θ − I1 φ2 sin θ cos θ + S φ sin θ +
∂V
∂θ
= 0
and
I1
d
dt
( φ sin2 θ) +
∂V
∂φ
= 0.
(2)
(3)
For small deviations from the equator (θ = 0), we can ex-
pand these equations in terms of x = sin θ cos φ ≃ θ cos φ,
y = sin θ sin φ ≃ θ sin φ. In these terms the equations of
motion become
∂V
∂x
∂V
∂y
I1 x + S y +
I1 y − S x +
= 0,
(4)
(5)
= 0.
2 VH(x2 + y2), and that
flattened halo will have the form 1
due to the satellite at position θS, φS will be −VS(sin2 θS −
x sin 2θS cos φS − y sin 2θS sin φS) where VH and VS are con-
stants representing the strengths of the halo torque and of
the quadrupole of the tidal field from the satellite, respec-
tively. Hence we find
I1 x + S y + VHx + VS sin 2θS cos φS = 0,
I1 y − S x + VHy + VS sin 2θS sin φS = 0.
(6)
(7)
l=270
y
l=0
x
If furthermore the satellite orbit is circular and polar in the
x = 0 plane, θS = ΩSt, φS = 90, and the solution to the
equations of motion is
x =
2ΩSS
∆
VS cos 2ΩSt;
y =
4I1Ω2
S − VH
∆
VS sin 2ΩSt
(8)
S)2 − 4Ω2
plus free precession and nutation terms, where ∆ = (VH −
4I1Ω2
SS 2. (A more general quasiperiodic satellite
orbit yields a solution which can be written as a sum of
such terms.) Notice that the satellite provokes an elliptical
precession about the halo symmetry axis, with axis ratio
dependent on the halo flattening and on the satellite orbit
frequency. For example, for an exponential disk of mass M ,
scale length h and with a flat rotation curve of amplitude
v, I3 = 2I1 = 6M h2 and S = 2hvM . For such a disk in
a spherical (or absent) halo (VH = 0), a satellite orbiting
at radius rS has frequency ΩS = v/r, and hence the axis
ratio of the forced precession is (x : y) = rS/3h. Hence the
response of the disk to a distant satellite is mainly to nod
perpendicular to the satellite orbit plane. This result can be
understood as the classic orthogonal response of a gyroscope
to an external torque: a distant satellite has a sufficiently low
orbital frequency that the disk responds as if the torque were
static.
For a slightly flattened potential of the form 1
2 v2 ln[R2 +
(z/(1 − ǫ))2], VH = M v2ǫ. With non-zero ǫ, the axis ratio of
the precession cone becomes [(4h/rS)/(ǫ − 12h2/r2
S)]: again
the oscillation in x is larger than that in y except for very
flattened halos.
The amplitudes generated by tidal perturbation from
a satellite such as the LMC are small, less than a degree.
The largest amplitude of oscillation is in the y-direction.
The potential energy of the disk due to the tidal field of the
satellite can be shown to be (see Appendix)
VS =
3GMSI1
2r3
S
.
(9)
Hence equation 8 yields, to leading order in h/rS, an x-
amplitude of
9
8
GMS
v2rS
h
rS
≃ 0.09◦
(10)
for the LMC (orbital radius about 50 kpc, and rS/h ≃ 15).
This number increases only slightly (a factor 2) for halo
flattenings up to 0.2 (see Figure 2).
It is clear from this calculation that simple tidal tilting
of a disk by an LMC-like satellite does not provide a good
model for the warping in the Galaxy, because the orientation
of the warp is not perpendicular to the orbital plane of the
LMC. This constraint is independent of the strength of the
perturbation VS.
For small x, y, the potential energy of the disk due to the
The amplitudes are much too small, but we have only
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The Warp of the Galaxy and the Orientation of the LMC Orbit
3
Figure 3. Rotation curve showing the contribution from the disk
(dotted) and halo (dashed) to the total (solid)
more rings and more particles per ring, without important
changes in the results described below.
The satellite is modelled as having a Plummer distribu-
tion. To avoid relative movements of the galaxy with respect
to the satellite we have used two satellites instead of one,
symmetrically placed with respect to the centre of the halo-
disk system. This causes the dipole term of the tidal field
to be zero, avoiding relative movements of the galaxy with
respect to the satellites. It is equivalent to only keeping the
even-l terms in the potential of the satellite, neglecting the
dipole, l = 1, terms in the potential, and concentrating on
the warp (which result from the quadrupole, l = 2 terms).
The first run was made with a satellite in a circular
orbit, to try to reproduce the predictions in §2. Later a non-
circular orbit is considered, and the difference between both
simulations analysed. The non-circular orbit has a pericen-
ter at 50 kpc and apocenter at 100 kpc, consistent with re-
cent determination of the orbit of the Clouds (Lin, Jones &
Klemola 1995). In the non-circular simulations the satellite
starts at its apocenter at the beginning of the simulation,
where the perturbation on the disk is the smallest possible.
The units of the model translate to the Galaxy (disk
scale-length 4.5 kpc, and the rotation velocity at 8.5 kpc of
220 km s−1) as follows: length unit = 4.5 kpc, velocity unit
= 340 km s−1, time unit = 1.30 × 107 years, mass unit=
1.20 × 1011M⊙. With these numbers, the disk mass of our
model is 6.1 × 1010 M⊙, and the satellite (LMC) has a mass
of 1.5 × 1010 M⊙, the biggest current mass estimate for the
Clouds (Schommer et al. 1992). In the coordinate system of
the simulations, the z = 0 is the disk plane, and the orbit of
the satellite lies in the x = 0 plane.
3.2 Code used to evolve the system
The disk is modelled as a system of pivoted spinning rings,
fixed at the centre of the halo. Each ring is realized as 36
azimuthally-spaced particles, and the potential generated by
the rings is calculated with a tree code (Barnes & Hut 1986).
The forces on the individual "ring-particles" are used to cal-
culate the torque on each ring. The force exerted by a satel-
lite on the ring particles was evaluated directly using the
Plummer law.
The Euler equations for rings and axisymmetric bodies
can be rewritten in a form so that the time derivatives of
the instantaneous angular velocities about the body axes are
linear combinations of the angular velocities, torques and
body normal vector components. This allows the derivation
Figure 2. The oscillation of the axis of a rigid exponential disk
subjected to the tidal field of an orbiting satellite. The amplitude
is calculated assuming a satellite of mass 1.5 × 1010M⊙, orbiting
at radius 54kpc in the z = 0 plane. The direction of the tilt of
the Galactic disk with respect to the Magellanic Clouds' orbital
plane is indicated by the arrow. The dots mark the expected
position of the disk axis given the current phase of the LMC
orbit for (bottom to top) halo potential ellipticities ǫ = 0 (solid
symbol),0.05, 0.1, 0.15 and 0.2 (open circles).
considered the tilting of a rigid disk, and the situation can
change when the floppiness of the disk is considered.
3 SIMULATION DETAILS
To test this scenario, and in particular to get beyond the
rigid tilting considered above, we have performed some N-
body simulations. We assume the halo to be a background
potential which does not respond to the disk or the satellite.
The N-body code used for this work models the disk as
a set of concentric spinning rings embedded in a spherical,
rigid halo. This description allows warps to be described,
but not in-plane distortions of the disk such as bars or lop-
sidedness.
3.1 Initial conditions
We have performed simulations with two types of disks: a
rigid disk and a exponential disk. The rigid disk run tells us
how good the analytic predictions are, and the exponential
disk is used later for a more realistic approach.
We use a King model for the halo, in order to obtain
a reasonable flat rotation curve (Figure 3). This is accom-
2 = −6, a tidal radius of 44
plished with a model of Ψ0/σ0
(200 kpc. for a 4.5 kpc scale-length disk), and mass of 10
disk masses.
Each of the disks is made of 1000 rings, each of them
consisting of 36 particles. Various runs where made with
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4 Garc´ıa-Ruiz et al.
Figure 4. Precession path followed by the rigid disk: x = θ cos φ,
y = θ sin φ. The dots indicate the disk's state when the satellite
has the LMC's orbital phase.
of a second order explicit time-centred leapfrog integration
scheme that can be used to solve the coupled equations for
the rings and make it easy to merge with an N-body code
(Dubinski 2000).
4 RIGID DISK
As a first approach, we have evolved a rigid disk and anal-
ysed its evolution under the influence of an orbiting satellite.
The result of our simulation is in good agreement with the
analytic predictions. The disk wobbles under the influence
of the satellite, describing an ellipse elongated in the direc-
tion perpendicular to the satellite's orbital plane. The path
followed by the disk is plotted in Figure 4, where it can be
seen that most of the time the maximum of the warp is lo-
cated in the direction perpendicular to the orbital plane of
the satellite (l = 0◦ and l = 180◦). The ellipse isn't as regu-
lar as in Figure 2 for two reasons: the assumption that the
disk is much smaller than the orbital radius of the satellite is
not completely fulfilled; and there are some transient terms
present because of the initial conditions of the simulation.
This are also the cause for the precession ellipse of the disk
not to be centred in the origin.
The position of the warp when the satellite is at the
location of the LMC is indicated by the dots in Figure 4, and
the location of them resembles the predicted one in Figure 2
(for ǫhalo = 0) remarkably well.
5 EXPONENTIAL SELF-GRAVITATING DISK
We now consider a more realistic disk: an exponential disk
model, in which we have considered also the disk's self-
gravity. The first thing that draws our attention in this
simulation is a peak we see in the inclination at around
6.5 scale-lengths. Simulations done with a different rotation
curve showed that this peak occurs at the locations on the
disk that satisfy Ωs/wz = 2, 3, ..., that are caused by reso-
nances with the satellite's orbital frequency. This happens
because the non-linear behaviour of the outer parts of the
disk, where the assumption rs ≫ rdisk is worse than it is
further in.
This is not the kind of warp we are looking for, due
to the fact that it is the result of a satellite with a single
frequency, and in the real case the eccentric orbit of the
satellite will wash out this peak. Looking at the evolution
of the disk it is clear that the warp looses its coherence at
a radius about 4.5 scale-lengths (at larger radius the line
of nodes winds up), so we will measure the warp properties
considering that the disk finishes there.
In the case of a floppy disk it is not straight-forward
to define a single inclination and position angle. We have
separated the disk into two components: the inner disk and
the outer (warped) disk. The inner disk consists on the first
2 scale-lengths, and remains practically flat along the simu-
lation. The warping angle is then calculated as the angle
between the inner and outer disk vectors. We have chosen
to use the disk vectors and not the angular momentum, for
example, not to penalise the outer less massive rings. The
results presented here do not change significantly when the
definition of the inner disk is altered.
It has to be kept in mind that the warping angles quoted
here are different than the maximum amplitude of the warp,
who usually are larger by a factor not greater than 5.
Using this method we obtain a plot similar to Figure 4
for the exponential disk, which is shown in Figure 5. Only
the path after t=160 is shown, that is the moment when the
disk behaviour reaches an equilibrium.
Note that the predictions for the Galactic Warp don't
really change with the floppiness of the disk: it is clearly close
to l = 0◦, as chapter §2 predicted, and not at l =≃ 90◦, as
we observe it in the Galaxy.
6 NON-CIRCULAR ORBIT, AND FLATTENED
HALOS
We also considered non-circular orbits, to allow for the fact
that the orbit derived for the Clouds has a pericenter of 50
kpc and an apocenter of 100 kpc (Murai & Fujimoto 1980).
The changing radius of the satellite causes a fluctuating tidal
field amplitude, which could be important for the dynamics
of the disk. Here we show that in fact the effect does not
change our conclusions materially.
First, to have an idea of what to expect, we integrated
the analytic equations of section §2 with a satellite in this
kind of orbit. The result was, as before, that the disk's
precession path was contained within an ellipse, elongated
along the direction perpendicular to the satellite's plane.
This causes the warp maxima to be most of the time close
to the direction perpendicular to the satellite's orbit.
We then performed simulations with this type of or-
bits. The first thing we observe in these simulations is that
the resonance peak we found in the circular orbit simulation
has disappeared. Now the satellite doesn't have a single fre-
quency, so the result is not surprising. The energy of the
resonance gets distributed along different parts of the disk
now, and no coherent pattern can be maintained across the
disk, winding up the outer parts of the disk. When we look at
the inner 4.5 scale-lengths as before, the precession pattern
remains similar to the simulation with the circular orbit, so
does the prediction of the warp's longitude at LMC's actual
orbital phase. So our conclusions are not modified by the
non-circularity of the orbit.
The halos considered in all these simulations are spher-
ical, which means that they don't contribute to the gener-
ation of torques on the disk. We know that halos are not
spherical, which creates a preferred plane in which the disk
settles. Ellipticities of the order of 0.05 in the potential make
the precession paths described before yet more elongated,
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The Warp of the Galaxy and the Orientation of the LMC Orbit
5
Figure 5. Warp orientation followed by the Exponential Disk: x = θ cos φ, y = θ sin φ. The dots indicate the disk's state when the
satellite has the LMC's orbital phase.
which would make the chances of finding the warp maxima
in the satellites' direction even more unlikely.
V = −GMSZ Σdx dy (r2
S − 2yrS sin θS + x2 + y2)−1/2.(A2)
7 CONCLUSION
We show by means of analytic work and N-body simulations
that the precession path of a warp generate by an orbiting
satellite galaxy is elongated along the direction perpendic-
ular to the satellite's orbital plane. Applying our result to
the Milky Way, if the Galactic Warp is generated by the
Magellanic Clouds, the direction of maximum amplitude of
the warp would lie close to l ≃ 0◦, as compared to the ob-
served direction of ≃ 90◦. Even if the halo's self-gravitating
tidal response to the satellite amplifies the effect of the satel-
lite (Weinberg 1998), this response will be mostly in phase
with the satellite, and the alignment problems will persist.
Possibly the Sagittarius dwarf galaxy, whose orbit lies at
90◦ to that of the LMC, can be the cause of the warp in-
stead (Ibata & Razoumov 1998).
A limitation of the present work is that the halo has not
been considered as a live, self-gravitating component. It has
been shown (Dubinski & Kuijken 1995; Nelson & Tremaine
1995) that the back-reaction of the halo on a re-aligning
disk can have important consequences. Such effects will be
the subject of a further paper.
APPENDIX A: POTENTIAL OF
AXISYMMETRIC DISK DUE TO A SATELLITE
The potential energy of an disk of surface density Σ(r) and
in the gravitational field due to a satellite at position r S is
given by
V = −Z d2
r GΣ(r)
MS
r − rS
.
(A1)
Choosing spherical coordinates for the satellite's posi-
tion (see Figure 1), and Cartesian coordinates in the disk
plane so that the satellite has x = 0, we have
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Assuming that the disk is small compared to rS, we can
expand the integrand in x and y. For an axisymmetric disk
the second-order terms are the first ones that generate a
potential gradient: they are
V = −
GMS
S Z Σdx dy [− 1
r3
2 (1 − 3 sin2 θS)y2 − 1
2 x2]
which results in
3GMSI1
V = −
2r3
S
sin2 θS + constant.
(A3)
(A4)
REFERENCES
Barnes, J., Hut, P., 1986, Nature, 324, 446
Binney, J., 1992, Annu. Rev. Astron. Astrophysics, 30, 51
Burke, B. F., 1957, AJ, 62, 90
Dubinski, J., 2000, in preparation.
Dubinski, J., Kuijken, K., 1995, ApJ, 442, 492
Hunter, C., Toomre, A., 1969, ApJ, 155, 747
Ibata, R. A., Razoumov, A. O., 1998, A&A, 336, 130
Kerr, F. J., 1957, AJ, 62, 93
Kuijken, K., 1997, in ASP Conference Series, Vol. 117, Dark and
Visible Matter in Galaxies, ed. Massimo Persic and Paolo
Salucci, pp. 220
Lin, D. N. C., Jones, B. F., Klemola, A. R., 1995, ApJ, 439, 652
Lynden-Bell, D. 1985, in The Milky Way Galaxy, ed. H. van Wo-
erden (Dordrecht: Reidel), 461
Murai, T., Fujimoto, M., 1980, Publ. Astron. Soc. Japan, 32, 581
Nelson, R. W., Tremaine, S., 1995, MNRAS, 275, 897
Schommer, R. A., Olszewski, E. W., Suntzeff, N. B., Harris, H.
C., 1992, AJ, 103, 447
Sparke, L. S., Casertano, S., 1988, MNRAS, 234, 873
Weinberg, M. D., 1998, MNRAS, 299, 499
|
astro-ph/0412695 | 1 | 0412 | 2004-12-30T12:03:41 | Rapid X-ray Variability of Seyfert 1 Galaxies | [
"astro-ph"
] | The rapid and seemingly random fluctuations in X-ray luminosity of Seyfert galaxies provided early support for the standard model in which Seyferts are powered by a supermassive black hole fed from an accretion disc. However, since EXOSAT there has been little opportunity to advance our understanding of the most rapid X-ray variability. Observations with XMM-Newton have changed this. We discuss some recent results obtained from XMM-Newton observations of Seyfert 1 galaxies. Particular attention will be given to the remarkable similarity found between the timing properties of Seyferts and black hole X-ray binaries, including the power spectrum and the cross spectrum (time delays and coherence), and their implications for the physical processes at work in Seyferts. | astro-ph | astro-ph |
Rapid X-ray Variability of Seyfert 1 Galaxies
S. Vaughan ([email protected])
X-Ray and Observational Astronomy Group, University of Leicester, Leicester,
LE1 7RH, U.K.
A. C. Fabian and K. Iwasawa
Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge
CB3 0HA, U. K.
Abstract. The rapid and seemingly random fluctuations in X-ray luminosity of
Seyfert galaxies provided early support for the standard model in which Seyferts
are powered by a supermassive black hole fed from an accretion disc. However, since
EXOSAT there has been little opportunity to advance our understanding of the
most rapid X-ray variability. Observations with XMM-Newton have changed this.
We discuss some recent results obtained from XMM-Newton observations of
Seyfert 1 galaxies. Particular attention will be given to the remarkable similarity
found between the timing properties of Seyferts and black hole X-ray binaries,
including the power spectrum and the cross spectrum (time delays and coherence),
and their implications for the physical processes at work in Seyferts.
1. In the beginning...
X-ray variability appears to be ubiquitous in Active Galactic Nuclei
(AGN). The rapid and seemingly random fluctuations in the X-ray
luminosity of Seyfert galaxies provided early support for the standard
black hole/accretion disc model (Rees, 1984) by implying compact
emission regions and high luminosity densities (Barr & Mushotzky,
1986).
2. The EXOSAT era
EXOSAT (1983 -- 1986) was the first mission to provide long (∼ 3 day),
uninterrupted X-ray observations of Seyfert galaxies. From these ob-
servations the X-ray power spectra (see van der Klis, 1989) of Seyfert
galaxies were measured for the first time (Lawrence et al., 1987; Green
et al., 1993; Lawrence & Papadakis, 1993). The EXOSAT observations
showed that the power spectra of Seyferts above ∼ 10−5 Hz could be
approximated by a power-law: P(f ) ∝ f −α where P(f ) is the power at
frequency f and α is the power spectrum slope. The measured slopes
from the EXOSAT observations were typically α ≈ 1.5. Processes such
as these, which have broad-band power spectra with more power at
c(cid:13) 2018 Kluwer Academic Publishers. Printed in the Netherlands.
vaughan_edited.tex; 29/09/2018; 13:02; p.1
2
Vaughan, Fabian & Iwasawa
Figure 1. XMM-Newton light curve of Mrk 766 binned to 100 second resolution.
lower frequencies, are called "red noise" (see Press, 1978). It was noted
early on (Lawrence et al., 1987) that this red noise variability of Seyferts
is similar to that observed in Galactic Black Hole Candidates (GBHCs;
Belloni & Hasinger, 1990; Nowak et al., 1999; McClintock & Remillard,
2004), perhaps suggesting that the same physical processes operate in
these sources that differ in black hole mass by factors of ∼
> 105.
3. Low frequency power spectra from RXTE
The steep slopes found in the EXOSAT power spectra required there to
be a flattening at even lower frequencies (so that the integrated power
remains finite). In recent years long RXTE monitoring observations
have detected these breaks (e.g. Uttley et al., 2002; Markowitz et al.,
2003; see also the article by Ian McHardy in these proceedings). Below
the break the slope is typically αlo ≈ 1 and at frequencies above the
break the slope is αhi ≈ 2. (The EXOSAT power spectra spanned
intermediate frequencies and often measured an intermediate slope
over the break.) The breaks represent "characteristic timescales" in
the aperiodic variability of Seyferts and, significantly, appear to scale
linearly with the mass of the central black hole: fbreak ∝ 1/MBH.
4. XMM-Newton results: high frequency power spectra
Until the launch of XMM-Newton (Jansen et al., 2001) high frequency
timing studies of Seyferts were not able to substantially improve on
the EXOSAT results. XMM-Newton's success is due to a combination
of high throughput, broad energy bandpass and long (∼ 2 day) orbit.
Figure 1 shows an example of a broad-band (0.2 − 10 keV) light curve
from a single orbit observation.
vaughan_edited.tex; 29/09/2018; 13:02; p.2
Rapid variability of Seyfert 1s
3
Figure 2. XMM-Newton power spectra of three Seyfert 1 galaxies. Shown are the
binned data (histograms) unfolded using the best-fitting broken power law con-
tinuum model (solid line). Note that the ordinate is in f × P (f ) units (a slope
of α = 1 would appear flat). Note that the figure shows unfolded data, meaning
the data/model residuals multiplied by the best-fitting model. This is necessarily a
model-dependent procedure and is not advisable if sharp features are present in the
spectrum. However, for these rather broad, smooth spectra unfolding does provide
a clear, if slightly crude, impression of the shape of the underlying spectrum free
from sampling effects.
Several Seyfert 1 galaxies have been studied with long XMM-Newton
observations and yielded interesting power spectra. These include: NGC
4051 (McHardy et al., 2004), Mrk 766 (Vaughan & Fabian, 2003),
MCG−6-30-15 (Vaughan et al., 2003), NGC 4395 (Vaughan et al., 2004)
(and also Ark 564; Vignali et al., 2004). Figure 2 shows the power spec-
tra for three of these. The XMM-Newton results clearly reveal similar
high frequency breaks in the power spectra (also measured by RXTE in
some cases) but clearly show a substantial object-to-object differences
in the normalisation of the power spectrum (which describes the overall
variability amplitude).
These XMM-Newton observations have also demonstrated the en-
ergy dependence of the power spectrum. Above the break frequency the
slope αhi tends to be steeper at lower energies. This is clearly observed
in MCG−6-30-15 (Vaughan et al., 2003) and NGC 4051 (McHardy et
al., 2004) but is not constrained by the other observations. This energy
dependence was also measured in NGC 7469 using an intensive RXTE
monitoring campaign (Nandra & Papadakis, 2001).
vaughan_edited.tex; 29/09/2018; 13:02; p.3
4
Vaughan, Fabian & Iwasawa
5. XMM-Newton results: high frequency cross spectrum
In addition to the energy dependence of the power spectrum, the excel-
lent quality XMM-Newton light curves have allowed the cross spectrum
to be investigated in several Seyfert 1s for the first time. Prior to XMM-
Newton only Papadakis et al. (2001) had measured the cross spectrum
for a Seyfert (NGC 7469 using RXTE).
The cross spectrum compares the variations in one band with those
in another as a function of frequency. The amplitude of the cross
spectrum gives the coherence (Vaughan & Nowak, 1997) while the
argument gives the phase lag (time delay; Nowak et al., 1999). The
coherence quantifies any (linear) correlation between the variations
in the two bands, irrespective of any time delays. The observations
typically show high coherence at the lowest frequencies (i.e. strong
correlation) with a decrease at higher frequencies which implies there
are independent variations between the two bands occurring on short
timescales (Vaughan et al., 2003; McHardy et al., 2004; Vaughan et
al., 2004). At low frequencies, where the coherence is high, the data
also exhibit small time delays, with the soft leading the hard variations
(Vaughan et al., 2003; McHardy et al., 2004). The magnitude of the
time delay decreases with increasing frequency (although the functional
form of the relation is poorly constrained).
6. Summary of results
XMM-Newton has already made significant progress towards improving
our understanding of the high frequency variability of Seyfert galaxies.
The timing studies have revealed:
− Similar broken power spectra in Seyferts but object-to-object dif-
ferences in normalisation (variability amplitude) and high frequency
slope.
− Energy-dependent high frequency power spectrum slope (steeper
at lower energies).
− High coherence at low frequencies, falling off at high frequencies.
− Small (∆T ∼ 0.01/f ) soft-to-hard time delays
vaughan_edited.tex; 29/09/2018; 13:02; p.4
Rapid variability of Seyfert 1s
5
Figure 3. The MBH − fbr relation for 11 Seyfert galaxies. The masses are from
reverberation mapping experiments except for the four objects marked using dotted
error bars. Also shown are typical break frequencies for Cyg X-1 in both its low/hard
(H) and high/soft (S) states. The dotted lines show example MBH ∝ 1/fbr relations
consistent with the Cyg X-1 points.
7. Comparison with GBHCs
The frequencies of the breaks in the power spectra are broadly consis-
tent with the long-held notion that the characteristic frequencies should
scale as ∝ 1/MBH right down to stellar mass black holes. Figure 3
shows the available data for 11 Seyferts. Although the uncertainties
are rather large, the data seem consistent with an extrapolation of
the fbr ∝ 1/MBH relation from the well-studied GBHC Cygnus X-
1 (Belloni & Hasinger, 1990; Cui et al., 1996; Nowak et al., 1999;
McClintock & Remillard, 2004). Note that the break frequency mea-
surements come from a combination of XMM-Newton and RXTE ob-
servations. Long XMM-Newton observations are sensitive to breaks in
the range ∼ 10−4 − 10−2 Hz while the RXTE monitoring campaigns
are sensitive to breaks at lower frequencies.
The connection between Seyferts and GBHCs is reinforced by the
similarity between their cross spectra. It is well known that GBHCs
show highly coherent variations at low frequencies with the coher-
ence fading away at the highest frequencies plus frequency dependent
time lags similar to those measured in Seyferts (Nowak et al., 1999;
vaughan_edited.tex; 29/09/2018; 13:02; p.5
6
Vaughan, Fabian & Iwasawa
McClintock & Remillard, 2004). These all argue for a common mech-
anism responsible for producing the X-ray variability in Seyferts and
GBHCs.
One may ask what advantage is gained by studying Seyferts in X-
rays if GBHCs operate with the same physics but provide much higher
quality data? One answer is that Seyferts can in fact provide data
that are in some senses better than that from GBHCs for studying
the highest frequencies. For example, comparing MCG−6-30-15 and
Cygnus X-1 we see the timescales are longer by ∼ 105 in the Seyfert, but
the X-ray flux is smaller by ∼ 103. This means that per characteristic
timescale the Seyfert provides ∼ 102 more photons! Of course, GBHC
enthusiasts can argue that GBHCs reclaim much of their advantage
even here because one can always observe many (∼ 105) more samples
of a given timescale in a fixed amount of observing time, even if many
less photons are recorded per timescale. Even so the power spectrum
of MCG−6-30-15 could be measured up to ∼ 5 × 10−3 Hz using the
XMM-Newton data. This is equivalent to probing ∼ 500 Hz in Cygnus
X-1, a challenge for even the best RXTE observations (Revnivtsev et
al., 2000).
8. Implications for the emission processes
The X-ray emission mechanism operating in Seyfert galaxies (and GB-
HCs) is usually thought to be inverse-Compton scattering. In the sim-
plest models harder photons are expected to lag behind the softer
photons due to the larger number of scatterings required to produce
harder photons; the delay should be of order the light-crossing time
of the corona. The direction and magnitude of the observed time lags
in Seyfert 1s are consistent with an origin in a Comptonising corona.
However, if the lags are frequency dependent (as expected by analogy
with Cygnus X-1) the lags at lower temporal frequencies would become
much longer than expected for a compact corona (see discussion in
Nowak et al., 1999). In addition, some models of Compton scattering
coronae predict the high frequency PSD should be steeper for higher
energy photons, due to the high-frequency fluctuations being washed
out by multiple scatterings (Nowak & Vaughan, 1996), contrary to
the observations. Alternatively, the time delay between soft and hard
bands could be due to the spectral evolution of individual X-ray events
(Poutanen & Fabian, 1999) or propagation of accretion rate variations
through a extended emission region (Kotov et al., 2001).
vaughan_edited.tex; 29/09/2018; 13:02; p.6
Rapid variability of Seyfert 1s
7
Acknowledgements
Based on observations obtained with XMM-Newton, an ESA science
mission with instruments and contributions directly funded by ESA
Member States and the USA (NASA). SV and KI are supported by
the UK PPARC. ACF is supported by the Royal Society.
References
Barr P., Mushotzky R. F., 1986, Nature, 320, 421
Belloni T., Hasinger G., 1990, A&A, 227, L33
Cui W., Zhang S. N., Focke W., Swank J. H., ApJ, 484, 383
Green A. R., McHardy I. M., Lehto H. J., 1993, MNRAS, 265, 664
Jansen F. et al. 2001, A&A, 365, L1
Kotov O., Churazov E., Gilfanov M., 2001, MNRAS, 327, 799
Lawrence A., Watson M. G., Pounds K. A., Elvis M., 1987, Nature, 325, 694
Lawrence A., Papadakis I., 1993, ApJ, 414, L85
Markowitz A. et al. 2003, ApJ, 593, 96
McClintock J. E., Remillard R. A., 2004, in Lewin W. H. G., van der Klis M.,
eds., Compact Stellar X-ray Sources, Cambridge University Press (Cambridge),
in press (astro-ph/0306213)
McHardy I. M., 1989, in J. Hunt, B. Battrick, eds, Two Topics in X Ray Astronomy,
(ESA SP-296; Noordwijk: ESA), p1111
McHardy I. M., Papadakis I. E., Uttley P., Page M., Mason K., 2004, 348, 783
Nandra K., Papadakis I. E., 2001, ApJ, 554, 710
Nowak M. A., Vaughan B. A., 1996, MNRAS, 280, 227
Nowak M. A., Vaughan B. A., Wilms J., Dove J. B., Begelman M. C., 1999, ApJ,
510, 874
Papadakis I. E., Nandra K., Kazanas D., 2001, ApJ, 554, L133
Poutanen J., Fabian A. C., 1999, MNRAS, 306, L31
Press W. H., 1978, Comments on Astrophysics, 7, 103
Rees M. J., 1984, ARA&A, 22, 471
Revnivtsev M., Gilfanov M., Churazov E., 2000, A&A, 363, 1013
Uttley P., McHardy I. M., 2001, MNRAS, 323, L26
Uttley P., McHardy I. M., Papadakis I., 2002, MNRAS, 332, 231
van der Klis M., 1989, in H. Ogelman, E. P. J. van den Heuvel eds., Timing Neutron
Stars, Kluwer (Dordrecht), NATO ASI Series C 262, p27
Vaughan B. A., Nowak M. A., 1997, ApJ, 474, L43
Vaughan S., Fabian A. C., 2003, MNRAS, 341, 496
Vaughan S., Fabian A. C., Nandra K., 2003, MNRAS, 339, 1237
Vaughan S., Iwasawa K., Fabian A. C., Hayashida K., 2004, MNRAS, submitted
Vignali C., Brandt W. N., Boller Th., Fabian A. C., Vaughan, S., 2004, MNRAS,
347, 854
vaughan_edited.tex; 29/09/2018; 13:02; p.7
vaughan_edited.tex; 29/09/2018; 13:02; p.8
|
0711.4241 | 1 | 0711 | 2007-11-27T12:47:16 | Radial Dependence of Extinction in Parent Galaxies of Supernovae | [
"astro-ph"
] | The problem of extinction is the most important issue to be dealt with in the process of obtaining true absolute magnitudes of core-collapse (including stripped-envelope) supernovae (SNe). The plane-parallel model, widely used in the past, was shown not to describe extinction adequately. We try to apply an alternative model which introduces radial dependance of extinction in parent galaxies of supernovae. For calculating extinction in our Galaxy we use two different methods and compare the results obtained. Our analysis is primarily focused on a chosen sample of stripped-envelope SNe (Ib/c) for which we find intrinsic peak absolute magnitude $\mathrm{M}_{\mathrm{B}}^{0}=-17.80\pm 0.43$. | astro-ph | astro-ph |
Documentation (2008)
(2008)
RADIAL DEPENDENCE OF EXTINCTION IN PARENT GALAXIES OF
SUPERNOVAE
D. Oni´c, B. Arbutina and D. Urosevi´c
Department of Astronomy, Faculty of Mathematics, University of Belgrade, Serbia
Version 3.16, 2003/04/07
RESUMEN
El problema de la extinci´on es el asunto m´as importante que se ocupar´a en
del proceso de obtener magnitudes absolutas verdaderas de corazo'n-se derrumba
(pelar-sobre incluyendo) las supernovas (SNe). El modelo plano-paralelo, usado ex-
tensamente en el pasado, fue demostrado para no describir la extinci´on adecuada-
mente. Intentamos aplicar un modelo alternativo que introduzca la dependencia
radial de la extinci´on en galaxias del padre de supernovas. Para la extinci´on cal-
culadora en nuestra galaxia utilizamos dos diversos m´etodos y comparamos los
resultados obtenidos. Nuestro an´alisis se centra sobretodo en una muestra elegida
del pelar-sobre SNe (Ib/c) para el cual encontramos la magnitud absoluta m´axima
intr´ınseca M0
B = −17.80 ± 0.43.
ABSTRACT
The problem of extinction is the most important issue to be dealt with in the
process of obtaining true absolute magnitudes of core-collapse (including stripped-
envelope) supernovae (SNe). The plane-parallel model, widely used in the past, was
shown not to describe extinction adequately. We try to apply an alternative model
which introduces radial dependance of extinction in parent galaxies of supernovae.
For calculating extinction in our Galaxy we use two different methods and com-
pare the results obtained. Our analysis is primarily focused on a chosen sample of
stripped-envelope SNe (Ib/c) for which we find intrinsic peak absolute magnitude
M0
B = −17.80 ± 0.43.
Key Words: SUPERNOVAE: GENERAL -- GALAXIES: SPIRAL -- ISM:
DUST, EXTINCTION.
1. INTRODUCTION
As well known, the supernovae (hereafter SNe)
type Ia are widely used by astronomers as distance
indicators due to their small dispersion in peak ab-
solute magnitude and similar light curves. Today,
they are thought to originate primarily in the explo-
sion of a Chandrasekhar mass C-O white dwarf in
close binary system, where the secondary star is fill-
ing its Roche lobe and transferring mass through the
Lagrange point to the white dwarf companion. SNe
type II, on the other hand, are a quite heterogenous
class. They come from stars of initial mass greater
than approximately 8 M⊙, which can have quite dif-
ferent properties. The situation is still unclear re-
garding stripped-envelope SNe (Ib/c). The progen-
itors of these SNe are massive stars that have lost
most or all of their hydrogen/helium envelopes, by
strong winds such as in Wolf-Rayet stars or through
mass transfer to a companion star in Roche lobe
overflow or a common envelope phase.
Whatever
the
exact
scenario for
stripped-
envelope SNe is, there is some quite unique physics
involved in producing these events characterized by
a complete loss of hydrogen/helium. This "unique-
ness" may lead to a smaller dispersion in their obser-
vational properties (e.g. peak brightness), at least in
comparison to more heterogeneous SNe type II. SNe
Ib/c thus may be the second best "standard candles"
among supernovae, after SNe Ia. Our intention is to
analyze this possibility. We will primarily focus on
finding the peak absolute magnitude for SNe Ib/c.
In doing so we must be able to eliminate extinction,
which in the case of all core-collapse SNe should be
significant.
1
2
ONI ´C, ARBUTINA & UROSEVI ´C
The problem of extinction is the most important
issue to be dealt with in the process of obtaining
true absolute magnitudes of core-collapse (including
stripped-envelope) SNe. The plane-parallel model
which gives absorption dependent on galaxy incli-
nation, Ag = A0 sec i, widely used in the past, was
shown not to describe extinction adequately (Cap-
pellaro 1997). We try to apply an alternative model
which introduces radial dependance of extinction
(Hatano et al. 1997, 1998).
A certain trend of dimmer SNe with decreasing
distance from the center of a galaxy was already
found by Arbutina (2007a,b). In present more de-
tailed analysis we have increased the number of SNe
in the sample, applied a new model for extinction
in the parent galaxy, and more carefully calculated
Galactic extinction.
2. THE MODEL
Because of their long-lived progenitors, only SNe
Ia have been observed in elliptical galaxies, which
are, practically, gas and dust free. In spirals SNe Ia
can be found in inter-arm regions and galaxy's halos.
Hence, extinction does not influence their luminosi-
ties as much as in the case of core-collapse SNe (Ib/c
and II) which are all observed in spiral and irregular
galaxies which have significant amounts of dust.
As just mentioned in the previous section, we
have chosen to work with SNe Ib/c because we ex-
pect them to be relatively homogenous class, at least
in comparison to SNe II. Ideally, we should further
separate SN Ib from Ic, but the sample is already
too small for doing this, as we shall see. On the
other hand, a sample of stripped-envelope SNe also
comprises peculiar SNe Ic (the so called hypernovae)
which can have very different properties, and thereby
were not included in this analysis.
Extinction has a selective nature, meaning it de-
pends on wavelength. Shorter wavelengths are more
weakened and therefore we expect blue magnitudes
(B) to be more extincted than, for example, visual
ones (V). Apparent peak B magnitudes for stripped-
envelope SNe that we have focused on have also
turned out to be more frequently available in our
primary data source (Asiago Supernova Catalogue,
Barbon et al. 1999).1 For the absolute peak blue
magnitude we can generally write:
M0
B = mB − µ − AG − Ag = MB − Ag,
(1)
where mB is apparent magnitude, µ = 5 log(d/Mpc)
is distance modulus, AG and Ag are Galactic extinc-
tion and extinction in parent galaxy, respectively.
1See Richardson et al. (2006) for a somewhat different anal-
ysis of SNe Ib/c peak magnitudes in the V band.
W
E
N
S
y
y'
*
x'
y'
i
x
Fig. 1. The figure shows how radial position of a super-
nova in its parent galaxy was calculated.
Bearing in mind the short life of their progenitors,
we may assume that the stripped-envelope SNe are
practically in the galactic plane (Z = 0). The radial
position of supernova in a galaxy is then (see Figure
1):
r2 = d2(x′2 + y′2 sec2 i) =
d2(x2 + y2)(cos2(arctan(
y
x
) + Π − 90◦) +
+ sin2(arctan(
y
x
) + Π − 90◦) sec2 i),
(2)
where x and y give SN offset from the center of the
galaxy in radians, Π is the position angle of the
major axis and d is the distance to the galaxy.
If
not given, offsets can be calculated from right as-
cension and declination; x ≈ (αSN − αg) cos δg and
y ≈ (δSN − δg).
If we assume that
Ag = A0e−ar/R,
then
MB = M0
B + A0e−ar/R.
(3)
(4)
R is the radius of a galaxy, and A0 and a parameters
which can be obtained from the fit.
Extinction has been already assumed to follow an
exponential law by Hatano et al. (1997). It is known
that dust density in our Galaxy and M31 actually
does not peak at the center (nor in bulge) but well
out in the disk, in a molecular ring where most of the
current star formation is taking place (see Figure 9
in Sodroski et al. 1997, Hatano et al. 1998). This
shortage of the model is not that important for two
reasons: (i) there is a selection effect against seeing
P
RADIAL DEPENDENCE OF EXTINCTION IN PARENT GALAXIES OF SUPERNOVAE
3
SNe in already bright bulge (Shaw 1979); (ii) we do
not expect (many) stripped-envelope SNe in old pop-
ulation rich bulge anyway. We also did not include
dependence of inclination in this preliminary model,
which would produce large errors if we had edge-on
galaxies with i → 90◦. Finally, spiral galaxies are
not azimuthally symmetric (there are spiral arms, of
course) and local environment in which star explodes
can be very different from one case to another, an ef-
fect that we can not account for.
3. DATA
3.1. Distances
When possible, the Cepheid-calibrated distance
to the parent galaxy, or a galaxy in the same
group of galaxies, was used. The distance mod-
ulus and its uncertainty were taken directly from
the literature (Richardson et al. 2006, Macri et al.
2001, Thim et al. 2003). The second choice was
the distance given in the Nearby Galaxy Catalogue
(Tully 1988), rescaled from H0 = 75km s−1 Mpc−1
to H0 = 65km s−1 Mpc−1 for consistency with Ar-
butina (2007a,b). Uncertainty of 0.2 mag in the dis-
tance modulus was adopted as in Richardson et al.
(2006). The third choice was the distance calculated
from the redshift of the parent galaxy (we have re-
stricted our analysis to the local universe, z < 0.03)
assuming H0 = 65km s−1 Mpc−1. The redshifts cor-
rected to the Cosmic Microwave Background (CMB)
Reference Frame were taken from the June 2007
version of the NED.2 The uncertainties ware calcu-
lated assuming a peculiar velocity of 300km s−1 as
in Richardson et al. (2006).
The distance modulus and its uncertainty for
SN 1994I were taken from Richmond et al. (1996)
(see the references therein) who adopted the surface
brightness fluctuations method, while for SN 2002ap
they were taken from Foley et al. (2003) (and refer-
ences therein) who used photometry of the brightest
red and blue stars in the system.
3.2. Peak Apparent Magnitudes
For most SNe, peak B magnitude was taken from
the June 2007 version of the Asiago Supernova Cat-
alogue (hereafter ASC, Barbon et al. 1999).3 Uncer-
tainty of 0.3 in apparent magnitudes listed in the
ASC, estimated by Richardson et al. (2002), was
adopted. The apparent B magnitude at the time
2The NASA/IPAC Extragalactic Database (NED)
is
operated by the Jet Propulsion Laboratory, California
Institute of Technology, under contract with the Na-
tional Aeronautics and Space Administration (available at
http://nedwww.ipac.caltech.edu/).
3Available at http://web.pd.astro.it/supern/
of maximum and its uncertainty for SN 1994I were
taken from Richmond et al. (1996), for SN 1999ex
from Stritzinger et al. (2002) and for SN 2004aw
from Taubenberger et al. (2006). The peak appar-
ent B magnitude for SN 1990B was estimated from
Clocchiatti et al. (2001) with the highest uncertainty
associated.
The peak apparent V magnitude and its uncer-
tainty were taken directly from Richardson et al.
(2006) except for SN 1972R and SN 2004aw for which
they were taken from Leibundgut et al. (1991) and
Taubenberger et al. (2006), respectively.
3.3. Galactic extinction
Two different methods for calculation of Galactic
extinction were used. We thought it would be useful
to use values from both methods and see what effect
this will have on the results (see Willick 1999).
The first choice were values taken from RC3 cata-
logue (de Vaucouleurs et al. 1991) who used the older
Burstein and Heiles maps (hereafter BH), which are
based on HI column density and faint galaxy counts
(Burstein & Heiles 1982). On the basis of the re-
sults of the investigation,
it is concluded that a
reasonable estimate of the relative accuracy of the
HI/galaxy counts method reported by BH is 0.01
mag in E(B − V) or 10% of the reddening, whichever
is larger (Burstein & Heiles 1982). The second choice
were values for Galactic extinction by the Schlegel,
Finkbeiner, and Davis maps (hereafter SFD), based
on IRAS/DIRBE measurements of diffuse IR emis-
sion (Schlegel et al. 1998), with values taken from
the NED. The reddening estimates from Schlegel et
al. (1998) have an accuracy of 16%.
Galactic extinction ASFD
G (V) and its uncertainty
were taken directly from Richardson et al. (2006)
except for SN 1972R and SN 2004aw for which they
were taken from NED.
3.4. Other data
The position angle of the major axis of the par-
ent galaxy (measured North Eastwards) was taken
from ASC, except for SN 1983N, SN 1984I, SN 1991N
and SN 1998bw for which it was taken from the July
2007 version of the SAI supernova catalogue.4 The
SN type, the parent galaxy type, the inclination of
the polar axis with respect to the line of sight in de-
grees (0◦ for face on systems), the SN offset from the
galaxy nucleus in arcseconds, and decimal logarithm
4Sternberg Astronomical Institute Supernova Catalogue
Tsvetkov, N.N. Pavlyuk, O.S. Bar-
(SAI) by D.Yu.
tunov and Yu.P. Pskovskii, Sternberg Astronomical
In-
stitute, Moscow University, Moscow, Russia (available at
http://www.sai.msu.su/sn/sncat/).
4
ONI ´C, ARBUTINA & UROSEVI ´C
of the apparent isophotal diameter in units of 0.1
arcminutes, were taken from ASC.
There is some discrepancy regarding the type of
SN 1972R (Ib in ASC, Ia in Patat et al. 1997, Ib? in
Leibundgut et al. 1991, Ipec in SAI supernova cata-
logue) and SN 1999ex (Ib/c in ASC, Ib in Richardson
et al. 2006). We have marked these as Ib? and Ib/c,
respectivley.
4. ANALYSIS AND RESULTS
Table 1 gives a sample of SNe Ib/c with the
known peak apparent B magnitudes and the calcu-
lated SN radial positions. Table 2 gives redshift and
distance modulus for each supernova in the sample.
Table 3 and 4 give absolute B magnitudes at max-
imum light, uncorrected for the parent galaxy ex-
tinction, for the same sample of SNe, with BH and
SFD model used for calculation of Galactic extinc-
tion, respectively. Table 5 gives peak V absolute
magnitudes, uncorrected for the parent galaxy ex-
tinction, with only SFD method for calculation of
Galactic extinction used.
We see in Figure 2 and 4 that there is a certain
trend of dimmer SNe with decreasing distance from
the center of a galaxy which can be attributed to
extinction. It can also be seen in Figure 2 that SNe
Ic show larger dispersion then SNe Ib.
Peculiar SNe Ic, also known as hypernovae, show
very large dispersion (see Figure 2), they may have
very different properties and are thereby excluded
from the fit. SN 2004aw appears to be a link be-
tween a normal Type Ic supernova like SN 1994I
and the group of broad-lined SNe Ic like 1998bw
and 2002ap (Taubenberger et al. 2006). NGC 3997,
host of SN 2004aw, could be a merger system of
two spiral galaxies showing tidally deformed spiral
arms (Taubenberger et al. 2006), which may explain
the SN position. Another supernova that was ex-
cluded from the fit is SN 1990B. It shows some kind
of anomalous extinction. The reasons may lie in a
specific environment. Unusually high value MB/V
could be due to the high value of the host galaxy
extinction (as estimated in Richardson et al. 2006).5
In Figure 3 The SN V absolute magnitudes, un-
corrected for the parent galaxy extinction, are plot-
ted against relative radial position of a SN in the
galaxy, with SFD method for calculation of Galactic
extinction used. Contrary to the case of B magni-
tudes, a significant trend of dimmer SNe with de-
5There are few SN Ib/c from ASC excluded from this anal-
ysis; SN 1954A is located in an irregular galaxy, SN 1966J was
shown to be SN Ia (Casebeer et al. 2000), SN 2001B was ex-
cluded because of the unknown peak B luminosity, SN 2006tq
because of some other unknown properties.
-20
-19
-18
-17
D
F
S
B
M
-16
-15
-14
-13
SNe Ib
SNe Ic
SNe Ic pec
0,0
0,2
0,4
0,6
0,8
1,2
1,4
1,6
1,8
2,0
1,0
r/R
Fig. 2. The SN B absolute magnitudes, uncorrected for
the parent galaxy extinction, are plotted against the rel-
ative radial positions of a SN in the galaxy. The SFD
method for calculation of Galactic extinction is used.
creasing radius cannot be observed here. This is
understandable because the extinction should have
weaker influence on V magnitudes than B magni-
tudes.
-19
-18
-17
D
F
S
V
M
SNe Ic
SNe Ib
-16
0,0
0,2
0,4
0,6
0,8
1,0
r/R
Fig. 3. The SN V absolute magnitudes, uncorrected for
the parent galaxy extinction, are plotted against relative
radial positions of a SN in the galaxy. The SFD method
for the calculation of Galactic extinction is used. SNe
Ic pec are excluded, as well as SN 1990B.
B = −17.86 ± 0.46, for BH method used and M0
If we assume that extinction is negligible when
r/R → ∞ a fit to data can give us intrinsic absolute
magnitude for SNe Ib/c (see Figure 4). These are:
M0
B =
−17.74 ± 0.40, for SFD method used. Figure 5 shows
the effect of BH/SFD method used in calculation of
Galactic extinction on the assumed extinction trend
and resulting peak absolute magnitude. As we can
RADIAL DEPENDENCE OF EXTINCTION IN PARENT GALAXIES OF SUPERNOVAE
5
see the effect is not large. We will thereby adopt the
mean value:
M0
B = −17.80 ± 0.43.
(5)
Curves in Figures 4 and 5 have the form of equa-
tions (3) i.e. (4). In Table 6 we give all fit parameters.
-19
-18
-17
D
F
S
B
M
-16
SNe Ib/c
-15
0,0
0,2
0,4
0,6
0,8
1,0
r/R
Fig. 4. The curve represents the best fit for the data
sample of 12 SNe Ib/c. The SFD method for calculation
of Galactic extinction is used. SN 1990B is excluded from
the fit.
BH
SFD
-19
-18
-17
D
F
S
B
M
-16
SNe Ib/c
-15
27
28
29
30
31
32
33
34
35
Fig. 6. The MSFD
B
dependance on distance modulus.
5. CONCLUSIONS
Measuring the radial dependence of controlled
samples of SNe in galaxies appears to be a promising
way to constrain the amount and the distribution of
dust in galaxies (Hatano et al. 1998). As pointed
out by Hatano et al. (1998) the extinction by dust
in parent galaxies can also affect the observed SNe
properties.
In this paper we analysed a sample of SNe Ib/c
in neighboring spiral galaxies. We have chosen to
work with SNe Ib/c since we expect them to be rel-
atively homogenous class, at least in comparison to
SNe II. B magnitudes were considered because the
extinction has stronger influence on them, as well
as because of the larger quantity of data available
(in ASC). A certain trend of dimmer SNe with de-
creasing distance from the center of a galaxy was
observed. Such an effect can be attributed to ex-
tinction. We have adopted the mean value for SNe
Ib/c intrinsic absolute magnitude:
-19
-18
-17
B
M
-16
-15
0,0
0,2
0,4
0,6
0,8
1,0
r/R
M0
B = −17.80 ± 0.43,
(6)
Fig. 5. The effect of BH/SFD method used in determin-
ing peak absolute magnitude on the results.
It is still possible that there is a substantial range
of absolute magnitudes for Ib/c events. In that case
we will have the Malmquist bias -- only brighter
SNe would be observed at large distances (high z)
(see Richardson et al. 2006). Figure 6 shows MSFD
against distance modulus for analysed Ib/c super-
novae. We do not see any distance dependance,
which is quite understandable since we have limited
our analysis to the local universe (z < 0.03).
B
obtained from the fit.
We made comparison of the results obtained
by using values for AG from different methods for
Galactic extinction calculation. The effect is not sig-
nificant. Both methods give similar results.
To conclude, analysing the dependence of SN ob-
servational properties on radial position in their par-
ent galaxies may hold some promise. Nevertheless,
more data are needed to reach stronger conclusions.
The authors would like to thank the referee Prof.
David Branch for useful comments. This paper is
6
ONI ´C, ARBUTINA & UROSEVI ´C
TABLE 6
FIT PARAMETERS
Method
BH
SFD
M0
B
A0
a
−17.86 ± 0.46 2.77 ± 0.51 3.89 ± 2.55
−17.74 ± 0.40 2.60 ± 0.71 4.55 ± 3.38
a part of the projects "Gaseous and Stellar Compo-
nents of Galaxies: Interaction and Evolutions" (No.
146012) and "Physics of Sun and Stars" (No. 146003)
supported by the Ministry of Science of Serbia.
REFERENCES
Arbutina, B. 2007a, IJMPD, 16, 1219
Arbutina, B. 2007b, AIP Conf. Proc., 938, 202
Barbon, R., Boundi, V., Cappellaro, E., & Turatto, M.
1999, A&A, 139, 531
Burstein D. & Heiles, C. 1982, AJ, 87, 1165
Cappellaro, E., Turatto, M., Tsvetkov, D. Yu., Bartunov,
O. S., Pollas, C., Evans R., & Humuy, M. 1997, A&A,
322, 431
Casebeer, G. et al. 2000, PASP, 112, 1433
Clocchiatti, A. et al 2001, ApJ, 553, 886
Foley, R. J. et al. 2003, PASP, 115, 1220
Hatano, K., Branch, D., Fisher, A. & Starrfield, S. 1997,
MNRAS, 290, 113
Hatano, K., Branch, D., & Deaton, J. 1998, ApJ, 502,
177
Leibundgut, B. et al. 1991, AA, 89, 537
Macri, L. M. et al. 2001, ApJ, 559, 243
Patat, F. et al. 1997, A&A, 317, 423
Richardson, D., Branch, D., & Baron, E. 2006, AJ, 131,
2233
Richardson, D., Branch, D., & Baron, E. 2002, AJ, 123,
745
Richmond, M. W. et al. 1996, AJ, 111, 327
Schlegel, D. et al. 1998, ApJ, 500, 525
Shaw, R. L. 1979, A&A, 76, 188
Sodroski T. J., Odegard, N., Arendt, R. G., Dwek, E.,
Weiland, J. L., Hauser, M. G., & Kelsall, T. 1997,
ApJ, 480, 173
Stritzinger, M. et al. 2002, AJ, 124, 2100
Taubenberger, S. et al. 2006, MNRAS, 371, 1459
Thim, F. et al. 2003, ApJ, 590, 256
Tully, R. B. 1988, Nearby Galaxy Catalogue (Cambridge
University Press)
de Vaucouleurs, G., de Vaucouleurs, A., Corwin, H.G.,
Buta, R.J., Paturel G. & Foque, P. 1991, Third Ref-
erence Catalogue of Bright Galaxies (Springer-Verlag,
New York)
Willick, J. A. 1999, ApJ, 522, 647
D. Oni´c, B. Arbutina and D. Urosevi´c: Department of Astronomy, Faculty of Mathematics, University of
Belgrade, Studentski trg 16, 11000 Belgrade, Serbia (donic, arbo, [email protected])
RADIAL DEPENDENCE OF EXTINCTION IN PARENT GALAXIES OF SUPERNOVAE
7
DATA AND RADIAL POSITIONS FOR SUPERNOVAE FROM THE SAMPLE
TABLE 1
Inclination Position Diameter
D [kpc]
Supernova
Galaxy
SN
type
Galaxy
type
NGC 2841
Ib?
SN 1972R
NGC 5236
Ib
SN 1983N
E323-G99
Ib
SN 1984I
NGC 991
Ib
SN 1984L
IC 454
Ib
SN 2000H
IC 5179
Ib/c
SN 1999ex
NGC 1073
Ic
SN 1962L
NGC 4051
Ic
SN 1983I
NGC 1365
SN 1983V
Ic
NGC 2715
SN 1987M Ic
NGC 3310
Ic
SN 1991N
NGC 5194
SN 1994I
Ic
SN 1990B
NGC 4568
Ic
SN 2004aw Ic pec NGC 3997
SN 1998bw Ic pec E184-G82
SN 2002ap
Ic pec NGC 628
Sb
SBc
SBcd
SBc
SBab
Sbc
SBc
SBbc
SBb
SBc
SBbc
Sbc
Sbc
SBb pec
SB
Sc
i[◦]
65
21
25
28
58
61
25
35
58
74
19
48
65
68
33
24
angle
Π[◦]
147
45
10
60
140
57
15
135
32
22
139
163
23
130
150
25
SN radial
position
r[kpc]
12.6
3.5
11.7
4.1
9.5
2.5
6.6
7.2
7.2
2.4
0.9
1.1
0.9
33.4
2.7
10.3
Apparent
magnitude
mB
12.85 ± 0.30
11.70 ± 0.30
16.60 ± 0.30
14.00 ± 0.30
17.90 ± 0.30
17.35 ± 0.02
13.94 ± 0.30
13.70 ± 0.30
14.67 ± 0.30
15.30 ± 0.30
15.50 ± 0.30
13.77 ± 0.02
16.89 ± 0.60
18.06 ± 0.04
14.09 ± 0.30
13.11 ± 0.30
32
17
32
19
32
33
24
34
57
33
17
24
20
38
9
22
8
ONI ´C, ARBUTINA & UROSEVI ´C
TABLE 2
SN REDSHIFT AND DISTANCE MODULUS
Supernova
Redshift
z
0.002128
SN 1972R
0.001711
SN 1983N
0.010737
SN 1984I
0.005110
SN 1984L
0.013159
SN 2000H
0.011415
SN 1999ex
0.004030
SN 1962L
0.002336
SN 1983I
SN 1983V
0.005457
SN 1987M 0.004466
SN 1991N
0.003312
0.001544
SN 1994I
SN 1990B
0.007522
SN 2004aw 0.015914
SN 1998bw 0.008670
SN 2002ap
0.002192
Distance
modulus
µ
30.74 ± 0.23
28.25 ± 0.15
33.64 ± 0.20
31.68 ± 0.20
33.99 ± 0.17
33.44 ± 0.22
31.22 ± 0.20
31.72 ± 0.14
31.27 ± 0.05
31.86 ± 0.20
31.67 ± 0.20
29.60 ± 0.30
30.92 ± 0.05
34.46 ± 0.14
32.93 ± 0.28
29.30 ± 0.14
Ref.
1
2
3
4
3
3
4
5
5
4
4
6
5
3
3
7
REFERENCES. - (1) Macri et al. 2001; (2) Thim et al. 2003; (3) NED (References are for redshifts); (4) Tully 1988;
(5) Richardson et al. 2006; (6) Richmond et al. 1996; (7) Foley et al. 2003.
RADIAL DEPENDENCE OF EXTINCTION IN PARENT GALAXIES OF SUPERNOVAE
9
TABLE 3
ABSOLUTE MAGNITUDE MBH
B
Supernova
Galactic
absorption
ABH
G (B)
Absolute
magnitudea
MBH
B
0.00 ± 0.04 −17.89 ± 0.57
SN 1972R
0.15 ± 0.04 −16.70 ± 0.49
SN 1983N
0.45 ± 0.04 −17.49 ± 0.54
SN 1984I
0.00 ± 0.04 −17.68 ± 0.54
SN 1984L
1.44 ± 0.04 −17.53 ± 0.51
SN 2000H
0.00 ± 0.04 −16.09 ± 0.28
SN 1999ex
0.07 ± 0.04 −17.35 ± 0.54
SN 1962L
0.00 ± 0.04 −18.02 ± 0.48
SN 1983I
0.00 ± 0.04 −16.60 ± 0.39
SN 1983V
SN 1987M 0.02 ± 0.04 −16.58 ± 0.54
0.00 ± 0.04 −16.17 ± 0.54
SN 1991N
0.00 ± 0.04 −15.83 ± 0.36
SN 1994I
SN 1990B
0.01 ± 0.04 −14.04 ± 0.69
SN 2004aw 0.00 ± 0.04 −16.40 ± 0.22
SN 1998bw 0.00 ± 0.04 −18.84 ± 0.62
SN 2002ap
0.13 ± 0.04 −16.32 ± 0.48
a MBH
B is absolute magnitude uncorrected for extinction in the parent galaxy.
10
ONI ´C, ARBUTINA & UROSEVI ´C
TABLE 4
ABSOLUTE MAGNITUDE MSFD
B
Supernova
Galactic
absorption
ASFD
G (B)
Absolute
magnitudea
MSFD
B
0.067 ± 0.011 −17.96 ± 0.54
SN 1972R
0.284 ± 0.045 −16.83 ± 0.50
SN 1983N
0.452 ± 0.072 −17.49 ± 0.57
SN 1984I
0.119 ± 0.019 −17.80 ± 0.52
SN 1984L
0.991 ± 0.159 −17.08 ± 0.63
SN 2000H
0.087 ± 0.014 −16.18 ± 0.25
SN 1999ex
0.169 ± 0.027 −17.45 ± 0.53
SN 1962L
0.056 ± 0.009 −18.08 ± 0.45
SN 1983I
SN 1983V
0.088 ± 0.014 −16.69 ± 0.36
SN 1987M 0.110 ± 0.018 −16.67 ± 0.52
0.097 ± 0.016 −16.27 ± 0.52
SN 1991N
0.150 ± 0.024 −15.98 ± 0.34
SN 1994I
SN 1990B
0.141 ± 0.023 −14.17 ± 0.67
SN 2004aw 0.089 ± 0.014 −16.49 ± 0.19
SN 1998bw 0.253 ± 0.040 −19.09 ± 0.62
SN 2002ap
0.301 ± 0.048 −16.49 ± 0.49
a MSFD
B
is absolute magnitude uncorrected for extinction in the parent galaxy.
TABLE 5
ABSOLUTE MAGNITUDE MSFD
V
mV
Supernova
SN 1972R
13.4 ± 0.3
SN 1983N
11.3 ± 0.2
SN 1984I
15.98 ± 0.20
SN 1984L
13.8 ± 0.2
SN 2000H
17.30 ± 0.03
SN 1999ex
16.63 ± 0.04
SN 1962L
13.13 ± 0.10
SN 1983I
13.6 ± 0.3
SN 1983V
13.80 ± 0.20
SN 1987M
14.7 ± 0.3
SN 1991N
13.9 ± 0.3
SN 1994I
12.91 ± 0.02
SN 1990B
15.75 ± 0.20
SN 2004aw 17.30 ± 0.03
SN 1998bw 13.75 ± 0.10
SN 2002ap
12.37 ± 0.04
ASFD
G (V)
MSFD
V
a
0.052 ± 0.008 −17.39 ± 0.54
0.228 ± 0.037 −17.18 ± 0.39
0.344 ± 0.055 −18.00 ± 0.46
0.091 ± 0.015 −17.97 ± 0.42
0.760 ± 0.122 −17.45 ± 0.32
0.067 ± 0.011 −16.88 ± 0.27
0.130 ± 0.021 −18.22 ± 0.32
0.043 ± 0.007 −18.16 ± 0.45
0.068 ± 0.011 −17.54 ± 0.26
0.085 ± 0.014 −17.25 ± 0.51
0.075 ± 0.012 −17.85 ± 0.51
0.115 ± 0.018 −16.81 ± 0.34
0.108 ± 0.017 −15.28 ± 0.27
0.069 ± 0.011 −17.23 ± 0.18
0.194 ± 0.031 −19.37 ± 0.41
0.161 ± 0.026 −17.09 ± 0.21
a MSFD
V
is absolute magnitude uncorrected for extinction in the parent galaxy.
|
astro-ph/0406591 | 1 | 0406 | 2004-06-25T19:54:01 | The Great Observatories Origins Deep Survey - VLT/FORS2 Spectroscopy in the GOODS-South Field | [
"astro-ph"
] | We present the first results of the ESO/GOODS program of spectroscopy of faint galaxies in the Chandra Deep Field South (CDF-S). 399 spectra of 303 unique targets have been obtained in service mode with the FORS2 spectrograph at the ESO/VLT, providing 234 redshift determinations (the median of the redshift distribution is at 1.04). The typical redshift uncertainty is estimated to be sig(z) ~ 0.001. Galaxies have been color selected in a way that the resulting redshift distribution typically spans from z=0.5 to 2. The reduced spectra and the derived redshifts are released to the community through the ESO web page http://www.eso.org/science/goods/ Large scale structure is clearly detected at z ~ 0.67, 0.73, 1.10 and 1.61. Three Lyman-break galaxies have also been included as targets and are confirmed to have redshifts z=4.800, 4.882 and 5.828. In a few cases, we observe clear [OII]3727 rotation curves, even at the relatively low resolution (R = 860) of the present observations. Assuming that the observed velocity structure is due to dynamically-relaxed rotation, this is an indication of large galactic masses (few times 10^(11) solar masses) at z ~ 1. | astro-ph | astro-ph | Astronomy&Astrophysicsmanuscript no. Vanzella
(DOI: will be inserted by hand later)
September 3, 2018
4
0
0
2
n
u
J
5
2
1
v
1
9
5
6
0
4
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
The Great Observatories Origins Deep Survey
VLT/FORS2 Spectroscopy in the GOODS-South Field
E. Vanzella1,2, S. Cristiani2, M. Dickinson3, H. Kuntschner4, L. A. Moustakas5, M. Nonino2, P. Rosati6, D. Stern8,
C. Cesarsky6, S. Ettori6, H. C. Ferguson5, R.A.E. Fosbury4, M. Giavalisco5, J. Haase4, A. Renzini6, A. Rettura6,7,
P. Serra4, and the GOODS Team
1 Dipartimento di Astronomia dell'Universit`a di Padova, Vicolo dell'Osservatorio 2, I-35122 Padova, Italy.
2 INAF - Osservatorio Astronomico di Trieste, Via G.B. Tiepolo 11, 40131 Trieste, Italy.
3 National Optical Astronomy Obs., P.O. Box 26732, Tucson, AZ 85726.
4 ST-ECF, Karl-Schwarzschild Str. 2, 85748 Garching, Germany.
5 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218.
6 European Southern Observatory, Karl-Schwarzschild-Strasse 2, Garching, D-85748, Germany.
7 Universite' Paris-Sud 11, Rue Georges Clemenceau 15, Orsay, F-91405, France
8 Jet Propulsion Laboratory, California Institute of Technology, MS 169-506, 4800 Oak Grove Drive, Pasadena, CA 91109 ⋆
Received ; accepted
Abstract. We present the first results of the ESO/GOODS program of spectroscopy of faint galaxies in the Chandra Deep
Field South (CDF-S). 399 spectra of 303 unique targets have been obtained in service mode with the FORS2 spectrograph at
the ESO/VLT, providing 234 redshift determinations (the median of the redshift distribution is at 1.04). The typical redshift
uncertainty is estimated to be σz ≃ 0.001. Galaxies have been color selected in a way that the resulting redshift distribution
typically spans from z=0.5 to 2. The reduced spectra and the derived redshifts are released to the community through the ESO
web page http : //www.eso.org/science/goods/. Large scale structure is clearly detected at z ≃ 0.67, 0.73, 1.10 and 1.61. Three
Lyman-break galaxies have also been included as targets and are confirmed to have redshifts z = 4.800, 4.882 and 5.828.
In a few cases, we observe clear [OII]3727 rotation curves, even at the relatively low resolution (ℜ = 860) of the present
observations. Assuming that the observed velocity structure is due to dynamically-relaxed rotation, this is an indication of large
galactic masses (few times 1011 M⊙) at z ∼ 1.
Key words. Cosmology: observations -- Cosmology: deep redshift surveys -- Cosmology: large scale structure of the universe
-- Galaxies: evolution.
1. Introduction
The Great Observatories Origins Deep Survey (GOODS) is a public, multifacility project that aims to answer some of the
most profound questions in cosmology: how did galaxies form and assemble their stellar mass? When was the morphological
differentiation of galaxies established and how did the Hubble Sequence form? How did AGN form and evolve, and what role do
they play in galaxy evolution? How much do galaxies and AGN contribute to the extragalactic background light? Is the expansion
of the universe dominated by a cosmological constant? A project of this scope requires large and coordinated efforts from many
facilities, pushed to their limits, to collect a database of sufficient quality and size for the task at hand. It also requires that the
data be readily available to the worldwide community for independent analysis, verification, and follow-up.
The program targets two carefully selected fields, the Hubble Deep Field North (HDF-N) and the Chandra Deep Field South
(CDF-S), with three NASA Great Observatories (HST, Spitzer and Chandra), ESA's XMM-Newton, and a wide variety of ground-
based facilities. The area common to all the observing programs is 320 arcmin2, equally divided between the North and South
fields. For an overview of GOODS, see Dickinson et al. (2003), Renzini et al. (2002) and Giavalisco et al. (2004a).
Spectroscopy is essential to reach the scientific goals of GOODS. Reliable redshifts provide the time coordinate needed to
delineate the evolution of galaxy masses, morphologies, clustering, and star formation. They calibrate the photometric redshifts
Send offprint requests to: E. Vanzella ([email protected])
⋆ Based on observations made at the European Southern Observatory, Paranal, Chile (ESO programme 170.A-0788 The Great Observatories
Origins Deep Survey: ESO Public Observations of the SIRTF Legacy/HST Treasury/Chandra Deep Field South.)
2
E. Vanzella et al.: The Great Observatories Origins Deep Survey
that can be derived from the imaging data at 0.36-8µm. Spectroscopy will measure physical diagnostics for galaxies in the
GOODS field (e.g., emission line strengths and ratios to trace star formation, AGN activity, ionization, and chemical abundance;
absorption lines and break amplitudes that are related to the stellar population ages). Precise redshifts are also indispensable to
properly plan for future follow-up at higher dispersion, e.g., to study galaxy kinematics or detailed spectral-line properties.
The ESO/GOODS spectroscopic program is designed to observe all galaxies for which VLT optical spectroscopy is likely
to yield useful data. The program makes full use of the VLT instrument capabilities (FORS2 and VIMOS), matching targets
to instrument and disperser combinations in order to maximize the effectiveness of the observations. The magnitude limits and
selection bandpasses depend to some degree on the instrumental setup being used. The aim is to reach mag ∼ 24 − 25 with
adequate S/N, with this limiting magnitude being in the B band for objects observed with the VIMOS LR-Blue grism, in the V
band for those observed in the VIMOS LR-Red grism, and in the z band for the objects observed with FORS2. This is not only a
practical limit, however, but is also well matched to the scientific aims of the GOODS program. The ACS i775 imaging samples
B. This is also the practical limit
rest-frame optical (B-band) light out to z = 1, where i775 = 25 reaches 1.5 to 2 magnitudes past L∗
for high-quality, quantitative morphological measurements from the ACS images (cf. Abraham et al. 1996). Similarly, i775 = 25
is ∼ 1 mag fainter than the measured L∗ UV for z = 3 Lyman Break Galaxies (LBGs), and 0.5 mag fainter than that at z = 4
(Steidel et al. 1999). These are the limits to which GOODS/SIRTF IRAC data will robustly measure rest-frame near-IR light, and
hence constrain the stellar mass.
In this paper we report on the first spectroscopic follow-up campaign in the Chandra Deep Field South (CDF-S), carried out
with the FORS2 instrument at the ESO VLT in the period fall 2002 - spring 2003 (the first 9 masks, 348 slits). Further 17 masks
have been observed during the period 2003 and early 2004, for which the reduction process has started and will be presented
elsewhere (Vanzella et al., in preparation).
The paper is organized as follows: in Sect. 2 we describe the target selection and in Sect. 3 the observations and the reduc-
tion. The redshift determination is presented in Sect. 4. In Sect. 5 we discuss the data and in Sect. 6 the conclusions are pre-
sented. Throughout this paper the magnitudes are given in the AB system (AB ≡ 31.4 − 2.5 logh fν/nJyi), and the ACS F435W,
F606W, F775W, and F850LP filters are denoted hereafter as B435, V606, i775 and z850, respectively. We assume a cosmology with
Ωtot, ΩM, ΩΛ = 1.0, 0.3, 0.7 and H0 = 70 km s−1 Mpc−1.
2. Target Selection
Objects were selected as candidates for FORS2 observations primarily based on the expectation that the detection and measure-
ment of their spectral features would benefit from the high throughput and spectral resolution of FORS2, and its reduced fringing
at red wavelengths, relative to other instrumental options such as VIMOS. In particular, we expect that the main spectral emission
and absorption features for galaxies at 0.8 < z < 1.6 would appear at very red optical wavelengths, out to ∼ 1µm. Similarly, very
faint Lyman break galaxies at z & 4, selected as B435, V606 and i775 -- dropouts from the GOODS ACS photometry, also benefit
greatly from the red throughput and higher spectral resolution of FORS2.
In practice, several categories of object selection criteria were used to ensure a sufficiently high density of target candidates
on the sky to efficiently fill out multi-slit masks. Using ACS photometry in the AB magnitude system, these criteria were:
1. Primary catalog: (i775 − z850) > 0.6 and z850 < 24.5. This should ensure redshifts z & 0.7 for ordinary early-type galaxies
(whose strongest features are expected to be absorption lines), and higher redshifts for intrinsically bluer galaxies likely to
have emission lines.
2. Secondary catalog: 0.45 < (i775 − z850) < 0.6 and z850 < 24.5.
3. Photometric-redshift sample: 1 < zphot < 2 and z850 < 24.5, using an early version of GOODS photometric redshifts like
those described by Mobasher et al. 2004.
4. i775-dropout and V606 -- dropout Lyman break galaxy candidates, selected from the criteria of Dickinson et al. 2004a and
Giavalisco et al. 2004b, respectively.
5. A few miscellaneous objects, including host galaxies of supernovae detected in the GOODS ACS observing campaign.
Target selection and mask design for the 2002-3 GOODS/FORS2 campaign was carried out while the GOODS ACS obser-
vations were still in progress, and before final ACS data reduction or cataloging could be completed. The targets were therefore
selected based on interim data reductions and catalogs, initially based on only one epoch of ACS imaging, and later using the
three-epoch ACS stacks and preliminary catalogs described in Giavalisco et al. (2004a). Because of this, the actual magnitudes
and colors of the observed galaxies from the final, 5-epoch ACS image stack, which we report here in Table 2, may not exactly
match the intent of the original selection criteria. When designing the masks, we generally tried to avoid observing targets that
had already been observed in other redshift surveys of this field, namely, the K20 survey of Cimatti et al. (2002) and the survey
of X-ray sources by Szokoly et al. 2004.
In the present spectroscopic catalog there are 303 targets, 114 meeting the primary selection criterion and 56 meeting the
secondary selection criteria. The other targets belong to the remaining classes.
Table 1. Journal of the MXU Observations.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
3
Mask ID
990247
984829
985831
973934
952426
981451
995131
994852
990652
UT date
30Dec.2002 - 2,6Jan. 2003
9Dec.2002 - 3,4Jan. 2003
5Jan. - 4,7Feb. 2003
7, 30, 31Jan. 2003
6,7Jan. 2003
31Jan. - 24,27Feb. - 22Nov. - 17Dec. - 30Jul. - 1Aug. 2003
5-6 Oct. 2002
4 Oct. 2002
8Dec. 12Nov. 2002
exp.time (s)
12×1200
12×1200
15×1200 + 663
12×1200
12×1200
24×1200
8×1800
8×1800
14×1200 + 300 + 900
3. Observations and Data Reduction
The VLT/FORS2 spectroscopic observations were carried out in service mode during several nights in 2002 and 2003. A summary
is presented in Table 1. In all cases the 300I grism was used as dispersing element without order-separating filter. This grism
provides a scale of roughly 3.2Å/pixel. The nominal resolution of the configuration was ℜ = λ/∆λ=860, which corresponds to
about 9Å at 8000Å. The spatial scale of FORS2 was 0.126′′/pixel, the slit width was always 1′′. Dithering of the targets along
the slits was applied in order to effectively improve the sky subtraction an the removal of CCD cosmetic defects. The mean shift
applied was ±8 pixels.
3.1. Data Reduction
Data were reduced with a semi-automatic pipeline that we have developed on the basis of the MIDAS package (Warmels 1991,
using commands of the LONG and MOS contexts (Fig. 1). The frames have been bias-subtracted and flat-fielded. For each slit the
sky background was estimated with a second order polynomial fitting. In some cases, better results have been obtained adopting
a first order polynomial. The fit has been computed independently in each column inside two windows, above and below the
position of the object (if more objects are present in the slit, a suitable modification of the windows is applied).
The resulting dithered, sky-subtracted, two-dimensional frames for each object are then averaged, with the weighting deter-
mined based on exposure time, seeing, and meteorlogical conditions. Spatial median filtering has been applied to each dithered
exposures to clean the cosmic rays. The FORS2 instrument shows an exquisite response in the red domain (beyond 8000Å), in
practice no appreciable residual fringe pattern affects the extracted signal. Rather, the sky residuals dominate the noise in the
regions where the intensity and the density of the skylines increase (see Figure 1). The individual dithered sky-subtracted spectra
have been visually inspected to verify that the object is indeed in the expected region of the slit. This step is necessary since the
applied small spatial offsets between the science exposures can result in objects falling too close to the slit edge or even outside
the slit (in exceptional cases). After this visual screening, the spatial offset between different exposures of the same object was
calculated on the basis of the world coordinate system (WCS) information stored in the frame headers. The individual exposures
were co-added (including the rejection of bad pixels or cosmic ray hits) after applying these spatial shifts. The frames were
shifted in the spatial direction and only by integer numbers of pixels. As the objects were sufficiently well sampled (the pixel
scale was significantly smaller than the seeing), no significant blurring of the spectra was observed, while the statistical properties
of the individual pixels were preserved.
The position of the target on the detector was estimated by collapsing 700 columns in the dispersion direction and measuring
the center of the resulting profile. The 1-D object signal was obtained using the 'optimal extraction' method of MIDAS. This
procedure calculates a weighted average in each column, based on both the estimated object profile and photon statistics.
Wavelength calibration was calculated on (daytime) arc calibration frames, using three arc lamps (a He, and two Ar lamps)
providing sharp emission lines over the whole spectral range used (6000 -- 10800Å). The object spectra were then rebinned to
a linear wavelength scale. We have verified the accuracy of the wavelength calibration by checking the position of 25 narrow
skylines in the science exposures (from 5577 to 10400 Å). Systematic translations of the wavelength scale have been typically
measured to be of the order of ±1Å and corrected. The final (absolute) mean wavelength accuracy for all spectra is 0.9±0.1Å
RMS.
Relative flux calibration was achieved by observations of standard stars listed by Bohlin et al. 1995. Since the standard stars
are typically quite blue and second order light can be substantial, the calibration spectra were obtained both with and without
order-sorting filters, providing calibration across the entire optical window. As noted previously, we opted to obtain the science
target without an order-sorting filter, implying deleterious effects to the flux calibration, particularly for bluer objects and at longer
wavelengths. For the red objects which dominated the FORS2 target selection, we felt that the improved wavelength coverage
4
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Fig. 1. Typical FORS2 data products for an individual slit of multi-object mask. From the top of the figure: the 2-D spectrum of
the arc lines used for the wavelength calibration, a 2-D science exposure (1200 seconds), the final flat-fielded and sky-subtracted
2-D spectrum (co-addition of 12 exposures for a total of 4 h), and at the bottom the 1-D spectrum with the identification of the
main absorption and emission lines (in this example an elliptical galaxy at z=1.100, GDS J033217.46-275234.8).
more than compensated for the slightly comprimised flux calibration. Due to both this second order light and uncertain slit losses,
we caution against using the calibrated fluxes for scientific purposes.
4. Redshift Determination
Spectra of 399 objects have been extracted. From them we have been able to determine 234 redshifts. In the large majority of
the cases the redshift has been determined through the identification of prominent features of galaxy spectra: the 4000Å break,
Ca H and K, g-band, MgII 2798, AlII 3584 in absorption and Lyα, [O ]3727, [O ]5007, Hβ, Hα in emission. The redshift
estimation has been performed cross-correlating the observed spectrum with templates of different spectral types (S0, Sa, Sb,
Sc, Ell., Lyman Break, etc.), using the rvsao package in the IRAF environment. The redshift identifications are summarized in
Table 2 and are available at the URL http : //www.eso.org/science/goods/.
In Table 2, the column ID contains the target identifier, that is constructed out of the target position (e.g., GDS J033206.44-
274728.8) where GDS stands for GOODS South. The quality flag, indicates the reliability of the redshift determination. Quality
"A" indicates a solid redshift determination, "B" a likely redshift determination, "C" a tentative redshift determination and "X"
an inconclusive spectrum or three cases in which no extraction was possible. 150 objects have been classified with quality "A",
57 with quality "B", 27 with quality "C", 69 with inconclusive redshift determination "X".
The class flag groups the objects for which emission line(s) (em.), absorption-line(s) (abs.) or both (comp.) are detected
in the spectrum. The classification has been guided by the observed continuum level and slope blueward and redward of the
emission/absorption feature, by the broad-band colors and the morphology of the targets (see Figure 2). 11 objects have been
classified as stars.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
5
Fig. 2. Three examples of objects classified as "em." (emission-lines detected), "abs." (absorption lines) and "comp." (both emis-
sion and absorption lines detected).
In 38% of the cases the redshift is based only on one emission line, usually identified with [O ]3727 or Lyα. In these cases the
continuum shape, the presence of breaks, the absence of other spectral features in the observed spectral range and the broad band
photometry are particularly important in the evaluation. In general these solo-emission line redshifts are classified as "likely" (B)
or "tentative" (C).
Finally, the comments column contains additional information relevant to the particular observation. The most common ones
summarize the identification of the principal lines, the inclination of an emission line due to internal kinematics, the weakness of
the signal ("faint"), the low S/N of the extracted spectrum ("noisy"), the 20% light radius ("Flux-radius") for objects classified
as stars, etc.
There are two objects that are not present in the v1.0 catalog, with z = 0.957 and z = 4.882 (marked with a cross in the
Table 2). These two objects were not successfully deblended in the detection process from the brighter nearby galaxies.
The internal redshift accuracy can be estimated from a sample of 42 galaxies which have been observed twice (or more)
in independent FORS2 mask sets. The distribution of measured redshift differences is presented in Figure 4. The mean of the
distribution is close to zero (10−6) and the redshift dispersion σz = 0.00078 and mean absolute deviation < ∆z >= 0.00055,
fairly constant with redshift. These values can be considered as a lower limit to the redshift uncertainty.
5. Discussion
5.1. Reliabilityoftheredshift-comparisonwithVVDS
A practical way to assess the reliability of the redshifts reported in Table 2 is to compare the present results with independent
measurements of other surveys. From this point of view the recent release of the data of the VIMOS-VLT Deep Survey (VVDS,
Le Fevre et al. 2004) is particularly important. There are 39 VVDS objects in common with the first release of the FORS2
GOODS survey and Figure 5 shows the comparison of the redshift determinations. The reliability level of the redshift measure-
6
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Table 2. Spectroscopic redshift catalog.
ID(V1)
GDS J033206.44-274728.8
GDS J033210.73-274819.4
GDS J033210.79-274719.8
GDS J033210.92-274722.8
GDS J033210.93-274721.5
GDS J033212.00-275104.2
GDS J033212.47-274621.4
GDS J033212.61-274605.1
GDS J033212.79-274823.1
GDS J033213.53-274917.0
GDS J033214.05-275124.5
GDS J033214.33-274825.2
GDS J033214.38-274825.9
GDS J033214.69-275258.2
GDS J033214.71-275257.2
GDS J033214.81-274600.0
GDS J033214.93-274659.8
GDS J033215.01-274633.4
GDS J033215.09-275130.7
GDS J033215.88-274723.1
GDS J033216.02-274750.0
GDS J033216.17-275241.4
GDS J033216.28-274955.5
GDS J033216.34-275013.4
GDS J033216.37-275201.3
GDS J033216.69-275239.0
GDS J033216.91-274808.3
GDS J033216.95-274519.3
GDS J033216.98-275102.4
GDS J033217.29-274807.5
GDS J033217.29-275113.2
GDS J033217.31-275025.0
GDS J033217.34-274844.3
GDS J033217.46-275234.8
GDS J033217.47-274838.4
GDS J033217.48-275248.0
GDS J033217.56-274709.2
GDS J033217.56-274810.1
GDS J033217.62-275228.5
GDS J033217.63-274811.8
GDS J033217.77-274603.0
GDS J033217.78-274823.8
GDS J033217.80-275256.9
GDS J033217.91-274122.7
GDS J033217.94-274721.5
GDS J033218.01-274718.5
GDS J033218.03-274850.3
GDS J033218.07-274845.7
GDS J033218.19-274746.6
GDS J033218.24-274744.0
GDS J033218.58-274619.0
GDS J033218.61-274705.1
GDS J033218.67-274915.7
GDS J033218.70-274919.8
GDS J033218.78-274951.3
GDS J033218.79-274820.8
GDS J033218.81-274908.5
GDS J033218.81-274910.0
GDS J033219.15-274040.2
GDS J033219.23-274545.5
GDS J033219.30-275219.3
z850
21.07
24.33
25.64
19.98
22.19
23.00
24.06
24.05
23.03
23.92
22.69
24.91
27.14
23.82
24.27
24.24
23.75
23.68
24.31
21.75
22.83
22.55
23.72
23.25
23.28
22.57
25.37
23.61
23.54
21.97
22.73
23.50
24.88
21.93
22.20
21.82
24.01
24.36
21.18
22.57
24.23
22.57
24.30
22.10
20.04
19.41
23.20
23.19
23.74
23.63
23.71
23.16
24.18
22.69
23.82
23.41
23.90
23.20
21.11
23.44
22.00
(i775 − z850)
0.70
0.96
0.06
0.29
1.00
0.85
1.03
0.53
0.45
0.45
0.74
-0.012
0.24
0.24
0.48
0.54
0.12
0.45
0.32
0.28
0.96
0.78
0.41
0.52
0.11
0.88
0.37
0.78
0.40
0.45
0.85
0.19
0.35
0.96
0.18
1.08
0.43
0.13
0.78
0.52
0.91
0.08
0.50
0.96
0.53
0.61
0.18
0.75
1.49
0.39
0.56
0.41
0.28
0.77
0.44
0.30
0.43
0.34
1.14
1.35
1.14
zspec
1.022
1.396
-
0.417
1.222
1.018
-
1.378
1.316
-
1.220
-
-
1.101
1.360
1.370
-
1.000
1.229
0.896
1.298
1.094
-
class.
comp.
em.
-
em.
abs.
comp.
-
em.
em.
-
em.
-
-
em.
em.
em.
-
abs.
em.
em.
comp.
abs.
-
1.046
comp.
-
1.045
-
1.303
0.991
0.735
-
1.612
1.107
1.100
0.737
1.095
-
0.542
1.098
0.735
-
0.117
1.044
1.041
0.732
0.735
0.297
0.000
0.000
-
1.435
1.380
-
1.038
1.294
0.999
1.128
0.735
1.128
0.000
1.096
-
abs.
-
em.
em.
abs.
-
em.
em.
abs.
em.
abs.
-
em.
comp.
em.
-
em.
em.
abs.
abs.
comp.
em.
star
star
-
em.
em.
-
comp.
em.
em.
em.
comp.
abs.
star
abs.
Quality
A
C
X
B
A
A
X
A
A
X
A
X
X
A
B
C
X
C
A
A
B
B
X
A
X
A
X
A
A
A
X
A
C
A
A
A
X
C
A
A
X
B
B
A
A
A
A
B
B
X
A
A
X
A
A
B
B
A
A
C
A
comments
[OII], CaHK, MgI
faint, [OII]
faint(line@8069A?)
CaH, g-band, [OIII], Na, Hα
CaHK, MgI, Hδ
[OII], CaHK, g-band
faint
[OII], MgI
[OII], MgII, (CaHK faint)
faint
[OII], MgII
featureless continuum
faint
[OII]
[OII]?
[OII]?
abs@7080A
D4000 break?
[OII]
[OII], Hβ, TILT
CaHK, [OII](line6255?)
CaHK, g-band, MgI-noisy
featureless continuum
[OII](Sky-ABS), CaHK
bright, abs@6271,6982,7100
CaHK, g-band
faint
[OII]
[OII], [OIII], Hβ
CaHK
bad-row, featureless?
[OII]
[OII], faint
CaHK, MgI
[OII], [OIII], Hβ
CaHK, g-band
faint
[OII]?-faint
CaHK, MgI, g-band, AlII, [OII]
[OII], CaHK, [OIII]
faint
Hα,[OIII]
[OII]?(SKY.ABS)
CaHK, MgI
CaHK, g-band, MgI, Hδ, Hβ, AlII
[OII], CaHK(noisy)
[OIII], Hβ, Hα
star Flux-radius = 1.261
star Flux-radius = 1.256
featureless continuum
[OII]
[OII], MgII
faint,line@8200A?
[OII], CaHK
[OII], noisy
[OII]
[OII]?
[OII], CaHK
CaHK, MgI, g-band, bright
star? Flux-radius = 1.239
CaHK, MgI, g-band
Table 2. Spectroscopic redshift catalog.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
7
ID(V1)
GDS J033219.43-274928.2
GDS J033219.48-274216.8
GDS J033219.61-274831.0
GDS J033219.68-275023.6
GDS J033219.77-274204.0
GDS J033219.79-274609.9
GDS J033219.79-274839.3
GDS J033219.89-274517.8
GDS J033219.96-274449.8
GDS J033219.97-274547.6
GDS J033219.99-274443.2
GDS J033220.02-274104.2
GDS J033220.11-275329.8
GDS J033220.28-275233.0
GDS J033220.29-274718.2
GDS J033220.41-274641.7
GDS J033220.72-274932.6
GDS J033220.91-275344.0
GDS J033221.22-274625.9
GDS J033221.57-274941.6
GDS J033221.63-274800.2
GDS J033221.67-274056.0
GDS J033221.76-274442.1
GDS J033221.81-274352.3
GDS J033221.84-274434.4
GDS J033221.99-274655.9
GDS J033222.18-274659.7
GDS J033222.36-275018.4
GDS J033222.41-274858.0
GDS J033222.47-275047.4
GDS J033222.54-274603.8
GDS J033222.58-274425.8
GDS J033222.93-274919.1
GDS J033222.93-275104.6
GDS J033223.17-274219.6
GDS J033223.18-274921.5
GDS J033223.26-275101.8
GDS J033223.28-274744.7
GDS J033223.29-274742.6
GDS J033223.40-274316.6
GDS J033223.45-274709.0
GDS J033223.61-274601.0
GDS J033223.61-275306.3
GDS J033223.69-275324.4
GDS J033223.83-274639.4
GDS J033223.90-275326.2
GDS J033224.01-275039.0
GDS J033224.08-275214.6
GDS J033224.11-274102.1
GDS J033224.20-274257.5
GDS J033224.20-274952.9
GDS J033224.26-274126.4
GDS J033224.37-274315.2
GDS J033224.39-274624.3
GDS J033224.66-275051.9
GDS J033224.72-274120.4
GDS J033224.79-274912.9
GDS J033224.85-275052.6
GDS J033224.90-274715.0
GDS J033224.91-274923.7
z850
23.90
20.29
21.87
20.50
23.34
24.11
23.55
24.00
22.35
23.94
24.43
21.51
24.44
22.22
23.94
24.15
24.16
22.99
23.48
23.01
24.35
22.75
20.86
24.27
24.81
20.42
25.00
22.82
24.19
24.38
24.32
20.28
24.77
22.89
23.82
24.08
21.90
23.94
24.31
20.48
23.22
22.57
22.30
20.41
24.19
23.81
22.74
23.56
23.36
24.15
23.61
20.18
24.63
21.92
22.40
21.17
24.90
23.24
24.65
23.87
(i775 − z850)
0.58
0.31
0.10
0.25
0.86
0.88
0.52
0.15
0.24
0.47
0.23
0.52
0.75
0.92
0.32
0.35
0.80
0.71
0.50
0.61
0.08
0.65
0.18
0.58
0.36
0.47
1.04
0.27
0.46
1.79
0.25
0.33
0.51
0.38
0.78
0.55
0.85
0.24
0.41
0.33
0.49
1.06
0.85
0.40
0.84
1.07
0.98
0.92
0.88
0.32
0.21
0.44
-0.06
0.76
-0.28
0.80
1.61
0.61
0.71
0.08
zspec
1.048
0.382
0.671
0.559
1.044
1.221
1.357
-
0.783
1.219
-
0.682
1.385
1.119
-
1.227
-
1.044
1.221
1.110
-
1.045
0.295
1.308
-
class.
em.
comp.
em.
comp.
comp.
em.
em.
-
em.
em.
-
abs.
em.
abs.
-
em.
-
em.
em.
em.
-
em.
em.
em.
-
0.670
comp.
-
0.736
1.383
0.000
-
0.738
1.298
0.905
-
1.109
0.964
0.764
1.092
0.615
1.423
1.033
1.125
0.532
1.222
-
1.094
1.015
0.000
-
-
0.533
1.271
0.895
0.272
0.967
0.000
1.329
-
-
-
em.
em.
star
-
comp.
em.
em.
-
em.
abs.
em.
em.
comp.
em.
em.
abs.
comp.
em.
-
abs.
em.
star
-
-
comp.
em.
abs.
em.
abs.
star
em.
-
-
Quality
B
A
A
A
A
A
A
X
C
A
X
A
C
A
X
A
X
B
A
A
X
B
A
B
X
A
X
B
A
C
X
A
B
A
X
B
A
B
A
A
A
A
C
A
B
X
A
C
A
X
X
A
C
A
C
A
C
A
X
X
comments
[OII](SKY.ABS)
CaHK, low-z
[OII], [OIII], Hβ
[OII], CaHK, [OIII], Hα
[OII], CaHK
[OII], MgI
[OII]
bright, abs@8176,6716Å
[OII]?
[OII]
faint
CaHK, g-band, MgI
[OII]?
CaHK, MgI, g-band
faint, noisy
[OII]
faint
[OII](SKY.ABS), CaHK
[OII]
[OII]
featureless continuum
[OII](SKY.ABS), CaHK
Hβ, [OIII], Hα
[OII]
faint
[OII], CaHK, g-band
faint
[OII]
[OII], MgI, MgII
star? Flux-radius = 1.307
faint(line@9400Å?)
[OII], Hβ, CaHK,MgI
[OII]?
[OII], Hβ
faint(abs@8157,9200)
[OII]
CaHK, g-band
[OII]
[OII], CaK
[OII], CaHK, g-band, Hβ, [OIII]
[OII], MgII
[OII], red
CaHK? (noisy)
[OIII], Hα
[OII]
faint
CaHK,MgI
[OII]?, g-band?
star Flux-radius = 1.247
featureless continuum
diffuse, faint
[OII], [OIII], Hβ, g-band
[OII]?
CaHK, [OII] faint, (short-slit)
Hα, Mg?
CaHK, g-band, MgI, AlII
continuum+break, Flux-radius = 1.284
[OII]
faint
featureless bright continuum
8
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Table 2. Spectroscopic redshift catalog.
ID(V1)
GDS J033225.04-274718.2
GDS J033225.10-274219.5
GDS J033225.19-274735.3
GDS J033225.20-275009.4
GDS J033225.21-275335.0
GDS J033225.35-274502.8
GDS J033225.47-274327.6
GDS J033225.48-275211.6
GDS J033225.54-275209.1
GDS J033225.55-275108.2
GDS J033225.58-274529.0
GDS J033225.69-274347.1
GDS J033225.76-274347.0
GDS J033225.77-274247.7
GDS J033225.79-274352.3
GDS J033225.86-275019.7
GDS J033225.90-274341.2
GDS J033226.00-274150.6
GDS J033226.03-275147.7
GDS J033226.16-274946.5
GDS J033226.17-274603.6
GDS J033226.24-275005.6
GDS J033226.26-274209.6
GDS J033226.31-274722.4
GDS J033226.32-274232.3
GDS J033226.40-274228.2
GDS J033226.49-274035.5
GDS J033226.64-274028.2
GDS J033226.66-274025.1
GDS J033226.66-274029.8
GDS J033226.67-274758.8
GDS J033226.67-274834.8
GDS J033226.84-274545.3
GDS J033226.89-274541.9
GDS J033226.92-274239.8
GDS J033227.02-274407.2
GDS J033227.05-275318.4
GDS J033227.07-274404.7
GDS J033227.11-274922.0
GDS J033227.17-274957.8
GDS J033227.36-274204.8
GDS J033227.58-274051.7
GDS J033227.70-274043.7
GDS J033227.72-275040.8
GDS J033227.84-274136.8
GDS J033227.88-275140.4
GDS J033228.09-275202.4
GDS J033228.42-274700.2
GDS J033228.44-274703.7
GDS J033228.45-274419.3
GDS J033228.48-274059.6
GDS J033228.56-274055.7
GDS J033228.84-274132.7
GDS J033228.88-274129.3
GDS J033228.94-274600.6
GDS J033228.94-274128.2†
GDS J033228.99-274908.4
GDS J033229.07-274153.1
GDS J033229.22-274707.6
GDS J033229.32-274054.0
† not present in the catalog v1.0
z850
23.86
23.65
23.90
22.88
21.17
22.58
20.14
23.68
22.93
23.93
24.33
24.73
23.02
24.26
23.17
21.54
18.92
21.75
22.98
23.09
24.25
23.68
23.96
22.48
23.86
23.33
19.60
21.22
21.70
22.06
21.81
23.59
23.47
23.72
22.95
22.15
22.25
22.24
22.88
23.26
21.26
23.43
21.64
21.42
21.97
20.26
20.29
24.31
20.86
22.78
23.82
25.44
25.43
20.72
23.82
-
20.56
24.27
20.65
23.74
(i775 − z850)
0.51
1.09
0.60
0.95
0.72
0.16
0.51
0.39
0.45
0.30
0.15
1.01
1.11
0.62
0.58
0.48
0.88
0.23
0.53
0.52
0.54
0.73
0.53
0.23
0.24
0.42
0.19
0.21
0.29
0.89
0.20
0.55
0.52
0.13
0.37
0.84
0.70
0.16
-0.01
0.51
0.53
0.94
0.86
0.74
0.91
0.40
0.41
0.19
0.47
0.50
0.49
0.07
0.01
0.53
0.98
-
0.73
0.44
0.51
0.74
zspec
1.357
1.609
1.017
1.100
0.833
0.975
0.668
1.312
0.955
0.832
0.667
-
-
1.026
1.297
1.095
0.000
0.545
1.242
0.735
1.219
1.096
0.932
0.737
0.736
1.615
-
0.310
1.042
1.040
0.628
0.905
1.306
0.338
-
1.128
1.103
0.739
0.559
1.293
0.735
1.070
0.968
1.097
1.043
0.521
0.560
-
-
class.
em.
em.
em.
abs.
comp.
em.
abs.
em.
em.
em.
em.
-
-
em.
em.
em.
star
comp.
em.
comp.
em.
em.
em.
em.
em.
em.
-
em.
em.
abs.
em.
abs.
em.
em.
-
comp.
comp.
em.
em.
em.
comp.
em.
abs.
comp.
abs.
abs.
abs.
-
-
1.135
em.
-
-
4.800
0.733
-
4.882
1.095
-
0.668
-
-
-
em.
comp.
-
em.
abs.
-
abs.
-
Quality
A
B
B
B
A
A
A
B
B
A
A
X
X
B
A
A
C
A
A
A
B
B
B
A
B
A
X
A
C
A
C
C
A
A
X
A
A
A
A
A
A
C
A
A
A
A
A
X
X
A
X
X
C
A
X
B
A
X
A
X
comments
[OII], MgII
[OII]
[OII], faint
CaHK, AlII
[OII], [OIII], CaHK, Hδ, (noisy)
[OII], Hβ, [OIII]
CaHK, [OII], AlII
[OII]
[OII]?, noisy
[OII], [OIII], Hβ
[OII], [OIII]
faint
noisy, bad-row
[OII](SKY.ABS)
[OII], CaHK
bright [OII], TILT
star? Flux-radius = 1.296
[OII], Hβ, CaHK
[OII], MgI
[OII], CaHK
[OII]
[OII]
[OII]
[OII], Hβ, [OIII]
[OII]
[OII], MgII
faint
low-z, Hα, [OIII]
[OII]?(SKY.ABS)
CaHK, g-band, MgI
Hβ, [OIII]?
D4000break?
[OII]
[OII], Hβ, [OIII], Hα
featureless?
[OII], CaHK
[OII], CaHK
[OII], Hβ, [OIII]
[OII], Hβ, [OIII], Hα
[OII], CaHK
CaHK, MgI, AlII, g-band
faint, [OII]?
CaHK, g-band, AlII
[OII], CaHK
CaHK, g-band, MgI, AlII
CaHK, g-band
g-band, Na, Mg, [OIII]
faint, abs@7060
noisy
[OII], MgII
featureless continuum
faint
Lyα? No continuum
[OII], CaHK, MgI, g-band, Hδ, Hβ
abs@8555,8150,9200?
Lyα, (SiIV?)
CaHK, MgI, g-band
faint
CaHK, g-band
short-slit
Table 2. Spectroscopic redshift catalog.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
9
ID(V1)
GDS J033229.35-275048.5
GDS J033229.48-274036.7
GDS J033229.63-274511.3
GDS J033229.65-274524.7
GDS J033229.71-274507.2
GDS J033229.75-275147.1
GDS J033229.85-274520.5
GDS J033229.87-274317.7
GDS J033229.99-274322.6
GDS J033230.03-275026.8
GDS J033230.06-274523.5
GDS J033230.07-274319.0
GDS J033230.09-275100.3
GDS J033230.23-274519.9
GDS J033230.34-274523.6
GDS J033230.51-275004.4
GDS J033230.70-274928.7
GDS J033230.71-274617.2
GDS J033230.75-274306.9
GDS J033230.83-274931.8
GDS J033230.85-274621.7
GDS J033230.98-274434.9
GDS J033231.04-274050.2
GDS J033231.22-274052.2
GDS J033231.22-274532.7
GDS J033231.28-274820.2
GDS J033231.42-274324.1
GDS J033231.45-274435.0
GDS J033231.55-275028.8
GDS J033231.65-274504.8
GDS J033232.04-274451.7
GDS J033232.08-274119.4
GDS J033232.08-274155.2
GDS J033232.12-274359.3
GDS J033232.14-274349.9
GDS J033232.32-274343.6
GDS J033232.33-274345.8
GDS J033232.58-275053.9
GDS J033232.73-274538.8
GDS J033232.73-275102.5
GDS J033232.94-274543.9†
GDS J033232.96-274545.7
GDS J033233.00-275030.2
GDS J033233.01-274829.4
GDS J033233.02-274547.4
GDS J033233.08-275123.9
GDS J033233.25-274117.4
GDS J033233.28-274236.0
GDS J033233.41-274230.5
GDS J033233.71-274210.2
GDS J033233.82-274410.0
GDS J033233.85-274600.2
GDS J033234.00-274412.1
GDS J033234.05-274937.8
GDS J033234.08-274222.3
GDS J033234.82-274835.5
GDS J033234.85-274640.4
GDS J033235.08-274615.7
GDS J033235.11-275009.0
† not present in the catalog v1.0
z850
20.10
24.04
24.07
24.20
22.24
23.25
21.01
22.73
24.42
23.43
21.81
21.59
20.80
23.14
21.92
23.85
22.69
22.24
23.25
23.19
21.60
23.71
21.74
22.86
22.98
23.10
23.01
17.90
23.96
23.16
21.59
22.59
23.00
24.66
23.12
22.80
23.95
21.61
21.69
20.78
-
19.84
20.85
22.80
21.13
21.39
24.06
24.55
23.37
23.71
21.11
23.82
23.70
22.91
23.91
22.94
22.70
23.11
24.17
(i775 − z850)
0.30
0.42
0.70
0.73
0.17
0.54
0.56
0.79
0.98
0.83
0.50
0.64
0.52
0.09
0.07
1.03
0.16
0.66
0.25
0.79
0.74
0.50
0.77
0.58
1.00
0.65
0.82
0.49
0.68
0.54
0.68
0.53
0.39
0.18
0.36
0.07
0.42
0.21
0.31
0.58
-
0.33
0.44
0.12
0.63
0.47
0.86
1.09
0.06
0.46
0.42
1.05
0.53
0.71
0.72
0.66
0.61
0.63
0.44
zspec
0.415
1.221
1.033
-
0.736
1.315
0.953
1.097
-
1.005
0.955
1.101
0.733
0.523
1.223
-
-
1.307
0.860
-
1.018
1.222
1.037
1.333
1.097
1.173
1.025
0.000
-
1.098
0.895
1.036
0.960
-
0.973
0.533
1.025
0.669
-
0.735
0.957
0.366
0.669
0.664
0.953
0.735
-
1.215
0.975
1.043
0.667
1.910
0.896
0.832
1.476
1.245
1.099
1.316
1.295
class.
abs.
em.
em.
-
em.
em.
em.
comp.
-
abs.
em.
comp.
abs.
em.
comp.
-
-
em.
em.
-
comp.
em.
abs.
em.
abs.
comp.
em.
star
-
em.
abs.
em.
abs.
-
em.
em.
em.
em.
-
abs.
em.
abs.
em.
em.
abs.
abs.
-
em.
em.
em.
abs.
abs.
em.
comp.
em.
em.
em.
em.
em.
Quality
A
C
A
X
A
A
A
B
X
B
A
B
A
A
A
X
X
A
A
X
A
A
A
B
A
A
B
A
X
A
A
B
C
X
A
C
B
B
X
A
A
A
A
A
B
A
X
B
B
B
A
B
B
A
B
A
A
A
A
comments
Hβ, Mg, Na
[OII]?
[OII], MgI, CaH
faint
[OII], Hβ, [OIII]
[OII]
[OII], AlII
CaHK, faint [OII]
faint
CaHK, MgII, g-band
[OII]
[OII], CaHK
CaHK, g-band, MgI, AlII, Hδ, Hβ
[OII]
faint [OII], CaHK, MgI
faint
faint
[OII], TILT
[OII], Hβ
abs@7150,9056? (noisy)
CaHK, [OII]
[OII]
CaHK, g-band, MgI
[OII], noisy
CaHK, g-band
[OII]
[OII](SKY.ABS)
star Flux-radius = 1.299
faint red
[OII], MgI, CaHK, Hδ, Hγ
CaHK
[OII](SKY.ABS)
abs@7000,6400,7800
faint (line at 6217A?)
[OII]
[OIII]
[OII]
[OII], Hβ
short-slit(line5900A)
CaHK, g-band
[OII]
N[II]+abs.spec.
[OII], [OIII], Hβ
[OIII], Hβ
CaHK, [OII]
CaHK
continuum, faint
[OII], faint red (XCDFS265)
[OII]
[OII](SKY.ABS), TILT
CaHK, g-band
MgII
[OII], CaHK?
[OII], CaHK
[OII]
[OII], CaHK
[OII]
[OII]
[OII]
10
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Table 2. Spectroscopic redshift catalog.
ID(V1)
GDS J033235.19-275103.4
GDS J033235.26-275104.8
GDS J033235.78-274627.5
GDS J033235.79-274734.7
GDS J033236.04-275004.3
GDS J033236.39-274747.0
GDS J033236.43-274750.6
GDS J033237.19-274608.1
GDS J033237.26-274610.3
GDS J033237.56-274646.7
GDS J033238.27-274604.0
GDS J033238.49-274702.4
GDS J033239.01-274722.7
GDS J033239.35-275016.3
GDS J033239.56-274851.7
GDS J033239.60-274909.6
GDS J033239.64-274709.1
GDS J033239.67-274850.6
GDS J033239.99-275114.2
GDS J033240.01-274815.0
GDS J033240.67-275032.3
GDS J033240.79-275035.1
GDS J033240.92-274823.8
GDS J033241.21-274932.4
GDS J033241.41-274457.5
GDS J033241.48-274440.4
GDS J033241.59-275003.0
GDS J033241.67-274448.7
GDS J033241.76-274619.4
GDS J033242.07-274911.6
GDS J033242.21-274953.9
GDS J033242.25-274625.4
GDS J033242.32-274950.3
GDS J033242.38-274707.6
GDS J033242.56-274550.2
GDS J033242.97-274649.9
GDS J033244.18-274729.4
GDS J033244.20-274733.5
GDS J033244.23-275039.5
GDS J033244.29-275009.7
GDS J033244.43-274641.8
GDS J033244.62-274632.2
GDS J033244.80-274920.6
GDS J033245.15-274940.0
GDS J033245.21-274858.0
GDS J033245.90-274517.2
GDS J033247.45-274603.9
GDS J033248.56-274504.6
GDS J033249.04-275015.5
GDS J033249.09-274519.2
GDS J033249.11-274524.2
GDS J033249.49-274534.2
GDS J033249.85-274757.8
GDS J033250.69-274732.2
GDS J033251.34-274742.7
GDS J033251.57-275044.7
GDS J033252.87-275114.7
GDS J033252.88-275119.8
GDS J033253.01-275000.5
GDS J033253.34-275104.6
GDS J033255.00-275051.6
z850
24.41
22.79
22.76
23.65
23.40
23.18
22.41
20.90
22.17
24.89
24.46
21.26
24.40
22.05
22.55
20.72
22.70
24.55
23.75
25.33
24.56
21.69
24.00
24.26
23.02
23.00
22.32
23.83
20.93
23.42
23.83
22.50
20.33
23.18
22.21
23.44
23.55
21.53
24.72
22.62
22.55
23.20
23.69
21.92
24.65
23.79
23.98
22.19
22.36
21.46
21.81
23.94
22.78
23.51
24.09
22.48
21.20
21.84
24.05
25.50
21.70
(i775 − z850)
0.23
0.44
0.89
0.88
0.81
0.17
0.13
1.07
0.19
0.78
1.07
0.76
0.40
0.42
0.81
0.89
0.99
0.21
0.62
1.44
0.21
0.12
0.45
0.94
0.80
0.84
0.66
0.14
0.35
1.30
0.37
0.63
0.29
1.11
0.16
0.89
0.44
0.20
0.40
0.86
0.13
0.44
0.84
1.04
0.51
0.39
0.51
0.88
0.80
0.37
0.60
0.17
0.67
-0.02
0.76
0.52
0.63
0.47
0.09
-0.08
0.37
zspec
0.981
0.734
1.094
1.223
1.612
-
0.127
1.096
0.736
-
-
0.953
-
-
0.000
0.980
1.317
-
-
5.828
-
0.213
1.244
-
-
1.296
0.000
-
0.333
0.000
1.377
1.288
-
1.314
0.218
-
1.220
0.737
1.122
1.038
0.215
1.426
-
1.123
1.463
1.036
-
1.115
1.122
-
1.094
1.609
1.146
-
1.298
0.980
1.002
1.220
-
0.912
-
class.
em.
abs.
comp.
em.
em.
-
em.
abs.
comp.
-
-
abs.
-
-
star
abs.
comp.
-
-
em.
-
em.
em.
-
em.
em.
star
-
em.
star
em.
em.
-
abs.
em.
-
em.
em.
em.
abs.
em.
em.
-
abs.
em.
em.
-
abs.
em.
-
em.
em.
em.
-
em.
comp.
comp.
em.
-
em.
-
Quality
B
A
A
A
A
X
B
A
A
X
X
A
X
X
C
A
A
X
X
A
X
A
B
X
X
B
B
X
C
B
A
A
X
B
A
X
A
A
B
A
C
A
X
A
A
A
X
C
B
X
A
A
B
X
B
A
A
A
X
C
X
comments
[OII]?
CaHK, Hδ, MgI
[OII], CaHK
[OII], [NeIII]
[OII], MgII
abs@6500,6586
Hα, [OI]6300Å, Na
CaHK
[OII], CaHK
abs@8195, em@8949?
faint
CaHK, MgI
faint
smoothly-red
star Flux-radius = 1.269
CaHK, g-band, AlII
[OII], CaHK, MgI
featureless continuum
lines:8210,8800?
Lyα
bad-row, faint
Hα, S[II], (2d-order-light)
[OII]?
faint
em.lines@6815,8208
[OII], faint
compact, Flux-radius = 1.279
faint, short-slit
low-z, Hα
star Flux-radius = 1.253
[OII], MgI
[OII], MgII
noisy
[OII], CaHK
[OIII], Hβ, Hα
faint
[OII]
[OII], [OIII], Hβ
[OII]?
CaHK, g-band
Hα?
[OII], MgII
faint
CaHK, g-band, MgI
[OII]
[OII], CaHK
faint
CaHK, noisy
[OII]
featureless continuum
[OII], Hβ-faint
[OII]
[OII]?
em@7100A, abs8100A
[OII], noisy
[OII], Hβ
CaHK, [OII]
[OII], CaHK
faint
noisy, [OII]?
featureless continuum
E. Vanzella et al.: The Great Observatories Origins Deep Survey
11
Fig. 3. Color-magnitude diagram for the spectroscopic sample as a function of the quality flag. The uncertainties in the redshift
determination increase with increasing z850 magnitude. Few bright sources (often serendipitous; z850 < 22) have inconclusive
redshift determinations due to the dithering procedure, which has positioned these sources off the slitlets for many of the expo-
sures.
ments in the VVDS is indicated by a quality flag. Flags 2, 3, 4 are the most secure with a confidence of 75%, 95% and 100%
respectively. Flag 1 is an indicative measurement, flag 9 indicates that there is only one secure emission line, and flag 0 indicates
a measurement failure with no features identified.
For 29 cases out of 39 (74%) the agreement is very good, with a mean difference < zFORS 2 − zVVDS > = 0.0016 ± 0.0021.
Assuming equipartition of the redshift uncertainties between FORS2 and VVDS, we can estimate a σz(FORS 2) ≃ 0.0015,
in reasonable agreement with the estimate of Sect. 4. In the following we will assume a typical uncertainty of the redshift
determinations of the present survey to be σz ≃ 0.001 (excluding "catastrophic" discrepancies).
Ten cases show "catastrophic" discrepancies, i.e. < zFORS 2 − zVVDS > greater than 0.015 and are reported in Table 3.
In the following we discuss case by case the origin of the discrepancy:
12
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Fig. 4. Redshift differences between objects observed twice or more in independent FORS2 observations. The distribution has a
dispersion σz = 0.00078.
1. GDS J033214.05-275124.5:
-- FORS2: the emission line [O ]3727 and the absorption lines Ca H and K are detected in the FORS2 spectrum at z=1.220.
The absorption line MgII 2798 is also present at 6210Å. The 4000Å Balmer Break is also evident, quality flag "A".
-- VVDS: the main emission feature in the VIMOS spectrum is identified with [O ]3727 at z=1.325, quality flag 3. We
note an absorption feature in the VIMOS spectrum (without identification) at ∼6200Å, consistent with the one measured
in the FORS2 spectrum.
2. GDS J033219.79-274839.3:
-- FORS2: flat continuum with an evident emission line at ∼8784Å. We interpret it as [O ]3727. No spectroscopic feature
is observed at ∼ 6500Å. Quality flag A.
-- VVDS: the main feature in the VIMOS spectrum is identified with [O ]5007 at z=0.568 (emission line at 7851Å),
quality flag 2.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
13
Fig. 5. VVDS spectroscopic redshift versus FORS2 spectroscopic redshift. There are 39 galaxies in common between the VVDS
sample and the sample presented in this work. 10 cases show (filled symbols) discrepant redshift determination with dz >0.015.
3. GDS J033221.67-274056.0:
-- FORS2: the emission line [O ]3727 and the absorption lines Ca H and K are detected in the FORS2 spectrum at z = 1.045,
quality flag "B". The [O ]3727 line is attenuated by the sky absorption band at ∼7600Å.
-- VVDS: Ca H and K are identified in the VIMOS spectrum at z=0.977, quality flag 1.
4. GDS J033226.03-275147.7:
-- FORS2: the emission line [O ]3727 (at 8356Å) and the absorption lines Ca H and K, MgI and B2630 are detected in the
FORS2 spectrum at z = 1.242, quality flag "A".
-- VVDS: the main feature is identified with Hα at z=0.264 (at 8296Å), quality flag 9. No emission lines are present in the
FORS2 spectrum at this wavelength.
5. GDS J033231.65-274504.8:
14
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Table 3. Galaxies with discrepant redshifts between the FORS2 and VVDS surveys.
ID
z(FORS2) QF(FORS2)
z(VVDS) QF(VVDS)
N
GDS J033214.05-275124.5
1
GDS J033219.79-274839.3
2
GDS J033221.67-274056.0
3
GDS J033226.03-275147.7
4
GDS J033231.65-274504.8
5
GDS J033232.32-274343.6
6
GDS J033242.38-274707.6
7
GDS J033242.56-274550.2
8
9
GDS J033249.04-275015.5
10 GDS J033249.85-274757.8
† Adopted FORS2 value has been updated to VVDS value in Table 2.
A (em.)
A (em.)
B (em.)
A (em.)
A (em.)
C (em.)
B (abs.)
A (em.)
B (em.)
B (em.)
1.220
1.357
1.045
1.242
1.098
1.059†
1.314
0.218
1.122
1.146
1.325
0.568
0.978
0.264
1.443
0.533
0.688
0.635
1.072
1.254
3
2
1
9
1
3
2
2
2
2
z(FORS2)-z(VVDS)
-0.105
0.789
0.067
0.978
-0.350
0.526
0.626
-0.417
0.05
-0.105
-- FORS2: the emission line [O ]3727 and the absorption lines Ca H and K, MgI, Hδ and are detected in the FORS2
spectrum at z = 1.098; the emission line Hγ is also detected at ∼9105Å, quality flag "A".
-- VVDS: in the VIMOS spectrum a line is detected at ∼9105Å, interpreted as [O ]3727 at z=1.443, quality flag 1.
6. GDS J033232.32-274343.6:
-- FORS2: for this object (at the border of the FORS2 field of view) the spectrum starts at ∼6400Å. We detect a weak
emission line at ∼7654Å (close to a sky absorption band), that we originally interpreted to be [O ]3727 at z∼1.059,
assigning to the redshift a quality flag "C".
-- VVDS: in the VIMOS spectrum an emission line at ∼5713Å is detected, interpreted as [O ]3727 at z=0.533 and quality
flag 3 (the absorption feature Ca H at ∼6085Å is also present). It is consistent with the interpretation [O ]3727 at
z = 0.533 with the FORS2 ∼ 7654Å emission line identified as [O ]5007 at z = 0.533. We have therefore updated the
entry in Table 2 to a redshift z = 0.533.
7. GDS J033242.38-274707.6:
-- FORS2: this is a red object (i775 − z850 = 1.11), we detect two clear absorption features in the ∼ 9100Å sky free region
interpreted as Ca H and K, faint [O ]3727 seems to be present, quality flag "B".
-- VVDS: red spectrum, Ca H and K are identified in the VIMOS spectrum at z=0.688, quality flag 2.
8. GDS J033242.56-274550.2:
-- FORS2: the emission lines [O ]5007 (at 6098Å), Hβ (at 5921 Å) and Hα (at 7994Å) are detected in the FORS2 spectrum,
z=0.218, quality flag "A".
-- VVDS: the main emission feature (at 6094Å) is identified with [O ]3727 at z=0.635, quality flag 2.
9. GDS J033249.04-275015.5:
-- FORS2: the spectrum starts at ∼6400Å. It shows continuum with a evident emission line at ∼7909Å interpreted as
[O ]3727 (z=1.122), a discontinuity consistent with the 4000Å Balmer Break is present, quality flag "B".
-- VVDS: the main feature in the VIMOS spectrum is an emission line at 7723Åidentified with [O ]3727 at z=1.072,
quality flag 2.
10. GDS J033249.85-274757.8:
-- FORS2: object red with bright continuum, the emission line [O ]3727 and the absorption lines MgII 2798 and Hζ are
detected in the FORS2 spectrum, the 4000Å Balmer Break is also evident, quality flag "B".
-- VVDS: flat continuum, the NeV absorption line is identified at z=1.254, quality flag 2.
In summary, out of ten highly discrepant cases we have found only one that can be ascribed to an evident error in the
identification of the features in the FORS2 spectrum (and the original quality flag for this object was "C"). We conclude that the
probable fraction of "catastrophic" misidentifications in Table 2 is at most a few percent.
5.2. Reliabilityoftheredshifts-diagnosticdiagrams
As mentioned above, the photometric information and its relation with the redshift provides useful indications about possible
errors in the redshift measurement and/or magnitude estimation. The Figures 6, 7 and 8 show the redshift-magnitude, the color-
redshift and the color-magnitude distributions for the spectroscopic sample (the quality flag "A" and "B" have been selected in
the Figures 7 and 8, while all sources have been plotted in the Figure 6). In figure 7 the two populations of "emission-line" and
"absorption-line" (typically elliptical) galaxies are clearly separated. The mean color of the "absorption-line" objects increases
from i775 − z850 = 0.46 ± 0.079 at < z > = 0.6 to i775 − z850 = 0.86 ± 0.18 at < z > = 1.0, consistent but increasingly bluer than the
colors of a non-evolving L⋆ elliptical galaxy (estimated integrating the spectral templates of Coleman, Wu & Weedman (1980)
through the ACS bandpasses).
E. Vanzella et al.: The Great Observatories Origins Deep Survey
15
Fig. 6. Spectroscopic redshift versus magnitude for the entire FORS2 sample (quality flag "A", "B", "C" and "X"). Stars are
denoted by star-like symbols at zero redshift. Inconclusive spectra are placed at z = −1.
The "emission-line" objects show in general a bluer i775 − z850 color and a broader distribution than the "absorption-line"
sources: i775 − z850 = 0.16 ± 0.13 at < z > = 0.6 and i775 − z850 = 0.52 ± 0.21 at < z > = 1.1. The broader distribution, with
some of the "emission-line" objects entering the color regime of the ellipticals, is possibly explained by dust obscuration, high
metallicity or strong line emission in the z850 band.
5.3. RedshiftdistributionandLargeScaleStructure
Figure 9 shows the redshift distribution of the objects observed in the present survey. The majority of the sources are at redshift
around ∼1 (the median of the redshift distribution is at 1.04), in agreement with the main criterion for the target selection (see
Sect. 2). Table 4 shows the fraction of determined redshifts as a function of the spectral features identified, i.e. emission lines,
absorption lines, emission & absorption lines, and no reliable spectral features (unclassified). There are 49 galaxies identified
16
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Fig. 7. Color-redshift diagram of the spectroscopic sample. Only redshifts with quality flag "A" and "B" have been selected.
Filled pentagons symbols are objects identified with absorption features only ("abs." sources), while open pentagons are objects
showing only emission lines ("em." sources). The intermediate cases are shown by filled triangles ("comp." sources). The long-
dashed line and the short dashed line show the colors of a non-evolving L⋆ elliptical galaxy and an Scd galaxy, respectively,
estimated integrating the spectral templates of Coleman, Wu & Weedman (1980) through the ACS bandpasses.
with absorption lines only (mainly Ca H and K) in the range of redshift between 0.4-1.3; an example is shown in Figure 1. In 46%
of the total sample we have measured emission lines (mainly [O ]3727), many of them entering the so-called "spectroscopic
desert" up to z=1.61.
The main peaks in the redshift distribution are at z ∼ 0.73 (21 galaxies) and 1.1 (25 galaxies). Two concentrations at z ∼ 1.6
(with 5 galaxies at the mean redshift < z >= 1.612 ± 0.003, see the two dimensional spectra in Figure 11) and z ∼ 0.67 (9
galaxies) are also apparent. The presence in the CDF-S of large scale structure, (LSS) at z ∼ 0.73 and z ∼ 0.67 is already known
(Cimatti et al. (2002), Gilli et al. 2003, Le Fevre et al. 2004). The peak at z ∼ 1.1 seems to be a new indication of large scale
E. Vanzella et al.: The Great Observatories Origins Deep Survey
17
Fig. 8. Color-magnitude diagram for the spectroscopic sample. Only redshifts with quality flag "A" and "B" have been selected.
The symbols are the same as in Figure 7.
structure, of the 25 galaxies in the range 1.09<z<1.11, 10 show emission lines, 9 are ellipticals and 6 are intermediate-type
galaxies.
The significance of the LSS at z = 1.61 is confirmed by:
1. the observations of Gilli et al. (2003) who found a peak in the redshift distribution of X-ray sources at z=1.618 (5 galaxies)
and measured a Poissonian probability of 3.8×10−3 for a chance distribution ;
2. three more galaxies at z = 1.605, 1.610, 1.615 in the K20 survey Cimatti et al. (2002);
The structure at z ≈ 1.61 is extending across a transverse size of ∼ 5 Mpc in a wall-like pattern rather than a group structure
(see Fig. 10).
18
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Table 4. Fractions of sources with different spectral features.
Spectral class
emission
absorption
em. & abs.
stars
unclassified
zmean
1.131
0.950
0.897
0.000
-
zmin
0.117
0.366
0.382
0.000
-
zmax
5.828
1.910
1.317
0.000
-
Fraction
46%
16%
12%
4%
22%
5.4. Highredshiftgalaxies
As discussed in Sect. 2, the target selection includes mainly low redshift objects (z < 2). For three galaxies, however, a redshift
larger than four was measured: the galaxy GDS J033240.01 − 274815.0 at z = 5.828 the only i775-dropout (see Sect. 2) actually
targeted in the present observations and two serendipitously-observed high redshift sources, GDS J033228.84 − 274132.7 and
one object at α = 3h 32m 28.94s, δ = −27◦ 41′ 28.19′′ not present in the catalog v1.0, measured at z = 4.800 and z = 4.882,
respectively.
The i-dropout candidate has been observed with both the Keck and VLT telescopes (Dickinson et al. 2004a). In Figure 12
the FORS2 spectrum of the i-dropout source is shown. The Lyα line is clearly detected at z = 5.828 and shows the blue cut -- off
characteristic of high -- redshift Lyα emitters and the Lyα forest continuum break.
Figure 13 shows a peculiar system of three sources: two emission-line sources above (∼1.5 arcsecond) and below (∼3 arcsec-
ond) the main galaxy GDS J033228.88 − 274129.3, clearly visible in the ACS color image and in the two dimensional spectrum.
The same target has been observed in two different masks adopting the same orientation of the slits. The total exposure time is
≃43 ks. The extracted one dimensional spectra are shown in the right side of the Figure 13.
The main galaxy GDS J033228.88 − 274129.3 has a redshift z = 0.733 with both emission and absorption lines measured
(quality flag "A"): [O ]3727, MgI, Ca H and K, g-band, etc. The bottom object (GDS J033228.84 − 274132.7) shows a solo-
emission line at 7052Å (see the 1-D spectrum), and is not detected in the ACS B band, we interpret this line as Lyα at z = 4.800
with quality "C".
The source above GDS J033228.88 − 274129.3 is most probably a Lyα emitter at redshift z = 4.882 (quality "B"). The
spectrum has been extracted subtracting the contamination of the tail of the main galaxy. After the subtraction the shape of the
spectrum shows the blue cut -- off and the Lyα forest continuum break, typical of the LBGs.
5.5. Dynamicalmassesofgalaxiesatz ∼ 1
Three galaxies, GDS J033215.88 − 274723.1, GDS J033225.86 − 275019.7 and GDS J033230.71 − 274617.2, at redshift
z=0.896, 1.095 and z=1.307 respectively show a spatially resolved [O ]3727 line with a characteristic "tilt" indicative of a high
rotation velocity (see Figure 14).
Various studies have been carried out on the internal kinematics of distant galaxies (Vogt at al. 1996, Vogt at al. 1997,
Moorwood et al. 2001, Pettini et al. 2001 and van Dokkum & Stanford 2001). Rigopoulou et al. 2002 have determined velocity
profiles with a medium resolution grating R∼5000 of three galaxies at z∼0.6 and one at z∼0.8, detected by ISOCAM in the
HDF -- S. For one object they have derived a rotational velocity of 460 km s−1 containing a mass of 1012M⊙ (within a radius of 20
Kpc) significantly higher than the dynamical masses measured in most other local and high redshift spirals.
In the case of GDS J033215.88 − 274723.1, GDS J033225.86 − 275019.7 and GDS J033230.71 − 274617.2, the spectra, in
spite of the relatively low resolution ℜ ∼860, clearly show a tilt of several pixels (corresponding to about 10Å). The measured
velocity increases with increasing distance from the center of the objects reaching a value of the order of and greater than 400
km s−1 at the extremes. For the object GDS J033225.86 − 275019.7 we have measured a displacement between the two extreme
peaks of 11.5Å (top panel of the Figure 14), while a displacement of 9.6Å has been measured in the case of GDS J033215.88 −
274723.1 (middle panel of the Figure 14)
Assuming that the observed velocity structure is due to dynamically-relaxed rotation, then it is possible to estimate the dy-
1.6
sin2(i) ×1011M⊙ for the galaxy GDS J033215.88 −
namical mass for the three galaxies shown in Figure 14 (e.g. Lequeux (1983)):
274723.1 (within a radius of 7.8 Kpc) and 3.1
sin2(i) ×1011M⊙ for the galaxy GDS J033225.86 − 275019.7 (within a radius of 9.8
Kpc). The noisy spectrum of the galaxy GDS J033230.71 − 274617.2 allows us to roughly measure a dynamical mass of the
1.5
sin2(i) ×1011M⊙ (within a radius of 7.5 Kpc). The estimates should be considered a lower limit to the total dynamical mass
order of
because more external parts of the rotating structure might have a lower surface brightness and remain undetected.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
19
Fig. 9. Redshift distribution for the spectroscopic sample with quality A, B and C (23 redshift determinations out of 224 have
quality C). Three objects at z>4 are not shown in the histogram.
5.6. GDS J033210.93 − 274721.5: aspectrumcontaminatedbyanearbygalaxy.
The spectrum of the galaxy GDS J033210.93 − 274721.5 simultaneously shows features corresponding to the redshifts z=1.222
and z=0.417 (Figure 15). The origin of the overlap is the presence of a nearby galaxy (z850 = 19.98, GDS J033210.92−274722.8)
offset by 1.3 arcsecond with a redshift z = 0.417. Light from the brighter z = 0.417 galaxy contaminates the spectrum of the
fainter (z850 = 22.19), higher redshift galaxy GDS J033210.93 − 274721.5 (see Figure 15). Such cases may represent a problem
and a source of error in large spectroscopic surveys, which require an highly automated data processing. A possible solution is
to evaluate a priori on the basis of imaging what are the cases subject of light contamination requiring a "special" reduction.
Alternatively, color-redshift diagrams (such as Figure 7), a comparison of spectroscopic and photometric redshifts or similar
diagnostics are required to carry out the necessary data quality control and identify possible misidentifications.
20
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Fig. 10. The spatial distribution of the galaxies at z ∼ 1.61 in the CDF-S. The background image is an exposure in the R band
obtained with the ESO wide-field imager (WFI). North is up and east on the left. The squares represent the FOV of the FORS2
pointings. The five FORS2 targets are given by their coordinates, and the three K20 sources with the identifiers: OBJ 00235,
OBJ 00237 and OBJ 00270. The numbers 31, 46, 60, 67, show the positions of z ∼ 1.61 X-ray sources (see text).
6. Conclusions
In the framework of the Great Observatories Origins Deep Survey a large sample of galaxies in the Chandra Deep Field South
has been spectroscopically targeted. A total of 303 objects with z850 ∼< 25.5 has been observed with the FORS2 spectrograph at
the ESO VLT providing 234 redshift determinations. From a variety of diagnostics the measurement of the redshifts appears to
be highly accurate (with a typical σz = 0.001) and reliable (with an estimated rate of catastrophic misidentifications at most few
percent). The reduced spectra and the derived redshifts are released to the community (http : //www.eso.org/science/goods/).
They constitute an essential contribution to reach the scientific goals of GOODS, providing the time coordinate needed to delin-
eate the evolution of galaxy masses, morphologies, and star formation, calibrating the photometric redshifts that can be derived
from the imaging data at 0.36-8µm and enabling detailed studies of the physical diagnostics for galaxies in the GOODS field.
Acknowledgements. We are grateful to the ESO staff in Paranal and Garching who greatly helped in the development of this programme.
The work of DS was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with NASA. L.A.M.
acknowledges support by NASA through contract number 1224666 issued by the Jet Propulsion Laboratory, California Institute of Technology
under NASA contract 1407. We thank the ASI grant I/R/088/02 (SC, MN, EV).
E. Vanzella et al.: The Great Observatories Origins Deep Survey
21
Fig. 11. Two dimensional spectra of 5 galaxies at z=1.61. The [O ]3727 emission line is marked with a circle at 9727.5Å. The
absorption sky feature (∼7600Å, A band) is indicated with an arrow. It is worth to note the optimal red sensitivity of FORS2.
References
Abraham, R., G., van den Bergh, S., Glazebrook, K., Ellis, R., S., Santiago, B., X., Surma, P., Griffiths, R., E., 1996, ApJ, 107, 1
Bahcall, J.N., Schmidt, M., Soneira, R.M. 1982, ApJ, 258, 17
Bohlin, R.,C., Colina, L., Finley, D., S., 1995, AJ, 110, 1316
Cimatti, A., Mignoli, M., Daddi, E., et al. 2002, A&A, 392, 395
Coleman, G., D., Wu, C.-C., & Weedman, D., W., 1980, ApJS, 43, 393
Dickinson et al. 2003, in the proceedings of the ESO/USM Workshop "The Mass of Galaxies at Low and High Redshift" (Venice, Italy, October
2001), eds. R. Bender and A. Renzini, astro-ph/0204213
Dickinson, M., et al., 2004, ApJ, 99, 122
Giavalisco, M., et al. 2004, ApJ, 600, L93
Giavalisco, M., Dickinson, M., Ferguson, H. C., Ravindranath, S., Kretchmer, C., Moustakas, L. A., Madau, P., Fall, S. M., Gardner, Jonathan
P., Livio, M., Papovich, C., Renzini, A., Spinrad, H., Stern, D., Riess, A., 2004, ApJ, 600, 103
Gilli, R., Cimatti, A., Daddi, E., Hasinger, G., Rosati, P., Szokoly, G., Tozzi, P., Bergeron, J., Borgani, S., Giacconi, R., Kewley, L., Mainieri,
V., Mignoli, M., Nonino, M., Norman, C., Wang, J., Zamorani, G., Zheng, W., Zirm, A., 2003, ApJ, 592, 721
Le Fevre, O., Vettolani, G., Paltani, S., Tresse, L., Zamorani, G., Le Brun, V., Moreau, C., and the VIMOS VLT Deep Survey team, submitted
to A&A, (astro-ph/0403628)
Lequeux, J. 1983, A&A, 125, 394
Moorwood A.F.M., van der Werf, P.P, Cuby, J.G., Oliva, E., 2000, A&A, 362, 9
Mobasher, B., Idzi, R., Bentez, N., Cimatti, A., Cristiani, S., Daddi, E., Dahlen, T., Dickinson, M., et al., 2004, ApJ, 600, 167
Oke, J.B., et al., 1995, PASP, 107, 375
Pettini, M., Shapley, A. E., Steidel C. C., Cuby, J.-G., Dickinson, M., A. F. M. Moorwood, Adelberger, K. L., Giavalisco, M., 2001, ApJ, 588,
65
Renzini et al. 2002, in the proceedings of the ESO/USM Workshop "The Mass of Galaxies at Low and High Redshift" (Venice, Italy, October
2001), eds. R. Bender and A. Renzini
Riess, A.G.,2 Strolger, L.-G., Tonry, J., Casertano, S., Ferguson, H.C., et al., 2004, ApJ, 607, 665
Rigopoulou, D., Franceschini, A., Aussel, H., Genzel, R., Thatte, N., Cesarsky, C. J., 2002, ApJ, 580, 789
Steidel, C.C., Adelberger, K.L., Giavalisco, M., Dickinson, M., Pettini, M., 1999, ApJ, 519, 1
Szokoly, G., P., Bergeron, J., Hasinger, G., Lehmann, I., Kewley, L., Mainieri, V., Nonino, M., Rosati, P., Giacconi, R., Gilli, R., Gilmozzi, R.,
Norman, C., Romaniello, M., Schreier, E., Tozzi, P., Wang, J., X., Zheng, W., Zirm, A., 2004, (astro-ph/0312324)
van Dokkum, P.G., & Stanford, S.A. 2001, ApJ, 562, 35
Vogt, P.N., et al. 1996, ApJ, 465, L15
Vogt, P.N., et al. 1996, ApJ, 479, L121
Warmels, R.H.: 1991, "The ESO-MIDAS System", in Astronomical Data Analysis Software and Systems I , PASP Conf. Series, Vol. 25, p.
115.
22
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Fig. 12. VLT spectrum of the i775 -- dropout galaxy GDS J033240.01-274815.0.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
23
Fig. 13. Simultaneous spectrum of three sources in the slit. On the right of the figure, the 1D spectra of the z=0.733 main galaxy
GDS J033228.88 − 274129.3, the single emission line ∼ 3 arcsecond below (GDS J033228.84 − 274132.7) and the object ∼ 1.5
arcsecond above are shown. The left-hand panel shows the ACS color image, 5 arcsec on a side. North is up, east is to the left.
The bottom panel shows the 2D spectrum, with the spatial profile obtained by collapsing 80 columns (256 Å), centered at 7150Å,
shown to the right. Candidate serendipitous Lyα emission lines are clearly marked. The object above the target source shows faint
continuum reward of the emission line.
24
E. Vanzella et al.: The Great Observatories Origins Deep Survey
Fig. 14. Three examples of tilted [O ]3727 emission line at redshift around 1. The two dimensional FORS2 spectra are shown
(object and sky lines). In the first two spectra (top and middle) a zoom of the [O ]3727 emission line is shown (the white
rectangle underline the region where the Gaussian fit has been performed to derive the line peak, small black crosses), in the
bottom spectrum the line is too faint to calculate a reliable peak (this object has been serendipitously-identified). In the right side
of the spectra the ACS images of the galaxies and the slits orientation are shown.
E. Vanzella et al.: The Great Observatories Origins Deep Survey
25
Fig. 15. The light merged case, two objects at different redshift superimposed in the slit (marked with white lines in the
left panel). In the right panel the same extracted spectrum with different identifications. An elliptical galaxy (the target,
GDS J033210.93 − 274721.5) at z=1.222 clearly identified with the Ca H and K, Hδ, MgI (quality flag "A"). The bright bluer
object (GDS J033210.92 − 274722.8) shows absorption and emission lines: Ca H, [O ]5007, Na, Hα at z = 0.417 (quality flag
"B"). The Ca K is contaminated by the sky line ∼5577Å.
|
astro-ph/9901114 | 3 | 9901 | 1999-04-08T03:03:08 | The Photo-Evaporation of Dwarf Galaxies During Reionization | [
"astro-ph"
] | During the period of reionization the Universe was filled with a cosmological background of ionizing radiation. By that time a significant fraction of the cosmic gas had already been incorporated into collapsed galactic halos with virial temperatures below about 10000 K that were unable to cool efficiently. We show that photoionization of this gas by the fresh cosmic UV background boiled the gas out of the gravitational potential wells of its host halos. We calculate the photoionization heating of gas inside spherically symmetric dark matter halos, and assume that gas which is heated above its virial temperature is expelled. In popular Cold Dark Matter models, the Press-Schechter halo abundance implies that about 50-90% of the collapsed gas was evaporated at reionization. The gas originated from halos below a threshold circular velocity of 10-15 km/s. The resulting outflows from the dwarf galaxy population at redshifts 5-10 affected the metallicity, thermal and hydrodynamic state of the surrounding intergalactic medium. Our results suggest that stellar systems with a velocity dispersion below about 10 km/s, such as globular clusters or the dwarf spheroidal galaxies of the Local Group, did not form directly through cosmological collapse at high redshifts. | astro-ph | astro-ph | The Photo-Evaporation of Dwarf Galaxies During Reionization
Institute for Advanced Study, Olden Lane, Princeton, NJ 08540
Rennan Barkana1
Astronomy Department, Harvard University, 60 Garden St., Cambridge, MA 02138
Abraham Loeb2
ABSTRACT
During the period of reionization the Universe was filled with a cosmological back-
ground of ionizing radiation. By that time a significant fraction of the cosmic gas had
already been incorporated into collapsed galactic halos with virial temperatures ∼< 104K
that were unable to cool efficiently. We show that photoionization of this gas by the
fresh cosmic UV background boiled the gas out of the gravitational potential wells of its
host halos. We calculate the photoionization heating of gas inside spherically symmetric
dark matter halos, and assume that gas which is heated above its virial temperature
is expelled. In popular Cold Dark Matter models, the Press-Schechter halo abundance
implies that ∼ 50 -- 90% of the collapsed gas was evaporated at reionization. The gas
originated from halos below a threshold circular velocity of ∼ 10 -- 15 km s−1. The re-
sulting outflows from the dwarf galaxy population at redshifts z = 5 -- 10 affected the
metallicity, thermal and hydrodynamic state of the surrounding intergalactic medium.
Our results suggest that stellar systems with a velocity dispersion ∼< 10 km s−1, such
as globular clusters or the dwarf spheroidal galaxies of the Local Group, did not form
directly through cosmological collapse at high redshifts.
Subject headings: cosmology:theory -- galaxies:formation -- galaxies:halos -- radiative
transfer
9
9
9
1
r
p
A
8
3
v
4
1
1
1
0
9
9
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
1.
Introduction
The formation of galaxies is one of the most important, yet unsolved, problems in cosmology.
The properties of galactic dark matter halos are shaped by gravity alone, and have been rigorously
parameterized in hierarchical Cold Dark Matter (CDM) cosmologies (e.g., Navarro, Frenk, & White
1997). However, the complex processes involving gas dynamics, chemistry and ionization, and
1email: [email protected]
2email: [email protected]
-- 2 --
cooling and heating, which are responsible for the formation of stars from the baryons inside these
halos, have still not been fully explored theoretically.
Recent theoretical investigations of early structure formation in CDM models have led to a
plausible picture of how the formation of the first cosmic structures leads to reionization of the
intergalactic medium (IGM). The bottom-up hierarchy of CDM cosmologies implies that the first
gaseous objects to form in the Universe have a low-mass, just above the cosmological Jeans mass of
∼ 104M⊙ (see, e.g., Haiman, Thoul, & Loeb 1996, and references therein). The virial temperature
of these gas clouds is only a few hundred K, and so their metal-poor primordial gas can cool only
due to the formation of molecular hydrogen, H2. However, H2 molecules are fragile, and were easily
photo-dissociated throughout the Universe by trace amounts of starlight (Stecher & Williams 1967;
Haiman, Rees, & Loeb 1996) that were well below the level required for complete reionization of
the IGM. Following the prompt destruction of their molecular hydrogen, the early low-mass objects
maintained virialized gaseous halos that were unable to cool or fragment into stars. Most of the
stars responsible for the reionization of the Universe formed in more massive galaxies, with virial
temperatures Tvir ∼> 104K, where cooling due to atomic transitions was possible. The corresponding
mass of these objects at z ∼ 10 was ∼ 108 M⊙, typical of dwarf galaxies.
The lack of a Gunn-Peterson trough and the detection of Lyα emission lines from sources out
to redshifts z = 5.6 (Weymann et al. 1998; Dey et al. 1998; Spinrad et al. 1998; Hu, Cowie, &
McMahon 1998) demonstrates that reionization due to the first generation of sources must have
occurred at yet higher redshifts; otherwise, the damping wing of Lyα absorption by the neutral
IGM would have eliminated the Lyα line in the observed spectrum of these sources (Miralda-
Escud´e 1998). Popular CDM models predict that most of the intergalactic hydrogen was ionized
at a redshift 8 ∼< z ∼< 15 (Gnedin & Ostriker 1997; Haiman & Loeb 1998a,c). The end of the
reionization phase transition resulted in the emergence of an intense UV background that filled the
Universe and heated the IGM to temperatures of ∼ 1 -- 2 × 104K (Haiman & Loeb 1998b; Miralda-
Escud´e, Haehnelt, & Rees 1998). After ionizing the rarefied IGM in the voids and filaments on
large scales, the cosmic UV background penetrated the denser regions associated with the virialized
gaseous halos of the first generation of objects. Since a major fraction of the collapsed gas had been
incorporated by that time into halos with a virial temperature ∼< 104K, photoionization heating by
the cosmic UV background could have evaporated much of this gas back into the IGM. No such
feedback was possible at earlier times, since the formation of internal UV sources was suppressed
by the lack of efficient cooling inside most of these objects.
The gas reservoir of dwarf galaxies with virial temperatures ∼< 104K (or equivalently a 1D
velocity dispersion ∼< 10 km s−1) could not be immediately replenished. The suppression of dwarf
galaxy formation at z > 2 has been investigated both analytically (Rees 1986; Efstathiou 1992) and
with numerical simulations (Thoul & Weinberg 1996; Quinn, Katz, & Efstathiou 1996; Weinberg,
Hernquist, & Katz 1997; Navarro & Steinmetz 1997). The dwarf galaxies which were prevented
from forming after reionization could have eventually collected gas at z = 1 -- 2, when the UV
background flux declined sufficiently (Babul & Rees 1992; Kepner, Babul, & Spergel 1997). The
-- 3 --
reverse process during the much earlier reionization epoch has not been addressed in the literature.
(However, note that the photo-evaporation of gaseous halos was considered by Bond, Szalay, & Silk
(1988) as a model for Lyα absorbers at lower redshifts z ∼ 4.)
In this paper we focus on the reverse process by which gas that had already settled into
virialized halos by the time of reionization was evaporated back into the IGM due to the cosmic
UV background which emerged first at that epoch. The basic ingredients of our model are presented
in §2. In order to ascertain the importance of a self-shielded gas core, we include a realistic, centrally
concentrated dark halo profile and also incorporate radiative transfer. Generally we find that self-
shielding has a small effect on the total amount of evaporated gas, since only a minor fraction of
the gas halo is contained within the central core. Our numerical results are described in §3. In
particular, we show the conditions in the highest mass halo which can be disrupted at reionization.
We also use the Press-Schechter (1974) prescription for halo abundance to calculate the fraction
of gas in the Universe which undergoes the process of photo-evaporation. Our versatile semi-
analytic approach has the advantage of being able to yield the dependence of the results on a wide
range of reionization histories and cosmological parameters. Clearly, the final state of the gas halo
depends on its dynamical evolution during its photo-evaporation. We adopt a rough criterion for
the evaporation of gas based on its initial interaction with the ionizing background. The precision
of our results could be tested in specific cases by future numerical simulations. In §4 we discuss the
potential implications of our results for the state of the IGM and for the early history of low-mass
galaxies in the local Universe. Finally, we summarize our main conclusions in §5.
2. A Model for Halos at Reionization
We consider gas situated in a virialized dark matter halo. We adopt the prescription for ob-
taining the density profiles of dark matter halos at various redshifts from the Appendix of Navarro,
Frenk, & White (1997, hereafter NFW), modified to include the variation of the collapse overdensity
∆c. Thus, a halo of mass M at redshift z is characterized by a virial radius,
rvir = 0.756
M
108 h−1 M⊙!1/3
(cid:20) Ω0
Ω(z)
or a corresponding circular velocity,
∆c
200(cid:21)−1/3(cid:18) 1 + z
10 (cid:19)−1
h−1 kpc ,
Vc = (cid:18) GM
rvir (cid:19)1/2
= 31.6(cid:18) rvir
h−1 kpc(cid:19)(cid:20) Ω0
Ω(z)
∆c
200(cid:21)1/2(cid:18) 1 + z
10 (cid:19)3/2
km s−1 .
The density profile of the halo is given by
3H 2
0
8πG
ρ(r) =
(1 + z)3 Ω0
Ω(z)
δc
cx(1 + cx)2 ,
(1)
(2)
(3)
where x = r/rvir and c depends on δc for a given mass M . We include the dependence of halo
profiles on Ω0 and ΩΛ, the current contributions to Ω from non-relativistic matter and a cosmological
constant, respectively (see Appendix A for complete details).
-- 4 --
Although the NFW profile provides a good approximation to halo profiles, there are indications
that halos may actually develop a core (e.g., Burkert 1995; Kravtsov et al. 1998; see, however, Moore
et al. 1998). In order to examine the sensitivity of the results to model assumptions, we consider
several different gas and dark matter profiles, keeping the total gas fraction in the halo equal to
the cosmological baryon fraction. The simplest case we consider is an equal NFW profile for the
gas and the dark matter. In order to include a core, instead of the NFW profile of equation (3) we
also consider the density profile of the form fit by Burkert (1995) to dwarf galaxies,
ρ(r) =
3H 2
0
8πG
(1 + z)3 Ω0
Ω(z)
δc
(1 + bx) [1 + (bx)2]
(4)
where b is the inverse core radius, and we set δc by requiring the mean overdensity to equal the
appropriate value, ∆c, in each cosmology (see Appendix A). We also consider two cases where the
dark matter follows an NFW profile but the gas is in hydrostatic equilibrium with its density profile
determined by its temperature distribution. In one case, we assume the gas is isothermal at the
halo virial temperature, given by
Tvir =
µV 2
c
2kB
= 36100
µ
0.6 mp (cid:18) rvir
h−1 kpc(cid:19)2 Ω0
Ω(z)
∆c
10 (cid:19)3
200 (cid:18) 1 + z
K ,
(5)
where µ is the mean molecular weight as determined by ionization equilibrium, and mp is the
proton mass. The spherical collapse simulations of Haiman, Thoul, & Loeb (1996) find a post-
shock gas temperature of roughly twice the value given by equation (5), so we also compare with
the result of setting T = 2 Tvir. In the second case, we let the gas cool for a time equal to the
Hubble time at the redshift of interest, z. Gas above 104K cools rapidly due to atomic cooling
until it reaches a temperature near 104K, where the cooling time rapidly diverges. In this case,
hydrostatic equilibrium yields a highly compact gas cloud when the halo virial temperature is
greater than 104K. In reality, of course, a fraction of the gas may fragment and form stars in these
halos. However, this caveat hardly affects our results since only a small fraction of the gas which
evaporates is contained in halos with Tvir > 104K. Throughout most of our subsequent discussion
we consider the simple case of identical NFW profiles for both the dark matter and the gas, unless
indicated otherwise.
We assume a helium mass fraction of Y = 0.24, and include it in the calculation of the ionization
equilibrium state of the gas as well as its cooling and heating (see, e.g., Katz, Weinberg, & Hernquist
1996). We adopt the various reaction and cooling rates from the literature, including the rates for
collisional excitation and dielectronic recombination from Black (1981); the recombination rates
from Verner & Ferland (1996), and the recombination cooling rates from Ferland et al. (1992) with
a fitting formula by Miralda-Escud´e (1998, private communication). Collisional ionization rates are
adopted from Voronov (1997), with the corresponding cooling rate for each atomic species given by
its ionization rate multiplied by its ionization potential. We also include cooling by Bremsstrahlung
emission with a Gaunt factor from Spitzer & Hart (1971), and by Compton scattering off the
microwave background (e.g., Shapiro & Kang 1987).
-- 5 --
In assessing the effect of reionization, we assume for simplicity a sudden turn-on of an external
radiation field with a specific intensity per unit frequency, ν,
Iν,0 = 10−21 I21(z) (ν/νL)−α erg cm−2 s−1 sr−1Hz−1 ,
(6)
where νL is the Lyman limit frequency. Our treatment of the response of the cloud to this radiation,
as outlined below, is not expected to yield different results with a more gradual increase of the
intensity with cosmic time. The external intensity I21(z) is responsible for the reionization of the
IGM, and so we normalize it to have a fixed number of ionizing photons per baryon in the Universe.
We define the ionizing photon density as
nγ = Z ∞
νL
4πIν,0
hνc
σHI (ν)
σHI (νL)
dν ,
(7)
where the photoionization efficiency is weighted by the photoionization cross section of HI, σHI (ν),
above the Lyman limit. The mean baryon number density is
nb = 2.25 × 10−4 (cid:18) 1 + z
10 (cid:19)3 Ωbh2
0.02 ! cm−3 .
(8)
Throughout the paper we refer to proper densities rather than comoving densities. As our standard
case we assume a post -- reionization ratio of nγ/nb = 1, but we also consider the effect of setting
nγ/nb = 0.1. For example, α = 1.8 and nγ/nb = 1 yield I21 = 1.0 at z = 3 and I21 = 3.5 at
z = 5, close to the values required to satisfy the Gunn-Peterson constraint at these redshifts (see,
e.g., Efstathiou 1992). Note that nγ/nb ∼> 1 is required for the initial ionization of the gas in the
Universe (although this ratio may decline after reionization).
We assume that the above uniform UV background illuminates the outer surface of the gas
cloud, located at the virial radius rvir, and penetrates from there into the cloud. The radiation
photoionizes and heats the gas at each radius to its equilibrium temperature, determined by equat-
ing the heating and cooling rates. The latter assumption is justified by the fact that both the
recombination time and the heating time are initially shorter than the dynamical time throughout
the halo. At the outskirts of the halo the dynamics may start to change before the gas can be
heated up to its equilibrium temperature, but this simply means that the gas starts expanding out
of the halo during the process of photoheating. This outflow should not alter the overall fraction
of evaporated gas.
The process of reionization is expected to be highly non-uniform due to the clustering of the
ionizing sources and the clumpiness of the IGM. As time progresses, the HII regions around the ion-
izing sources overlap, and each halo is exposed to ionizing radiation from an ever increasing number
of sources. While the external ionizing radiation may at first be dominated by a small number of
sources, it quickly becomes more isotropic as its intensity builds up with time (e.g., Haiman & Loeb
1998a,b; Miralda-Escud´e, Haehnelt, & Rees 1998). The evolution of this process depends on the
characteristic clustering scale of ionizing sources and their correlation with the inhomogeneities of
-- 6 --
the IGM. In particular, the process takes more time if the sources are typically embedded in dense
regions of the neutral IGM which need to be ionized first before their radiation shines on the rest
of the IGM. However, in our analysis we do not need to consider these complications since the total
fraction of evaporated gas in bound halos depends primarily on the maximum intensity achieved
at the end of the reionization epoch.
In computing the effect of the background radiation, we include self-shielding of the gas which
is important at the high densities obtained in the core of high redshift halos. For this purpose,
we include radiative transfer through the halo gas and photoionization by the resulting anisotropic
radiation field in the calculation of the ionization equilibrium. We also include the fact that the
ionizing spectrum becomes harder at inner radii, since the outer gas layers preferentially block
photons with energies just above the Lyman limit. We neglect self-shielding due to helium atoms.
Appendix B summarizes our simplified treatment of the radiative transfer equations.
Once the gas is heated throughout the halo, some fraction of it acquires a sufficiently high
temperature that it becomes unbound. This gas expands due to the resulting pressure gradient and
eventually evaporates back to the IGM. The pressure gradient force (per unit volume) kB∇(T ρ/µ)
competes with the gravitational force of ρGM/r2. Due to the density gradient, the ratio between
the pressure force and the gravitational force is roughly the ratio between the thermal energy ∼ kBT
and the gravitational binding energy ∼ µGM/r (which is ∼ kBTvir at rvir) per particle. Thus, if the
kinetic energy exceeds the potential energy (or roughly if T > Tvir), the repulsive pressure gradient
force exceeds the attractive gravitational force and expels the gas on a dynamical time (or faster
for halos with T ≫ Tvir).
We compare the thermal and gravitational energy (both of which are functions of radius) as
a benchmark for deciding which gas shells are expelled from each halo. Note that infall of fresh
IGM gas into the halo is also suppressed due to its excessive gas pressure, produced by the same
photo-ionization heating process.
This situation stands in contrast to feedback due to supernovae, which depends on the effi-
ciency of converting the mechanical energy of the supernovae into thermal energy of the halo gas.
The ability of supernovae to disrupt their host dwarf galaxies has been explored in a number of
theoretical papers (e.g., Larson 1974; Dekel & Silk 1986; Vader 1986, 1987). However, numerical
simulations (Mac-Low & Ferrara 1998) find that supernovae produce a hole in the gas distribution
through which they expel the shock-heated gas, leaving most of the cooler gas still bound.
In
the case of reionization, on the other hand, energy is imparted to the gas directly by the ionizing
photons. A halo for which a large fraction of the gas is unbound by reionization is thus prevented
from further collapse and star formation.
When the gas in each halo is initially ionized, an ionization shock front may be generated (cf.
the discussion of Lyα absorbers by Donahue & Shull 1987). The dynamics of such a shock front
have been investigated in the context of the interstellar medium by Bertoldi & McKee (1990) and
Bertoldi (1989). Their results imply that the dynamics of gas in a halo are not significantly affected
-- 7 --
by the shock front unless the thermal energy of the ionized gas is greater than its gravitational
potential energy. Furthermore, since gas in a halo is heated to the virial temperature even before
reionization, the shock is weaker when the gas is ionized than a typical shock in the interstellar
medium. Also, as noted above, the ionizing radiation reaching a given halo builds up in intensity
over a considerable period of time. Thus, we do not expect the ionization shock associated with the
first encounter of ionizing radiation to have a large effect on the eventual fate of gas in the halo.
3. Results
We assume the most popular cosmology to date (Garnavich et al. 1998) with Ω0 = 0.3 and
ΩΛ = 0.7. We illustrate the effects of cosmological parameters by displaying the results also for
Ω0 = 1, and for Ω0 = 0.3 and ΩΛ = 0. The models all assume Ωbh2 = 0.02 and a Hubble constant
h = 0.5 if Ω0 = 1 and h = 0.7 otherwise (where H0 = 100 h km s−1Mpc−1).
0
Figure 1 shows the temperature of the gas versus its baryonic overdensity ∆b relative to the
cosmic average (cf. Efstathiou 1992). The curves are for z = 8 and assume Ω0 = 0.3 and ΩΛ = 0.7.
We include intergalactic radiation with a flux given by equation (6) for α = 1.8 and nγ/nb = 1.
The dotted curve shows tH = tcool with no radiation field, where tH is the age of the Universe,
approximately equal to 6.5 × 109h−1(1 + z)−3/2Ω−1/2
years at high redshift. This curve indicates
the temperature to which gas has time to cool through atomic transitions before reionization. This
temperature is always near T = 104K since below this temperature the gas becomes mostly neutral
and the cooling time is very long. It is likely that only atomic cooling is relevant before reionization
since molecular hydrogen is easily destroyed by even a weak ionizing background (Haiman, Rees,
& Loeb 1996). The solid curve shows the equilibrium temperature for which the heating time
theat due to a UV radiation field equals the cooling time tcool. The decrease in the temperature at
∆b < 10 is due to the increased importance of Compton cooling, which is proportional to the gas
density rather than its square. At a given density, gas is heated at reionization to the temperature
indicated by the solid curve, unless the net cooling or heating time is too long. The dashed curves
show the temperature where the net cooling or heating time equals tH. By definition, points on
the solid curve have an infinite net cooling or heating time, but there is also a substantial regime
at low ∆b where the net cooling or heating time is greater than tH. However, this regime has only
a minor effect on halos, since the mean overdensity inside the virial radius of a halo is of order 200.
On the other hand, if gas leaves the halo and expands it quickly enters the regime where it cannot
reach thermal equilibrium.
Figure 2 presents an example for the structure of a halo with an initial total mass of M =
3 × 107M⊙ at z = 8. We assume the same cosmological parameters as in Figure 1. The bottom
plot shows the baryon overdensity ∆b versus r/rvir, and reflects our assumption of identical NFW
profiles for both the dark matter and the baryons. The middle plot shows the neutral hydrogen
fraction versus r/rvir, and the top plot shows the ratio of thermal energy per particle (TE = 3
2 kBT )
to potential energy per particle (PE = µφ(r), where φ(r) is the gravitational potential) versus
-- 8 --
r/rvir. The dashed curves assume an optically thin halo, while the solid curves include radiative
transfer and self-shielding. The self-shielded neutral core is apparent from the solid curves, but
since the point where TE/PE = 1 occurs outside this core, the overall unbound fraction does not
depend strongly on the radiative transfer in this case. Its value is 67% assuming an optically -- thin
halo, and 64% when radiative transfer is included and only a fraction of the external photons make
their way inside. Even when the opacity at the Lyman limit is large, some ionizing radiation still
reaches the central parts of the halo because, (i) the opacity drops quickly above the Lyman limit,
and (ii) the heated gas radiates ionizing photons inwards.
Figure 3 shows the unbound gas fraction after reionization as a function of the total halo mass.
We assume Ω0 = 0.3, ΩΛ = 0.7, and nγ/nb = 1. The three pairs of curves shown consist of a solid
line (which includes radiative transfer) and a dashed line (which assumes an optically thin halo).
From right to left, the first pair is for α = 1.8 and z = 8, the second is for α = 5 and z = 8,
and the third is for α = 1.8 and z = 20. In each case the self-shielded core lowers the unbound
fraction when we include radiative transfer (solid vs dashed lines), particularly when the unbound
fraction is sufficiently large that it includes part of the core itself. High energy photons above the
Lyman limit penetrate deep into the halo and heat the gas efficiently. Therefore, a steepening of the
spectral slope from α = 1.8 to α = 5 decreases the temperature throughout the halo and lowers the
unbound gas fraction. This is only partially compensated for by our UV flux normalization, which
increases I21 with increasing α so as to get the same density of ionizing photons in equation (7).
Increasing the reionization redshift from z = 8 to z = 20 increases the binding energy of the gas,
because the high redshift halos are denser. Although the corresponding increase of I21 with redshift
(at a fixed nγ/nb) counteracts this change, the fraction of expelled gas is still reduced due to the
deeper potential wells of higher redshift halos.
From plots similar to those shown in Figure 3, we find the total halo mass at which the unbound
gas fraction is 50%. We henceforth refer to this mass as the 50% mass. Figure 4 plots this mass
as a function of the reionization redshift for different spectra and cosmological models. The solid
line assumes α = 1.8 and the dotted line α = 5, both for Ω0 = 0.3 and ΩΛ = 0.7. The other
lines assume α = 1.8 but different cosmologies. The short-dashed line assumes Ω0 = 0.3, ΩΛ = 0
and the long-dashed line assumes Ω0 = 1. All assume nγ/nb = 1. Gas becomes unbound when its
thermal energy equals its potential binding energy. The thermal energy depends on temperature,
but the equilibrium temperature does not change much with redshift since we increase the UV flux
normalization by the same (1 + z)3 factor as the mean baryonic density. With this prescription for
the UV flux, the 50% mass occurs at a value of the circular velocity which is roughly constant with
redshift. Thus for each curve, the change in mass with redshift is mostly due to the change in the
characteristic halo density, which affects the relation between circular velocity and mass.
The cosmological parameters have only a modest effect on the 50% mass, and change it by up
to 35% at a given redshift. Lowering Ω0 reduces the characteristic density of a halo of given mass,
and so a higher mass is required in order to keep the gas bound. Adding a cosmological constant
reduces the density further through ∆c [see equations (10) and (11)]. For the three curves with
-- 9 --
α = 1.8, the circular velocity of the 50% mass equals 13 km s−1 at all redshifts, up to variations of
a few percent.
The spectral shape of the ionizing flux affects modestly the threshold circular velocity cor-
responding to the 50% mass, because assuming a steeper spectrum (i.e. with a larger α) reduces
the gas temperature and thus requires a shallower potential to keep the gas bound. A higher flux
normalization has the opposite effect of increasing the threshold circular velocity. The left panel
of Figure 5 shows the variation of circular velocity with spectral shape, for two normalizations
(nγ/nb = 1 and nγ/nb = 0.1 for the solid and dashed curves, respectively). The right panel shows
the complementary case of varying the spectral normalization, using two values for the spectral
slope (α = 1.8 and α = 5 for the solid and dashed curves, respectively). All curves assume an
Ω0 = 0.3, ΩΛ = 0.7 cosmology.
Obviously, 50% is a fairly arbitrary choice for the unbound gas fraction at which halos evapo-
rate. Figure 3 shows that for a given halo, the unbound gas fraction changes from 10% to 90% over
a factor of ∼ 60 in mass, or a factor of ∼ 4 in velocity dispersion. When 50% of the gas is unbound,
however, the rest of the gas is also substantially heated, and we expect the process of collapse and
fragmentation to be inhibited. In the extreme case where the gas expands until a steady state is
achieved where it is pressure confined by the IGM, less than 10% of the original gas is left inside the
virial radius. However, continued infall of dark matter should limit the expansion. Numerical sim-
ulations may be used to define more precisely the point at which gas halos are disrupted. Clearly,
photo-evaporation affects even halos with masses well above the 50% mass, although these halos
do not completely evaporate. Note that it is also clear from Figure 3 that not including radiative
transfer would have only a minor effect on the value of the 50% mass (typically ∼ 5%).
Given the values of the unbound gas fraction in halos of different masses, we can integrate to
find the total gas fraction in the Universe which becomes unbound at reionization. This calculation
requires the abundance distribution of halos, which is readily provided by the Press-Schechter mass
function for CDM cosmologies (relevant expressions are given, e.g., in NFW). The high-mass cutoff
in the integration is given by the lowest mass halo for which the unbound gas fraction is zero, since
halos above this mass are not significantly affected by the UV radiation. The low-mass cutoff is
given by the lowest mass halo in which gas has assembled by the reionization redshift. We adopt
for this low-mass cutoff the linear Jeans mass, which we calculate following Peebles (1993, §6). The
gas temperature in the Universe follows the cosmic microwave background temperature down to a
redshift 1 + zt ∼ 740(Ωbh2)2/5, at which the baryonic Jeans mass is 1.9 × 105(Ωbh2)−1/2M⊙. After
this redshift, the gas temperature goes down as (1 + z)2, so the baryon Jeans mass acquires a factor
of [(1 + z)/(1 + zt)]3/2. Until now we have considered baryons only, but if we add dark matter then
the mean density (or the corresponding gravitational force) is increased by Ω0/Ωb, which decreases
the baryonic Jeans mass by (Ω0/Ωb)−3/2. The corresponding total halo mass is Ω0/Ωb times the
baryonic mass. Thus the Jeans cutoff before reionization corresponds to a total halo mass of
-- 10 --
MJ = 6.9 × 103 Ω0h2
0.2 !−
1
2 Ωbh2
0.02 !−
3
2
3
5 (cid:18) 1 + z
10 (cid:19)
M⊙ .
(9)
This value agrees with the numerical spherical collapse calculations of Haiman, Thoul, & Loeb
(1996).
We thus calculate the total fraction of gas in the Universe which is bound in pre-existing halos,
and the fraction of this gas which then becomes unbound at reionization. In Figure 6 we show the
fraction of the collapsed gas which evaporates as a function of the reionization redshift. The solid
line assumes α = 1.8, and the dotted line assumes α = 5, both for Ω0 = 0.3, ΩΛ = 0.7. The other
lines assume α = 1.8, the short-dashed line with Ω0 = 0.3, ΩΛ = 0 and the long-dashed line with
Ω0 = 1. All assume nγ/nb = 1 and a primordial n = 1 (scale invariant) power spectrum. In each
case we normalized the CDM power spectrum to the present cluster abundance, σ8 = 0.5 Ω−0.5
(see, e.g., Pen 1998), where σ8 is the root-mean-square amplitude of mass fluctuations in spheres
of radius 8 h−1 Mpc. The fraction of collapsed gas which is unbound is ∼ 0.4 -- 0.7 at z = 6 and
it increases with redshift. This fraction clearly depends strongly on the halo abundance but is
relatively insensitive to the spectral slope α of the ionizing radiation. In hierarchical models, the
characteristic mass (and binding energy) of virialized halos is smaller at higher redshifts, and a
larger fraction of the collapsed gas therefore escapes once it is photoheated. Among the three
cosmological models, the characteristic mass at a given redshift is smallest for Ω0 = 1 and largest
for Ω0 = 0.3, ΩΛ = 0.
0
In Figure 7 we show the total fraction of gas in the Universe which evaporates at reionization.
The solid line assumes α = 1.8, and the dotted line assumes α = 5, both for Ω0 = 0.3, ΩΛ = 0.7.
The other lines assume α = 1.8, the short-dashed line with Ω0 = 0.3, ΩΛ = 0 and the long-dashed
line with Ω0 = 1. All assume nγ/nb = 1. For the different cosmologies, the total unbound fraction
goes up to 20 -- 25% if reionization occurs as late as z = 6 -- 7; in this case a substantial fraction of
the total gas in the Universe undergoes the process of expulsion from halos. However, this fraction
typically decreases at higher redshifts. Although a higher fraction of the collapsed gas evaporates
at higher z (see Figure 6), a smaller fraction of the gas in the Universe lies in halos in the first place.
The latter effect dominates except for the open model up to z ∼ 7. As is well known, the Ω0 = 1
model produces late structure formation, and indeed the collapsed fraction decreases rapidly with
redshift in this cosmological model. The low Ω0 models approach the Ω0 = 1 behavior at high z,
but this occurs faster for the flat model with a cosmological constant than for the open model with
the same value of Ω0.
Changing the dark matter and gas profiles as discussed in §2 has a modest effect on the results.
For example, with Ω0 = 0.3, ΩΛ = 0.7, and z = 8, and for our standard model where the gas and
dark matter follow identical NFW profiles, the total unbound gas fraction is 19.8% and the halo
mass which loses 50% of its baryons is 5.25 × 107M⊙. If we let the mass and the baryons follow
the profile of equation (4) the corresponding results are 20.0% and 5.31 × 107M⊙ for b = 10 in
-- 11 --
equation (4) and 20.9% and 6.84 × 107M⊙ for b = 5 (i.e. a larger core). With an NFW mass profile
but gas in hydrostatic equilibrium at the virial temperature, the unbound fraction is 19.2%, and
the 50% mass is 4.33 × 107M⊙. If we let the gas temperature be T = 2 Tvir, the unbound fraction
is 22.0% and the 50% mass is 1.18 × 108M⊙. For clouds of gas which condense by cooling for a
Hubble time, the unbound fraction is 18.2%, and the 50% mass is 3.38 × 107M⊙. We conclude
that centrally concentrated gas clouds are in general more effective at retaining their gas, but the
effect on the overall unbound gas fraction in the Universe is modest, even for large variations in
the profile. If we return to the NFW profile but adopt f = 0.01 instead of f = 0.5 in the NFW
prescription for finding the collapse redshift (see Appendix A), we find an unbound fraction of
20.3%, and a 50% mass of 6.06 × 107M⊙. Finally, lowering Ωb by a factor of 2 changes the unbound
fraction to 19.0% and the 50% mass to 5.44 × 107M⊙. Our predictions appear to be robust against
variations in the model parameters.
4.
Implications for the Intergalactic Medium and for Low Redshift Objects
Our calculations show that a substantial fraction of gas in the Universe may lie in virialized
halos before reionization, and that most of it evaporates out of the halos when it is photoionized and
heated at reionization. The resulting outflows of gas from halos may have interesting implications
for the subsequent evolution of structure in the IGM. We discuss some of these implications in this
section.
In the pre-reionization epoch, a fraction of the gas in the dense cores of halos may fragment
and form stars. Some star formation is, of course, needed in order to produce the ionizing flux
which leads to reionization. These population III stars produce the first metals in the Universe,
and they may make a substantial contribution to the enrichment of the IGM. Numerical models
by Mac-Low & Ferrara (1998) suggest that feedback from supernovae is very efficient at expelling
metals from dwarf galaxies of total mass 3.5 × 108 M⊙, although it ejects only a small fraction of
the interstellar medium in these hosts. Obviously, the metal expulsion efficiency depends on the
presence of clumps in the supernova ejecta (Franco et al. 1993) and on the supernova rate -- the
latter depending on the unknown star formation rate and the initial mass function of stars at high
redshifts. Reionization provides an alternative method for expelling metals efficiently out of dwarf
galaxies by directly photoheating the gas in their halos, leading to its evaporation along with its
metal content.3
Gas which falls into halos and is expelled at reionization attains a different entropy than if
it had stayed at the mean density of the Universe. Gas which collapses into a halo is at a high
3Note that we have assumed zero metallicity in calculating cooling. Even if some metals had already been mixed
into the IGM, the metallicity of newly formed objects was likely too low to affect cooling since even at z ∼ 3 the
typical metallicity of the Lyman alpha forest has been observed to be < 0.01 solar (Songaila & Cowie 1996; Tytler
et al. 1995).
-- 12 --
overdensity when it is photoheated, and is therefore at a lower entropy than if it were heated to
the same temperature at the mean cosmic density. However, the overall change in the entropy
density of the IGM is small for two reasons. First, even at z = 6 only about 25% of the gas
in the Universe undergoes evaporation. Second, the gas remains in ionization equilibrium and is
photoheated during its initial expansion. For example, if z = 6, Ω0 = 0.3, ΩΛ = 0.7, nγ/nb = 1,
and α = 1.8, then the recombination time becomes longer than the dynamical time only when the
gas expands down to an overdensity of 26, at which point its temperature is 22,400 K compared
to an initial (non-equilibrium) temperature of 19,900 K for gas at the mean density. The resulting
overall reduction in the entropy is the same as would be produced by reducing the temperature of
the entire IGM by a factor of 1.6. This factor reduces to 1.4 if we increase z to 8 or increase α to 5.
Note that Haehnelt & Steinmetz (1998) showed that differences in temperature by a factor of 3 -- 4
result in possibly observable differences in the Doppler parameter distribution of Lyα absorption
lines at redshifts 3 -- 5.
When the halos evaporate, recombinations in the gas could produce Lyα lines or radiation
from two-photon transitions to the ground state of hydrogen. However, a simple estimate shows
that the resulting luminosity is too small for direct detection unless these halos are illuminated
by an internal ionizing source. In an externally illuminated z = 6, 108M⊙ halo our calculations
imply a total of ∼ 1 × 1050 recombinations per second. Note that the number of recombinations
is dominated by the high density core, and if we did not include self-shielding we would obtain an
overestimate by a factor of ∼ 15. If each recombination releases one or two photons with a total
energy of 10.2 eV, then for Ω0 = 0.3 and ΩΛ = 0.7 the observed flux is ∼ 5 × 10−20 erg s−1 cm−2.
This flux is well below the sensitivity of the planned Next Generation Space Telescope, even if part
of this flux is concentrated in a narrow line.
The photoionization heating of the gaseous halos of dwarf galaxies resulted in outflows with
a characteristic velocity of ∼ 20 -- 30 km s−1. These outflows must have induced peculiar velocities
of a comparable magnitude in the IGM surrounding these galaxies. The effect of the outflows
on the velocity field and entropy of the IGM at z = 5 -- 10 could in principle be searched for in
the absorption spectra of high redshift sources, such as quasars. These small-scale fluctuations in
velocity and the resulting temperature fluctuations have been seen in recent simulations by Bryan
et al. (1998). However, the small halos responsible for these outflows were only barely resolved
even in these high resolution simulations of a small volume.
The evaporating galaxies could contribute to the high column density end of the Lyα forest (cf.
Bond, Szalay, & Silk 1988). For example, shortly after being photoionized, a z = 8, 5×107 M⊙ halo
has a neutral hydrogen column density of 2 × 1016 cm−2 at an impact parameter of 0.5 rvir = 0.66
kpc, 6×1017 cm−2 at 0.25 rvir, and 9×1020 cm−2 (or 9×1018 cm−2 if we do not include self-shielding)
at 0.1 rvir (assuming Ω0 = 0.3, ΩΛ = 0.7, α = 1.8, and nγ/nb = 1). These column densities will
decline as the gas expands out of the host galaxy. Abel & Mo (1998) have suggested that a
large fraction of the Lyman limit systems at z ∼ 3 may correspond to mini-halos that survived
reionization. Remnant absorbers due to galactic outflows can be distinguished from large-scale
-- 13 --
absorbers in the IGM by their compactness. Close lines of sight due to quasar pairs or gravitational
lensed quasars (see, e.g., Crotts & Fang 1998; Petry, Impey, & Foltz 1998, and references therein)
should probe different HI column densities in galactic outflow absorbers but similar column densities
in the larger, more common absorbers. Follow-up observations with high spectroscopic resolution
could reveal the velocity fields of these outflows.
Although much of the gas in the Universe evaporated at reionization, the underlying dark
matter halos continued to evolve through infall and merging, and the heated gas may have accu-
mulated in these halos at lower redshifts. This latter process has been discussed by a number of
authors, with an emphasis on the effect of reionization and the resulting heating of gas. Thoul &
Weinberg (1996) found a reduction of ∼ 50% in the collapsed gas mass due to heating, for a halo
of Vc = 50 km s−1 at z = 2, and a complete suppression of infall below Vc = 30 km s−1. The
effect is thus substantial on halos with virial temperatures well above the gas temperature. Their
interpretation is that pressure support delays turnaround substantially and slows the subsequent
collapse. Indeed, as noted in §2, the ratio of the pressure force to the gravitational force on the gas
is roughly equal to the ratio of its thermal energy to its potential energy. For a given enclosed mass,
the potential energy of a shell of gas increases as its radius decreases. Before collapse, each gas shell
expands with the Hubble flow until its expansion is halted and then reversed. Near turnaround,
the gas is weakly bound and the pressure gradient may prevent collapse even for gas below the
halo virial temperature. On the other hand, gas which is already well within the virial radius is
tightly bound, which explains our lower value of Vc ∼ 13 km s−1 for halos which lose half their gas
at reionization.
Three dimensional numerical simulations (Quinn, Katz, & Efstathiou 1996; Weinberg, Hern-
quist, & Katz 1997; Navarro & Steinmetz 1997) have also explored the question of whether dwarf
galaxies could re-form at z ∼> 2. The heating by the UV background was found to suppress infall
of gas into even larger halos (Vc ∼ 75 km s−1), depending on the redshift and on the ionizing
radiation intensity. Navarro & Steinmetz (1997) noted that photoionization reduces the cooling
efficiency of gas at low densities, which suppresses further the late infall at redshifts below 2. We
note that these various simulations assume an isotropic ionizing radiation field, and do not calcu-
late radiative transfer. Photoevaporation of a gas cloud has been calculated in a two dimensional
simulation (Shapiro, Raga, & Mellema 1998), and methods are being developed for incorporating
radiative transfer into three dimensional cosmological simulations (e.g., Abel, Norman, & Madau
1999; Razoumov & Scott 1999).
Our results have interesting implications for the fate of gas in low-mass halos. Gas evaporates
at reionization from halos below Vc ∼ 13 km s−1, or a velocity dispersion σ ∼ 10 km s−1. A similar
value of the velocity dispersion is also required to reach a virial temperature of 104 K, allowing
atomic cooling and perhaps star formation before reionization. Thus, halos with σ ∼> 10 km s−1
could have formed stars before reionization. They would have kept their gas after reionization, and
could have had ongoing star formation subsequently. These halos were the likely sites of population
III stars, and could have been the progenitors of dwarf galaxies in the local Universe (cf. Miralda-
-- 14 --
Escud´e & Rees 1998). On the other hand, halos with σ ∼< 10 km s−1 could not have cooled before
reionization. Their warm gas was completely evaporated from them at reionization, and could not
have returned to them until very low redshifts, possibly z ∼< 1, so that their stellar population
should be relatively young.
It is interesting to compare these predictions to the properties of dwarf spheroidal galaxies in
the Local Group which have low central velocity dispersions. At first sight this appears to be a
difficult task. The dwarf galaxies vary greatly in their properties, with many showing evidence for
multiple episodes of star formation as well as some very old stars (see the recent review by Mateo
1998). Another obstacle is the low temporal resolution of age indicators for old stellar populations.
For example, if Ω0 = 0.3 and ΩΛ = 0.7 then the age of the Universe is 43% of its present age at
z = 1 and 31% at z = 1.5. Thus, stars that formed at these redshifts may already be ∼ 10 Gyr old
at present, and are difficult to distinguish from stars that formed at z > 5.
Nevertheless, one of our robust predictions is that most early halos with σ ∼< 10 km s−1 could
not have formed stars in the standard hierarchical scenario. Globular clusters belong to one class
of objects with such a low velocity dispersion. Peebles & Dicke (1968) originally suggested that
globular clusters may have formed at high redshifts, before their parent galaxies. However, in
current cosmological models, most mass fluctuations on globular cluster scales were unable to cool
effectively and fragment until z ∼ 10, and were evaporated subsequently by reionization. We note
that Fall & Rees (1985) proposed an alternative formation scenario for globular clusters involving
a thermal instability inside galaxies with properties similar to those of the Milky Way. Globular
clusters have also been observed to form in galaxy mergers (e.g., Miller et al. 1997).
It is still
possible that some of the very oldest and most metal poor globular clusters originated from z ∼> 10,
before the UV background had become strong enough to destroy the molecular hydrogen in them.
However, primeval globular clusters should have retained their dark halos but observations suggest
that globular clusters are not embedded in dark matter halos (Moore 1996; Heggie & Hut 1995).
Another related population is the nine dwarf spheroidals in the Local Group with central
velocity dispersions σ ∼< 10 km s−1, including five below 7 km s−1 (e.g., Mateo 1998).
In the
hierarchical clustering scenario, the dark matter in a present halo was most probably divided at
reionization among several progenitors which have since merged. The velocity dispersions of these
progenitors were likely even lower than that of the final halo. Thus the dwarf galaxies could not
have formed stars at high redshifts, and their formation presents an intriguing puzzle. There are two
possible solutions to this puzzle, (i) the ionizing background dropped dramatically at low redshifts,
allowing the dwarf galaxies to form at z ∼< 1, or (ii) the measured stellar velocity dispersions of the
dwarf galaxies are well below the velocity dispersions of their dark matter halos.
Unlike globular clusters, the dwarf spheroidal galaxies are dark matter dominated. The dark
halo of a present-day dwarf galaxy may have virialized at high redshifts but accreted its gas at
low redshift from the IGM. However, for dark matter halos accumulating primordial gas, Kepner,
Babul, & Spergel (1997) found that even if I21(z) declines as (1 + z)4 below z = 3, only halos
-- 15 --
with Vc ∼> 20 km s−1 can form atomic hydrogen by z = 1, and Vc ∼> 25 km s−1 is required to form
molecular hydrogen.
Alternatively, the dwarf dark halos could have accreted cold gas at low redshift from a larger
host galaxy rather than from the IGM. As long as the dwarf halos join their host galaxy at a redshift
much lower than their formation redshift, they will survive disruption due to their high densities.
The subsequent accretion of gas could result from passages of the dwarf halos through the gaseous
tidal tail of a merger event or through the disk of the parent galaxy. In this case, retainment of cold,
dense, and possibly metal enriched gas against heating by the UV background requires a shallower
potential well than accumulating warm gas from the IGM. Simulations of galaxy encounters (Barnes
& Hernquist 1992; Elmegreen, Kaufman, & Thomasson 1993) have found that dwarf galaxies could
form but with small amounts of dark matter. However, the initial conditions of these simulations
assumed parent galaxies with a smooth dark matter distribution rather than clumpy halos with
dense sub-halos inside them. Simulations by Klypin et al. (1999) suggest that galaxy halos may
have large numbers of dark matter satellites, most of which have no associated stars. If true, this
implies that the dwarf spheroidal galaxies might be explained even if only a small fraction of dwarf
dark halos accreted gas and formed stars.
A common origin for the Milky Way's dwarf satellites (and a number of halo globular clusters),
as remnants of larger galaxies accreted by the Milky Way galaxy, has been suggested on independent
grounds. These satellites appear to lie along two (e.g., Majewski 1994) or more (Lynden-Bell &
Lynden-Bell 1995, Fusi-Pecci et al. 1995) polar great circles. The star formation history of the dwarf
galaxies (e.g., Grebel 1998) constrains their merger history, and implies that the fragmentation
responsible for their appearance must have occured early in order to be consistent with the variation
in stellar populations among the supposed fragments (Unavane, Wyse, & Gilmore 1996; Olszewski
1998). Observations of interacting galaxies (outside the Local Group) also suggest the formation
of "tidal dwarf galaxies" (e.g., Duc & Mirabel 1997).
Finally, there exists the possibility that the measured velocity dispersion of stars in the dwarf
spheroidals underestimates the velocity dispersion of their dark halos. Assuming that the stars are
in equilibrium, their velocity dispersion could be lower than that of the halo if the mass profile
is shallower than isothermal beyond the stellar core radius. As discussed in §2, halo profiles are
thought to vary from being shallow in a central core to being steeper than isothermal at larger
distances. The velocity dispersion and mass to light ratio of a dwarf spheroidal could also appear
high if it is non-spherical or the stellar orbits are anisotropic. Some dwarf spheroidals may even
not be dark matter dominated if they are tidally disrupted (e.g., Kroupa 1997). The observed
properties of dwarf spheroidals require a central mass density of order 0.1M⊙ pc−3 (e.g., Mateo
1998), which is ∼ 7 × 105 times the present critical density. The stars therefore reside either in
high-redshift halos or in the very central parts of low redshift halos. Detailed observations of the
velocity dispersion profiles of these stars could be used to discriminate between these possibilities.
-- 16 --
5. Conclusions
We have shown that the photoionizing background radiation which filled the Universe during
reionization likely boiled most of the virialized gas out of CDM halos at that time. The evaporation
process probably lasted of order a Hubble time due to the gradual increase in the UV background as
the HII regions around individual sources overlapped and percolated until the radiation field inside
them grew up to its cosmic value -- amounting to the full contribution of sources from the entire
Hubble volume. The precise reionization history depends on the unknown star formation efficiency
and the potential existence of mini-quasars in newly formed halos (Haiman & Loeb 1998a).
The total fraction of the cosmic baryons which participate in the evaporation process depends
on the reionization redshift, the ionizing intensity, and the cosmological parameters, but is not
very sensitive to the precise gas and dark matter profiles of the halos. The central core of halos
is typically shielded from the external ionizing radiation by the surrounding gas, but this core
typically contains < 20% of the halo gas and has only a weak effect on the global behavior of the
gas. We have found that halos are disrupted up to a circular velocity Vc ∼ 13 km s−1 for a shallow,
quasar-like spectrum, or Vc ∼ 11 km s−1 for a stellar spectrum, assuming the photoionizing sources
build up a density of ionizing photons comparable to the mean cosmological density of baryons.
At this photoionizing intensity, the value of the circular velocity threshold is nearly independent of
redshift. The corresponding halo mass changes, however, from ∼ 108M⊙ at z = 5 to ∼ 107M⊙ at
z = 20, assuming a shallow ionizing spectrum.
Based on these findings, we expect that both globular clusters and Local Group dwarf galaxies
with velocity dispersions ∼< 10 km s−1 formed at low redshift, most probably inside larger galaxies.
The latter possibility has been suggested previously for the Milky Way's dwarf satellites based on
their location along polar great circles.
We are grateful to Jordi Miralda-Escud´e, Chris McKee, Roger Blandford, Lars Hernquist, and
David Spergel for useful discussions. We also thank Renyue Cen and Jordi Miralda-Escud´e for
assistance with the reaction and cooling rates. RB acknowledges support from Institute Funds.
This work was supported in part by the NASA NAG 5-7039 grant (for AL).
APPENDIX A: Halo profile
We follow the prescription of NFW for obtaining the density profiles of dark matter halos, but
instead of adopting a constant overdensity of 200 we use the fitting formula of Bryan & Norman
(1998) for the virial overdensity:
for a flat Universe with a cosmological constant and
∆c = 18π2 + 82d − 39d2
∆c = 18π2 + 60d − 32d2
(10)
(11)
-- 17 --
for an open Universe, where d ≡ Ω(z) − 1. Given Ω0 and ΩΛ, we define
Ω(z) =
Ω0(1 + z)3 + ΩΛ + (1 − Ω0 − ΩΛ)(1 + z)2 .
Ω0(1 + z)3
In equation (3) c is determined for a given δc by the relation
δc =
∆c
3
c3
ln(1 + c) − c/(1 + c)
.
The characteristic density is given by
1 + z (cid:19)3
δc = C(f )Ω(z)(cid:18) 1 + zcoll
.
(12)
(13)
(14)
For a given halo of mass M , the collapse redshift zcoll is defined as the time at which a mass M/2
was first contained in progenitors more massive than some fraction f of M . This is computed using
the extended Press-Schechter formalism (e.g. Lacey & Cole 1993). NFW find that f = 0.01 fits
their z = 0 simulation results best. Since we are interested in high redshifts when mergers are very
frequent, we adopt the more natural f = 0.5 but also check the f = 0.01 case. [For example, the
survival time of a z = 8, 5 × 107 M⊙ halo before it merges is ∼ 30 -- 40% of the age of the Universe
at that redshift (Lacey & Cole 1993).] In both cases we adopt the normalization of NFW, which is
C(0.5) = 2 × 104 and C(0.01) = 3 × 103.
APPENDIX B: Radiative Transfer
We neglect atomic transitions of helium atoms in the radiative transfer calculation. We only
consider halos for which kBT is well below the ionization energy of hydrogen, and so following
Tajiri & Umemura (1998) we assume that recombinations to excited levels do not result in further
ionizations. On the other hand, recombinations to the ground state result in the emission of ionizing
photons all of which are in a narrow frequency band just above the Lyman limit frequency ν = νL.
We follow separately these emitted photons and the external incoming radiation. The external
photons undergo absorption with an optical depth at the Lyman limit determined by
dτνL
ds
= σHI (νL)nHI .
The emitted photons near νL are propagated by the equation of radiative transfer,
dIν
ds
= −σHI (ν)nHI Iν + ην .
(15)
(16)
Assuming all emitted photons are just above ν = νL, we can set σHI (ν) = σHI (νL) in this equation
and propagate the total number flux of ionizing photons,
F1 ≡ Z ∞
νL
Iν
hν
dν .
(17)
-- 18 --
The emissivity term for this quantity is
Z ∞
νL
ην
hν
dν =
ω
4π
αHI nHIIne ,
(18)
where αHI is the total recombination coefficient to all bound levels of hydrogen and ω is the fraction
of recombinations to the ground state. In terms of Table 5.2 of Spitzer (1978), ω = (φ1 − φ2)/φ1.
We find that a convenient fitting formula up to 64, 000 K, accurate to 2%, is (with T in K)
ω = 0.205 − 0.0266 ln(T ) + 0.0049 ln2(T ) .
(19)
When these photons are emitted they carry away the kinetic energy of the absorbed electron.
When the photons are re-absorbed at some distance from where they were emitted, they heat the
gas with this extra energy. Since kBT ≪ hνL we do not need to compute the exact frequency
distribution of these photons. Instead we solve a single radiative transfer equation for the total flux
of energy (above the ionization energy of hydrogen) in these photons,
F2 ≡ Z ∞
νL
Iν
hν
(hν − hνL)dν .
(20)
The emissivity term for radiative transfer of F2 is
Z ∞
νL
ην
hν
(hν − hνL)dν =
2.07 × 10−11
χ1(β) − χ2(β)
T 1/2
4π
nHIIne erg cm−3 s−1 sr−1 ,
(21)
where β = hνL/kT , T is in K, and the functions χ1 and χ2 are given in Table 6.2 of Spitzer (1978).
We find a fitting formula up to 64, 000 K, accurate to 2% (with T in K):
χ1(T ) − χ2(T ) = ( 0.78
−0.172 + 0.255 ln(T ) − 0.0171 ln2(T ) otherwise.
if T < 103 K
(22)
From each point we integrate along all lines of sight to find τνL, F1 and F2 as a function of angle.
Because of spherical symmetry, we do this only at each radius, and the angular dependence only
involves θ, the angle relative to the radial direction. We then integrate to find the photoionization
rate. For each atomic species, the rate is
Γγi = Z 4π
0
dΩZ ∞
νi
Iν
hν
σi(ν)dν s−1 ,
(23)
where νi and σi(ν) are the threshold frequency and cross section for photoionization of species
i, given in Osterbrock [1989; see Eq.
(2.31) for HI, HeI and HeII]. For the external photons
the UV intensity is Iν,0e−τν , with the boundary intensity Iν,0 = IνL,0(ν/νL)−α as before, and τν
approximated as τνL(ν/νL)−3. Since σi(ν) has the simple form of a sum of two power laws, the
frequency integral in Γγi can be done analytically, and only the angular integration is computed
numerically (cf. the similar but simpler calculation of Tajiri & Umemura 1998). There is an
-- 19 --
additional contribution to photoionization for HI only, from the emitted photons just above ν = νL,
0 dΩ σi(νL)F1. The photoheating rate per unit volume is niǫi, where ni is the number
given by R 4π
density of species i and
ǫi = Z 4π
0
dΩZ ∞
νi
Iν
hν
σi(ν)(hν − hνL)dν s−1ergs s−1 .
(24)
The rate for the external UV radiation is calculated for each atomic species similarly to the calcu-
lation of Γγi. The emitted photons contribute to ǫHI an extra amount of R 4π
0 dΩ σi(νL)F2.
Abel, T., & Mo, H. J. 1998, ApJ, 494, L151
REFERENCES
Abel, T., Norman, M. L., & Madau, P. 1999, submitted to ApJL, astro-ph/9812151
Babul, A., & Rees, M. 1992, MNRAS, 253, 31
Barnes, J. E., & Hernquist, L. 1992, Nature, 360, 715
Bertoldi, F. 1989, ApJ, 346, 735
Bertoldi, F., & McKee, C. F. 1990, ApJ, 354, 529
Black, J. H. 1981, MNRAS, 197, 553
Bond, J. R., Szalay, A. S., & Silk, J. 1988, ApJ, 324, 627
Bryan, G. L., Machacek, M., Anninos, P., & Norman, M. L. 1998, to appear in ApJ, astro-
ph/9805340
Bryan, G., & Norman, M. 1998, ApJ, 495, 80
Burkert, A. 1995, ApJ, 447, L25
Crotts, A. P. S., & Fang, Y. 1998, ApJ, 502, 16
Duc, P.-A., & Mirabel, I. F. 1997, Proceedings of IAU Symp. 186 (Kyoto), astro-ph/9711253
Dekel, A., & Silk, J. 1986, ApJ, 303, 39
Dey, A., Spinrad, H., Stern, D., Graham, J. R., & Chaffee, F. H. 1998, ApJ, 498, L93
Donahue, M., & Shull, J. M. 1987, ApJ, 323, L13
Elmegreen, B. G., Kaufman, M., & Thomasson, M. 1993, ApJ, 412, 90
Efstathiou, G. 1992, MNRAS, 256, 43
-- 20 --
Fall, S. M., & Rees, M. J. 1985, ApJ, 298, 18
Ferland, G. J., Peterson, B. M., Horne, K., Welsh, W. F., & Nahar, S. N. 1992, ApJ, 387, 95
Franco, J., Ferrara, A., Roczyska, M., Tenorio-Tagle, G., & Cox, D. P. 1993, ApJ, 407, 100
Fusi-Pecci, F., Ballazzini, M., Cacciari, C., & Ferraro, F. R. 1995, AJ, 100, 1664
Garnavich, P. M., et al. 1998, ApJ, 509, 74
Gnedin, N. Y., & Ostriker, J. P. 1997, ApJ, 486, 581
Grebel, E. 1998, invited review to appear in IAU Symp. 192, "The Stellar Content of the Local
Group", astro-ph/9812443
Haiman, Z., & Loeb, A. 1998a, ApJ, 503, 505
-- -- -- -- -- -- -- -- -- -- . 1998b, ApJ, in press, astro-ph/9807070
-- -- -- -- -- -- -- -- -- -- . 1998c, invited contribution to Proc. of 9th Annual October Astrophysics
Conference in Maryland, "After the Dark Ages: When Galaxies Were Young (the Universe
at 2 < z < 5)", College Park, October 1998, astro-ph/9811395
Haiman, Z., Rees, M., & Loeb, A. 1996, ApJ, 476, 458; erratum, ApJ, 484, 985
Haiman, Z., Thoul, A. A., & Loeb, A. 1996, ApJ, 464, 523
Haehnelt, M. G., & Steinmetz, M. 1998, MNRAS, 298, 21
Heggie, D. C., & Hut, P. 1995,
in IAU Symp. 174, Dynamical Evolution of Star Clusters-
Confrontation of Theory and Observations, ed. P. Hut & J. Makino (Dordrecht: Kluwer)
Hu, E. M., Cowie, L. L., & McMahon, R. G. 1998, ApJ, 502, L99
Katz, N., Weinberg, D. H., & Hernquist, L. 1996, ApJS, 105, 19
Kepner, J. V., Babul, A., & Spergel, D. N. 1997, ApJ, 487, 61
Klypin, A. A., Kravtsov, A. V., Valenzuela, O., & Prada, F. 1999, submitted to ApJ, astro-
ph/9901240
Kravtsov, A. V., Klypin, A. A., Bullock, J. S., & Primack, J. R. 1998, ApJ, 502, 48
Kroupa, P. 1997, NewA 2, 139
Larson, R. B. 1974, MNRAS, 271, 676L
Lacey, C. G., & Cole, S. M. 1993, MNRAS, 262, 627
Lynden-Bell, D., & Lynden-Bell, R. M. 1995, MNRAS, 275, 429
-- 21 --
Majewski, S. R. 1994, ApJ, 431, L17
Mateo, M. 1998, ARAA, 36, 435
Miller, B. W., Whitmore, B. C., Schweizer, F., & Fall, S. M. 1997, AJ, 114, 2381
Miralda-Escud´e, J. 1998, ApJ, 501, 15
Miralda-Escude, J., Haehnelt, M., & Rees, M. J. 1998, submitted to ApJ, astro-ph/9812306
Miralda-Escude, J., & Rees, M. J. 1998, ApJ, 497, 21
Moore, B. 1996, ApJ, 461, L13
Moore, B., Governato, F., Quinn, T., Stadel, J., & Lake, G. 1998, ApJ, 499, L5
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1997, ApJ, 490, 493 (NFW)
Navarro, J. F., & Steinmetz, M. 1997, ApJ, 478, 13
Olszewski, E. W. 1998, in Galactic Halos: A UC Santa Cruz Workshop, ed. D. Zaritski (San
Francisco: ASP)
Peebles, P. J. E. 1993, Principles of Physical Cosmology (Princeton: Princeton Univ. Press)
Peebles, P. J. E., & Dicke, R. H. 1968, ApJ, 154, 891
Pen, U.-L. 1998, ApJ, 498, 60
Petry, C. E., Impey, C. D., & Foltz, C. B. 1998, ApJ, 494, 60
Press, W. H., & Schechter, P. 1974, ApJ, 187, 425
Quinn, T., Katz, N., & Efstathiou, G. 1996, MNRAS 278, L49
Razoumov, A., & Scott, D. 1999, submitted to MNRAS, astro-ph/9810425
Rees, M. J. 1986, MNRAS, 218, 25
Shapiro, P. R., & Kang, H. 1987, ApJ, 318, 32
Shapiro, P. R., Raga, A. C., and Mellema, G. 1998, in Molecular Hydrogen in the Early Universe,
eds. E. Corbelli, D. Galli, and F. Palla, Memorie Della Societa Astronomica Italiana, 69,
pp. 463- 469, astro-ph/9804117
Songaila, A., & Cowie, L. L. 1996, AJ, 112, 335
Stecher, T. P., & Williams, D. A. 1967, ApJ, 149, L29
Spitzer, L., Jr. 1978, Physical Processes in the Interstellar Medium (New York: Wiley)
-- 22 --
Spitzer, L., Jr., & Hart, M. H. 1971, ApJ, 166, 483
Tajiri, Y., & Umemura, M. 1998, ApJ, 502, 59
Tytler, D. et al. 1995, in QSO Absorption Lines, ed. G. Meylan (Berlin: Springer)
Thoul, A. A., & Weinberg, D. H. 1996, ApJ, 465, 608
Unavane, M., Wyse, R. F. G.,& Gilmore, G. 1996, MNRAS, 278, 727
Vader, J. P. 1986, ApJ, 305, 669
Vader, J. P. 1987, ApJ, 317, 128
Verner, D. A., & Ferland, G. J. 1992, ApJS, 103, 467
Voronov, G. S. 1997, ADNDT, 65, 1
Weinberg, D. H., Hernquist, L., & Katz, N. 1997, ApJ, 477, 8
This preprint was prepared with the AAS LATEX macros v4.0.
-- 23 --
Fig. 1. -- Temperature T versus baryon overdensity ∆b. Dotted curve: tH = tcool, no radiation
field. Solid curve: theat = tcool, with radiation. Dashed curves: tH =net cooling/heating time, with
radiation. The curves assume Ω0 = 0.3, ΩΛ = 0.7, z = 8, α = 1.8 and nγ/nb = 1.
-- 24 --
Fig. 2. -- Halo structure after reionization, for a halo mass M = 3 × 107M⊙ at z = 8. Bottom
plot: Baryon overdensity (∆b) versus r/rvir. Middle plot: Neutral hydrogen fraction (nHI /nH )
versus r/rvir. Top plot: Ratio of thermal energy (TE) to potential energy (PE) versus r/rvir. The
dashed curves assume an optically thin halo, while the solid curves include radiative transfer and
self-shielding. All the curves assume Ω0 = 0.3, ΩΛ = 0.7, α = 1.8 and nγ/nb = 1.
-- 25 --
Fig. 3. -- Unbound gas fraction versus total halo mass. There are three pairs of curves, each
consisting of a solid line (with radiative transfer) and a dashed line (without radiative transfer).
From right to left, the three sets of curves correspond to α = 1.8, z = 8; α = 5, z = 8; and α = 1.8,
z = 20. All the curves assume Ω0 = 0.3, ΩΛ = 0.7, and nγ/nb = 1.
-- 26 --
Fig. 4. -- Total halo mass for which 50% of the gas is unbound versus reionization redshift. The
solid line assumes α = 1.8, and the dotted line assumes α = 5, both for an Ω0 = 0.3, ΩΛ = 0.7
cosmology. The other lines assume α = 1.8 but different cosmologies. The short-dashed line
assumes Ω0 = 0.3, ΩΛ = 0 and the long-dashed line assumes Ω0 = 1. All assume nγ/nb = 1.
-- 27 --
Fig. 5. -- Circular halo velocity at which 50% of the gas is unbound, as a function of ionizing
spectrum. Both panels show Vc on the vertical axis. The left panel varies the spectral slope α for
two values of the normalization, nγ/nb = 1 (solid curve) and nγ/nb = 0.1 (dashed curve). The right
panel varies the normalization for two spectral slopes, α = 1.8 (solid curve) and α = 5 (dashed
curve). All curves assume an Ω0 = 0.3, ΩΛ = 0.7 cosmology.
-- 28 --
Fig. 6. -- Fraction of the collapsed gas which evaporates from halos at reionization, versus the
reionization redshift. The solid line assumes α = 1.8, and the dotted line assumes α = 5, both for
an Ω0 = 0.3, ΩΛ = 0.7 cosmology. The other lines assume α = 1.8 but different cosmologies. The
short-dashed line assumes Ω0 = 0.3, ΩΛ = 0 and the long-dashed line assumes Ω0 = 1. All assume
nγ/nb = 1.
-- 29 --
Fig. 7. -- Total fraction of gas in the Universe which evaporates from halos at reionization, versus
the reionization redshift. The solid line assumes α = 1.8, and the dotted line assumes α = 5, both
for an Ω0 = 0.3, ΩΛ = 0.7 cosmology. The other lines assume α = 1.8 but different cosmologies.
The short-dashed line assumes Ω0 = 0.3, ΩΛ = 0 and the long-dashed line assumes Ω0 = 1. All
assume nγ/nb = 1.
|
astro-ph/0110294 | 1 | 0110 | 2001-10-12T11:15:39 | X-ray Properties of Young Stellar Objects in OMC-2 and OMC-3 from the Chandra X-ray Observatory | [
"astro-ph"
] | We report X-ray results of the Chandra observation of Orion Molecular Cloud 2 and 3. A deep exposure of \sim 100 ksec detects \sim 400 X-ray sources in the field of view of the ACIS array, providing one of the largest X-ray catalogs in a star forming region. Coherent studies of the source detection, time variability, and energy spectra are performed. We classify the X-ray sources into class I, class II, and class III+MS based on the J, H, and K-band colors of their near infrared counterparts and discuss the X-ray properties (temperature, absorption, and time variability) along these evolutionary phases. | astro-ph | astro-ph |
X-ray Properties of Young Stellar Objects in OMC-2 and OMC-3
from the Chandra X-ray Observatory
Department of Physics, Graduate School of Science, Kyoto University, Sakyo-ku, Kyoto, 606-8502, Japan
[email protected], [email protected]
Masahiro Tsujimoto and Katsuji Koyama
Department of Physics & Astrophysics, 525 Davey Laboratory, Pennsylvania State University, University
Yohko Tsuboi
Park, PA 16802, USA
[email protected]
and
National Astronomical Observatory of Japan, 650 North A'ohoku Place, Hilo, HI 96720, USA
Miwa Goto and Naoto Kobayashi
[email protected], [email protected]
ABSTRACT
We report X-ray results of the Chandra observation of Orion Molecular Cloud 2 and 3. A
deep exposure of ∼ 100 ksec detects ∼ 400 X-ray sources in the field of view of the ACIS array,
providing one of the largest X-ray catalogs in a star forming region. Coherent studies of the source
detection, time variability, and energy spectra are performed. We classify the X-ray sources into
class I, class II, and class III + MS based on the J, H, and K -band colors of their near infrared
counterparts and discuss the X-ray properties (temperature, absorption, and time variability)
along these evolutionary phases.
Subject headings: stars: pre-main sequence -- X-rays: stars -- infrared: stars -- Individual: OMC-2,
OMC-3
1.
INTRODUCTION
Stars are known to possess high energy phe-
nomena well before they reach the main sequence
(MS)1. The Einstein observatory discovered that
PMS stars (class III and class II stage with ages
of ∼ 107 and ∼ 106 years after the on-set of grav-
1In this paper, we define terminologies as follows; "Proto-
stars" are class 0 and class I objects. "Pre-main sequence
(PMS) stars" and "T Tauri Stars (TTS)" are class II and
class III objects. TTS comprise two subclasses, Classical T
Tauri Stars (CTTS) and Weak-line T Tauri Stars (WTTS),
corresponding to class II and class III, respectively. "Young
Stellar Objects (YSO)" are class 0, I, II, and III sources col-
lectively.
itational collapse, respectively) are strong X-ray
emitters (Feigelson & DeCampli 1981; Montmerle
et al. 1983). Successive observations with ASCA
and ROSAT detected hundreds of X-ray samples
of PMS stars with better spectral or spatial res-
olution. They revealed that the X-ray activities
in these objects are quite common. Koyama et al.
(1996) and Kamata et al. (1997) moved the start
of X-ray activity forward to protostars. They de-
tected hard X-rays from class I (∼ 105 years) pro-
tostars with ASCA. Recently, Tsuboi et al. (2001)
reported hard X-ray emission from two class 0 can-
didates in the Orion Molecular Cloud (OMC) us-
ing the Chandra X-ray Observatory (CXO). Now
1
YSOs at virtually all the evolutionary phases --
from class 0 to class III -- are known to emit X-
rays.
However, the X-ray emission mechanism has
not been well understood. Imanishi, Koyama, &
Tsuboi (2001) observed the ρ-Ophiuchi dark cloud
with the CXO and found that the X-ray emit-
ting regions of several sources are well beyond
their stellar surface. This may be explained if the
X-rays are due to star-disk magnetic interaction.
Grosso et al. (2000) reported, on the other hand,
that the X-ray emission does not depend on the ex-
istence of disks; no significant difference of the X-
ray luminosity function is seen between ROSAT -
detected class III objects (no disk) and class II
objects (with disks). Montmerle et al. (2000) ad-
vocated that the reconnection of magnetic loops
between stellar surface and disk is responsible for
quasi-periodic flares and extremely high luminos-
ity X-rays found in a protostar (YLW16) with
ASCA (Tsuboi et al. 2000). Schulz et al. (2001)
detected high temperature plasma with moderate
luminosity and no rapid variability from YSOs in
the Orion Nebula Cluster with the CXO. Hence
they argued that the stellar surface-disk arcade
model is unlikely but magnetically confined stel-
lar plasma is a more likely origin for the X-ray
activities, as was suggested by Skinner & Walter
(1998) based on the ASCA observation of a CTTS
(SU Aur). These arguments from previous studies
may not be disagreements with each other, but are
far from a unified picture of X-ray emission from
YSOs.
Orion Molecular Cloud 2 and 3 (OMC-2 and
OMC-3), located at a distance of ∼ 450 pc from
us, are the best sites to study X-ray properties of
YSOs at various evolutionary phases from class
0 through class III. X-rays from class 0 candi-
dates have only been reported in OMC-3 (Tsuboi
et al. 2001). Therefore, this cloud is the unique
site where we actually provide X-ray samples from
class 0 to class III. Moreover, the moderate angu-
lar size allows us to fully cover the clouds in the
CXO/ACIS field of view (FOV).
Here, we report results of the CXO observation
of the OMC-2 and OMC-3 star forming regions,
with one of the deepest exposures of star form-
ing regions. The purposes of this paper are 1) to
conduct a coherent study of the X-ray source de-
tection, time variability, and energy spectra, and
2
to provide a catalog of one of the largest samples
of X-ray sources in a star forming region, and 2) to
investigate the X-ray properties of YSOs as a func-
tion of evolutionary phase inferred from the near
infrared (NIR) data, and to compare them with
the results of other star forming regions.
For the second purpose we use the 2MASS data,
which is the deepest currently available in these
regions. We conducted deeper infrared surveys on
these clouds down to ∼ 18 mag in the J, H, and K -
band. Those results will be presented in a separate
paper.
2. OBSERVATION
The observation of OMC-2 and OMC-3 was car-
ried out on 2000 January 1 -- 2 using the CXO with
a nominal exposure time of 89.2 ksec. We used
four ACIS-I chips (I0, I1, I2, and I3) and one
ACIS-S chip (S2) on the focal plane of the mir-
ror system. All these detector chips utilize front-
illuminated CCDs, which have sensitivity over a
wide energy range (0.2 -- 10.0 keV) with moder-
ate energy resolution (∆E ∼ 200 eV at E =
6 keV). Together with the optics, ACIS achieves
sub-arcsec spatial resolution for on-axis sources,
and a 8.′4 × 8.′4 FOV for each CCD chip. De-
tails on the satellite and the detectors are found
in Weisskopf, O'dell, & van Speybroeck (1996) and
Garmire et al. (2001), respectively.
Its spectroscopic capability in the hard energy
band, together with the sub-arcsec spatial resolu-
tion, makes the CXO/ACIS an ideal instrument
for the observation of star forming regions, where
sources are heavily absorbed and crowded.
Figure 1 shows the FOV of our observation
overlaid on the 1300mm intensity contour map
(Chini et al. 1997). OMC-2 and OMC-3 are sepa-
rated into south and north by the dotted line. Our
observation fully covers both star forming clouds.
3. DATA REDUCTION AND ANALY-
SIS
3.1. Data Reduction
For the data reduction, we use the level 2
data "reprocessed" at the Chandra X-ray Center
(CXC). This version improves the aspect solution
and restores the degradation of energy gain and
resolution due to the increase of the charge trans-
fer inefficiency (CTI) of the ACIS2. For data ma-
nipulation, we use the Chandra Interactive Analy-
sis of Observations (CIAO) version 2.1 and FTOOLS
version 4.2.
In order to avoid the "spiking" problem re-
ported by the CXC3, we manually randomize the
values of ENERGY and PI columns in the event
file. After this procedure, we conduct the anal-
ysis in the following sections separately for the
ACIS-I and ACIS-S data. Throughout this paper,
we use X-ray photons in the 0.5 -- 8.0 keV energy
band. The photons of each source are accumu-
lated from an elliptical region. The major and
minor axes, and the rotation angles are derived
from the wavdetect command.
3.2. Source Finding
We use wavdetect command with the signifi-
cance criterion of 1 × 10−5 and the wavelet scales
ranging from 1 to 16 pixels in multiples of √2.
We remove a few spurious sources through care-
ful inspection by eye. We then detect 365 sources
in the 0.5 -- 8.0 keV band image. In order to pick
up either soft (less absorbed) or hard (highly ab-
sorbed) sources more effectively, we also apply the
same detection algorithm to the 0.5 -- 2.0 keV (soft)
and 2.0 -- 8.0 keV (hard) band image. Then 17 and
16 new sources are found in the hard band and the
soft band image, respectively. In total, we detect
398 X-ray sources in the ACIS-I and ACIS-S FOV.
For each detected source, we calculate the X-ray
photon count (0.5 -- 8.0 keV) and the hardness ratio
(Table 1).
For a systematic study, we divide all the X-ray
sources into two groups, "bright" (> 200 counts)
and "faint" (≤ 200 counts) sources, according to
the photon counts inside the accumulation re-
gion. Out of 398 sources, 136 sources are "bright",
and 262 are "faint".
If the X-ray counts of the
"bright" sources are less than 3 times the back-
ground counts in the photon accumulation region,
we remove these sources and define the remain-
ing 123 sources to be "bright2". This screening is
practically necessary to comprise a good sample
for spectral and timing analyses.
In particular,
those at large off-axis angles have the background
counts comparable to 200 due to a rather large ac-
cumulation area, hence the background has a large
impact on the quality of the source spectrum and
light curve.
3.3. Correlation with 2MASS Sources
For all the detected X-ray sources, we search for
a NIR counterpart using the Point Source Cata-
log in the 2MASS Second Incremental Data Re-
lease4.
It covers the whole ACIS-I and ACIS-S
FOVs, and provides us with the NIR source po-
sitions and their photometric data in the J, H,
and K s-band down to 15.8, 15.1, and 14.3 mag,
respectively.
In the ACIS-I and ACIS-S FOVs, we find 638
2MASS sources. First, we search for the 2MASS
sources nearest to each CXO source within 3′′ ra-
dius. Second, we conversely search for the CXO
source nearest to each 2MASS source within 3′′
radius. Thus we pick up the nearest CXO-2MASS
pairs. The systematic position off-sets of the CXO
sources from their 2MASS counterparts is found
to be −0.′′186 and 0.′′200 in the direction of right
ascension and declination. After correcting these
systematic off-sets of the X-ray position, we re-
apply the same procedure for the 2MASS coun-
terpart search. Finally we find that 238 out of
398 (∼ 60%) X-ray sources have a 2MASS coun-
terpart. The distance between X-ray sources and
their counterparts is found to be ∼ 0.′′5.
Together with the X-ray properties, the J -- H
and H -- K colors of their 2MASS counterparts are
given in Table 1. About 80% of the "bright"
sources have a 2MASS counterpart (111 out of
136), while for the "faint" sources, the ratio is
about 50% (127 out of 262).
3.4. Timing Analysis
We make the X-ray light curve (background is
not subtracted) for the "bright2" sources and per-
form χ2 fit with a constant flux assumption. We
discriminate the time-variable sources by the sig-
nificance criteria of 0.01, which are marked with †
in Table 2. About 40% (47 out of 123) of the sam-
ple sources are found to be time-variable. Most
of them show flare-like variability, a fast rise and
slow decay of the flux. Some light curves show
multiple flares during the observation. A variety
2see http://asc.harvard.edu/udocs/reprocessing.html
3see http://asc.harvard.edu/ciao/caveats/acis pi.html
4see http://www.ipac.caltech.edu/2mass/
3
of features are seen in the light curves, and it is
difficult to separate X-ray photons during the flare
and at quiescence in a coherent manner. We there-
fore deal with them in the same way.
3.5. Spectral Analysis
We next perform spectral analysis of
the
"bright2" sources, the same data set for the tim-
ing analysis. We combine each energy bin so as to
have more than 20 photons, then subtract a back-
ground spectrum. We use the sherpa program for
the spectral fitting.
First we fit the spectrum of all the sample
sources with a thin thermal plasma and a power-
law model, both with absorption of hydrogen col-
umn density (NH). The free parameters of the
former model are temperature (kBT ), metallic-
ity (Z), and normalization, while the latter's are
photon index (Γ) and normalization. The for-
mer model is accepted for 90 out of 123 sources
with the upper probability of larger than 0.01
(99% confidence). The latter model is accepted
for 64 sources, of which all except two are also ac-
cepted in the former model. On the contrary, both
the models are rejected for the other 31 sources.
Therefore, here and after, we use the results of the
thin-thermal plasma model fittings (Table 2).
Second, for the 31 sources which reject both
models, we try 2-component models, 1) a thin-
thermal plasma + power-law and 2) 2-temperature
thin-thermal plasma model. Additional free pa-
rameters are kBT (or Γ) and normalization. Both
models are acceptable for 8 and rejected for 6
sources out of 31 samples, while the latter model
are accepted by other 15 sources. We therefore use
the results of the two temperature thin-thermal
plasma model fittings (Table 3).
4. DISCUSSION
4.1. The Nature of X-ray and NIR Sources
4.1.1. The Nature of NIR Sources
Since some fraction of the 2MASS sources may
be background or foreground sources, we first es-
timate the contribution of background galaxies in
the following way. The number counts of galax-
ies per square degree per magnitude at a certain
K -band (2.2µm) magnitude is given by
dN
dK
= 4000 × 10α(K−17)
(1)
where α = 0.67 for 10 mag< K < 17 mag (Toku-
naga 2000). Therefore, the number in the range
of Kmin mag< K < Kmax mag is
Z Kmax
Kmin
dN
dK
dK =
4000
α log 10{10α(Kmax−17)−10α(Kmin−17)}.
(2)
The NIR counterparts of the CXO sources have a
K -band flux of 6 mag < Ks < 15 mag. We sub-
stitute Kmin = 6 and Kmax = 15 for simplicity,
though equation (2) is valid only for Kmin >10
mag. Still, this gives us a good estimate, because
the second term of the right hand side of equa-
tion (2) is negligible compared to the first term
in this case. Considering our FOV (8.′4 × 8.′4 for
each chip), the estimated number of galaxies with
6 mag< Ks < 15 mag is ∼12. This is only 1.9%
of the 2MASS source number in our FOV. In ad-
dition, the background galaxies in this direction
may suffer significant extinction due to molecu-
lar clouds, hence the contribution of extragalac-
tic sources to our 2MASS sources should be even
smaller than that estimated above.
We can also neglect the contribution of fore-
ground sources to our 2MASS sample, because
OMC-2 and OMC-3 lie off the galactic plane by
∼ 20 degrees. Thus, we can safely assume most of
the NIR sources to be cloud members (YSOs and
MS stars).
4.1.2. The Nature of X-ray Sources
Krishnamurthi et al. (2001) observed the core
of the Pleiades star cluster with the CXO for
36 ksec, and found that a significant fraction of
X-ray sources are likely to be AGNs. We there-
fore try to discriminate cloud members from ex-
tragalactic sources using the X-ray hardness ratio.
In Figure 2, the histogram of the hardness ra-
tio is given separately for X-ray sources with and
without a NIR counterpart. The hardness ratio
of the X-ray sources with a NIR counterpart has
the peak at −1.0 ∼ −0.8, while those without a
NIR counterpart has its peak at 0.2 ∼ 0.4. A
power-law spectrum of Γ =1.7 has the hardness
ratio of 0.1 ∼ 0.4 when absorbed with the column
density of NH=1 -- 2×1022cm−2, a typical value of
4
the cloud column density. This corresponds to the
peak of X-ray sources without a NIR counterpart.
On the contrary, a thin-thermal spectrum with
kBT = 1.2 keV, Z=0.20, and NH= 1021cm−2, typ-
ical values for class III and MS stars (see the fol-
lowing section), has the hardness ratio of ∼ −0.8.
This corresponds to the peak of the histogram of
X-ray sources with a NIR counterpart.
Therefore, X-ray sources with a NIR counter-
part are mostly cloud members, consistent with
the conclusion of the previous subsection. We thus
focus on the X-ray sources with a NIR counter-
part. X-ray sources without a NIR counterpart,
on the other hand, are probably AGNs, which are
not the main subject of this paper.
4.2. Classification
YSOs are observationally classified into class
0 through class III based on their spectral en-
ergy distribution (SED) from optical, NIR, mid in-
frared (MIR), and sub-mm bands (Andr´e & Mont-
merle 1994). Since no systematic catalog for wave-
lengths longer than MIR is published in both of
these clouds, we use the J -- H /H -- K color-color di-
agram of the 2MASS counterparts for the classi-
fication (Lada & Adams 1992; Strom, Kepner, &
Strom 1995).
Out of the 123 "bright2" sources, 108 are de-
tected either in the J, H, or K s-band, and the
remaining 15 are found in none of these bands.
Figure 3 shows the J -- H vs H -- K plots of the de-
tected X-ray sources.
Intrinsic colors for giants and dwarfs are taken
from Tokunaga (2000) with colors transformed
into the CIT system (Bessel & Brett 1988), while
CTTS locus is taken from Meyer, Calvet, & Hillen-
brand (1997). We assume the slope of the redden-
ing lines to be E(J -- H)/E(H -- K) = 1.63 (Martin
& Whittet 1990). All the 2MASS colors are also
translated into CIT color system with the transfor-
mation formula established by Carpenter (2001).
Based on the position in this diagram, we
classify the sources into three groups -- class I,
class II, and class III + MS -- in the following
way5. Class I and class II sources are character-
5Note that our classification based only on three NIR bands
may be less complete compared with the conventional clas-
sification scheme. The number of class I and class II sources
might be underestimated because the J-H/H-K diagram is
5
ized by H -- K color excess over J -- H color, originat-
ing from their disk emission, while class III and
MS sources are not. Therefore, sources between
the second and the third leftmost reddening lines,
and above the CTTS locus are either class I or
class II sources. Class I and class II sources in
this region can be distinguished from each other
by the amount of extinction, where class Is gen-
erally have higher extinction (typically J -- H >1.5)
than class IIs (Lada & Adams 1992; Strom, Kep-
ner, & Strom 1995). Therefore, sources with J --
H >1.5 are classified to be class I and J -- H ≤ 1.5
are class II. Sources between the leftmost and the
second leftmost reddening lines can be either class
I, class II, class III or MS stars.
In this region
again, we classify sources based on their amount
of extinction; sources with J -- H >1.5 are class I,
J -- H >0.8 are class II, and J -- H ≤0.8 are class III
+ MS. These criteria are based on the fact that,
in the Taurus-Auriga dark cloud complex, class
II sources rarely have larger extinction than J --
H >1.5 and class III sources rarely have larger ex-
tinction than J -- H >0.8 (Strom, Kepner, & Strom
1995). In addition, sources which lack 2MASS J
and/or H -band detection (the upper limit of the
H -band magnitude is given) are classified as class
I, because all of them have larger extinction of
H -- K >1.2 mag.
Then, out of the 108 sources, we find 19 class I,
18 class II, and 61 class III + MS sources. The
other 10 are outside of the classification regions.
The results are summarized in Table 2.
4.3. X-ray Properties of Different Classes
Based on the classification described in the pre-
vious section, we list the X-ray properties (absorp-
tion, metallicity, luminosity, and temperature) of
each class in Table 4 and compare them with each
other.
not very sensitive to the NIR excess (Olofsson et al. 1999;
Persi et al. 2000; Bontemps et al. 2001). On the other hand,
the number of class I and class II sources might be overesti-
mated because we can not discriminate mildly extinguished
class III + MS from them. Therefore it may be fair to use
the terminology "class I-like" instead of "class I" etc. For
simplicity, however, we use the latter terminology in this
paper.
4.3.1. Absorption
The X-ray absorption decreases as a YSO
evolves from class I through class III + MS. This
is basically consistent with the fact that we clas-
sify these sources using their NIR extinction. J -- H
derived from the NIR photometry is a good in-
dicator for the amount of material in solid-state,
while NH derived from X-ray spectroscopy repre-
sents the amount of gas. Therefore, NH should
go in proportion to J -- H, and the ratio gives us
the information of the dust-to-gas ratio in a star
forming cloud.
For all the "bright2" sources with NIR data, we
plot the relation between NH and J -- H (Figure 4).
A clear correlation is found. The best-fit linear
function is
NH/1022 cm−2 = (1.35±0.18){(J -- H)−(0.63±0.05)} mag.
(3)
The J -- H offset of 0.63±0.05 mag should be the
averaged intrinsic color after removing the redden-
ing, which corresponds to the spectral type K5 --
K7 in MS stars (Tokunaga 2000). This indicates
that low mass YSOs and MS stars are dominant in
these clouds. The slope of 1.35±0.18 gives us in-
formation on the dust-to-gas ratio. Together with
the relation between AV and J -- H given in Meyer
et al. (1997),
NH/1021 cm−2 = (1.49 ± 0.20) × AV mag
(4)
is derived. The slope of 1.49±0.20 is smaller than
those of the Galactic interstellar medium and the
ρ-Ophiuchi dark cloud, and is similar to that of
another star forming region, the Mon R2 cloud
(Predehl & Schmitt 1995; Imanishi et al. 2001;
Kohno, Koyama, & Hamaguchi 2001). Thus the
dust-to-gas ratio may scatter from cloud to cloud,
possibly with more massive star forming regions
having larger dust-to-gas ratio.
Padgett (1996) observed 30 G and K pre-main-
sequence stars in the nearby star forming regions
including Orion, and studied their photospheric
abundances using the iron absorption lines in the
optical band. All star forming regions are found
to have the solar abundance.
The discrepancy between the coronal abun-
dance and the photospheric abundance, which are
derived from the X-ray observations and the opti-
cal observations respectively, is often seen in other
samples Yamauchi et al. (1996); Imanishi et al.
(2001).
4.3.3. Luminosity and Temperature
The temperature (kBT ) and the luminosity
(LX ) are plotted on Figure 5, separately for class I,
class II, and class III + MS. The temperatures
of class I to class III + MS sources are ran-
domly distributed over a wide range of luminos-
ity, but the mean temperatures of class I and
class II sources are significantly higher (kBT ∼
3.0 keV) than that of class III + MS (kBT ∼
1.2 keV). We may argue that there are two types
of X-ray emission mechanisms; one exhibits higher
temperature, and the other has lower tempera-
ture. The former dominates in less evolved YSOs,
like class I and class II sources, while the latter
gradually appears as YSOs evolve to class III +
MS. In this context,
it is suggestive that most
sources with two temperature components (Ta-
ble 4) belong to class III + MS. The higher tem-
perature component of these sources has ∼ 2.3 keV
on average similar to the 1-temperature sources
of class I and class II, while the mean of the
lower temperature component is ∼ 0.8 keV sim-
ilar to the 1-temperature sources of class III +
MS. We hence suspect that the two component
class III + MS sources may be in the transition
phase between higher-temperature-dominant and
the lower-temperature-dominant stage.
4.3.2. Metallicity
4.3.4. Time Variation
We determine the metallicity for many sample
stars in a star forming region for the first time. A
hint of a decreasing trend toward evolved classes
is seen, although this depends on the way the sam-
ples are broken into class I, class II, and class III
+ MS. The metallicity, when all classes are com-
bined, is 0.45 (0.37 -- 0.52) solar.
Class II sources have a slightly higher fraction
of time-variable sources than other classes, though
it is not statistically significant. When class I
and class II are combined and are compared with
class III + MS, we see no significant difference in
time-variation rate (in both groups, ∼ 40% are
time-variable). Imanishi et al. (2001) also argues
6
phase, possibly early class III, plasma emis-
sions with different temperatures coexist.
7. The ratio of time-variable sources is nearly
the same among different classes. Whether
the time-variability is related to higher-
temperature- or lower-temperature-plasma
is still an open question.
The authors express their thanks to Dr. Yoshit-
omo Maeda for giving us invaluable information on
the analysis of the ACIS data, especially on the
source detection algorithm and detector calibra-
tions. The authors also acknowledge the careful
reading of the manuscript by Jun Yokogawa and
Leisa Townsley. M.T. and M.G. are financially
supported by the Japan Society for the Promotion
of Science.
that no significant difference of the flare rate is
seen among three classes in the ρ-Ophiuchi dark
cloud. Whether time-variable activity is related to
the higher-temperature or the lower-temperature-
component is not clear in our data set.
5. SUMMARY
We observed OMC-2 and OMC-3 with the
CXO/ACIS for ∼ 100 ksec. This is one of the
deepest observations ever performed in star form-
ing regions in the X-ray band. We detected ∼ 400
X-ray sources in our FOV. Coherent analyses on
these sources derived the following results.
1. Imaging analysis is performed for all the de-
tected sources, and their position, photon
counts, and hardness ratio are derived. This
is one of the largest catalogs of X-ray sources
in a star forming region.
2. Correlations with the 2MASS sources are
found using the 2MASS database. About
60% of the X-ray sources have a NIR coun-
terpart.
3. Spectral and timing analysis are performed
for "bright2" X-ray sources. A one tem-
perature thin-thermal plasma model can ex-
plain most of the spectra. About 40% of
the "bright2" sources are found to be time-
variable.
4. Most of the X-ray sources with a NIR coun-
terpart are likely cloud members, while those
with no NIR counterpart are probably back-
ground AGNs.
5. We classify the cloud members to be class
I, class II, and class III + MS based on the
J -- H /H -- K color-color diagram and conduct
a systematic comparison on the X-ray prop-
erties, such as absorption, luminosity, tem-
perature, and time-variation.
6. Class I and class II sources are found to have
higher temperatures than class III + MS. We
thus propose that two types of X-ray emis-
sion mechanisms exist (higher temperature
and lower temperature component). The
higher temperature plasma appears in the
earlier phase and lower temperature plasma
appears as YSOs evolve.
In the transition
7
REFERENCES
Persi, P. et al. 2000, A&A, 357, 219
Andr´e, P. & Montmerle, T. 1994, ApJ, 420, 837
Predehl, P & Schmitt, J. H. M. M. 1995, A&A,
Bessel, M. S. & Brett, J. M. 1988, PASP, 100,
1134
Bontemps, S. et al. 2001, A&A, 372, 173
Carpenter, J. M. 2001, AJ, 121, 2851
Chini, R., Reipurth, B., Ward-Thompson, D.,
Bally, J., Nyman, L-A, Sievers, A., & Billawala,
Y. 1997, ApJ, 474, L135
Feigelson, E. D. & DeCampli 1981, ApJ, 243, L89
Garmire, G. P., et al., 2001, in preparation
Grosso, N., Montmerle, T., Bontempts, S., Andr´e,
P., & Feigelson, E. D. A&A, 359, 113
Imanishi, K., Koyama, K., & Tsuboi, Y. 2001,
ApJ, in press
Kamata, Y., Koyama, K., Tsuboi, Y., & Ya-
mauchi, S. 1997, PASJ, 49, 461
293, 889
Schulz, N. S., Canizares, C., Huenemoerder, D.,
Kastner, J. H., Tayler, S. C., & Bergstorm, E.
J. ApJ, 549, 441
Skinner, S. L. & Walter, F. M. 1998, ApJ, 509,
761
Strom, K. M., Kepner, J., & Strom, S. E. 1995,
ApJ, 438, 813
Tokunaga, A. T. 2000, in Allen's Astrophysical
Quantities, ed. A. N. Cox, (4th ed.; New York:
Springer-Verlag), 143
Tsuboi, Y., Imanishi, K., Koyama, K., Grosso, N.,
& Montmerle, T. 2000, ApJ, 532, 1089
Tsuboi, Y., Koyama, K., Hamaguchi, K., Tatem-
atsu, K., Sekimoto, Y., Bally, J., & Reipurth,
B. 2001, ApJ, 554, in press
Weisskopf, M. C., O'dell, S., & van Speybroeck,
Kohno, M., Koyama, K., & Hamaguchi, K. 2001,
L. P. 1996, Proc. of SPIE, 2805, 2
in preparation
Yamauchi, S., Koyama, K., Sakano, M., & Okada,
Koyama, K., Hamaguchi, K., Ueno, S., &
K. 1996, PASJ, 48, 719
Kobayashi, N. 1996, PASJ, 48, L87
Krishnamurthi, A., Reynolds, C. S., Linsky, J.L.,
Martin, E., & Gagn´e, M. 2001, AJ, 11, 337
Lada, C. J. & Adams, F. C. 1992, ApJ, 393, 278
Martin, P. G. & Whittet, D. C. B. 1990, ApJ, 357,
113
Meyer, M. R., Calvet, N., & Hillenbrand, L. A.
1997, AJ, 114, 288
Montmerle, T., Koch-Miramond, L., Falgarone,
E., & Grindlay, J, E. 1983, ApJ, 269, 182
Montmerle, T., Grosso, N., Tsuboi, Y., &
Koyama, K. 2000, ApJ, 532, 1097
Morrison, R. & McCammon, D. 1983, ApJ, 270,
119
Oloffson, G. et al. 1999, A&A, 350, 883
Padgett, D. L. 1996, ApJ, 471, 847
This 2-column preprint was prepared with the AAS LATEX
macros v5.0.
8
class I
class II
class III & MS
AV = 5 mag
Fig. 3. -- The J -- H /H -- K color-color diagram in
CIT color system.
"Bright2" sources with the
2MASS J, H, and K s-band detection are plotted
(crosses). The typical error is ±0.05 mag. The
evolution tracks of giants and dwarfs, and CTTS
locus are given by solid lines, while the dotted lines
represent the extinction vector. X-ray sources are
classified into class I, class II, and class III + MS
sources based on their position on the diagram.
Fig. 1. -- FOVs of ACIS-I (large square region)
and ACIS-S2 (small square region at the top left)
are overlaid on the 1300mm intensity map (Chini
et al. 1997). ACIS-I covers the whole OMC-2 and
OMC-3 star forming regions. Squares and cir-
cles are X-ray sources with and without a 2MASS
counterpart, respectively. The coordinate is in the
equinox 1950B.
Fig. 2. -- The histogram of the hardness ratio
(from -- 1 to 1 with a step of 0.2) is shown by the
dashed and dotted line for the X-ray sources with
and without a NIR counterpart, respectively. The
total is given by the solid line.
Fig. 4. -- The NIR extinction (J -- H mag) and
the X-ray absorption (NH 1022cm−2) are plotted.
The best-fit linear function is expressed with the
dashed line.
9
Fig. 5. -- Luminosity (1029 erg s−1) in logarith-
mic scale and temperature (keV) of each source
are plotted separately for class I (squares), class II
(triangles), and class III + MS (crosses) sources.
10
Results of Imaging Analysis and NIR Identificationa
Table 1
IDb
Namec
X-rayd Hardnesse
Counts
Ratio
J -- H f
(mag)
H -- K f
(mag)
I1
I2
I3
I4
I5
I6
I7
I8
I9
I10
I11
I12
I13
I14
I15
I16
I17
I18
I19
I20
I21
I22
I23
I24
I25
I26
I27
I28
I29
I30
I31
I32
I33
I34
I35
I36
I37
I38
I39
I40
I41
I42
I43
I44
I45
I46
I47
I48
I49
I50
I51
I52
I53
I54
I55
I56
I57
I58
I59
I60
I61
I62
I63
I64
053438.60-050844.6
053440.86-050700.0
053443.10-050919.8
053444.89-050650.9
053445.19-051046.9
053448.29-050714.9
053448.37-050502.9
053449.18-050440.0
053450.33-050639.9
053450.43-051113.0
053450.59-050651.6
053450.64-050409.4
053453.03-050328.7
053453.42-051030.0
053453.45-050132.0
053453.71-050550.7
053455.38-050141.1
053455.66-050846.1
053455.76-050744.3
053456.07-050057.4
053456.13-050603.3
053456.64-050627.9
053456.67-050439.7
053456.71-051045.4
053456.82-051135.3
053457.35-050605.8
053458.19-050929.7
053458.21-051156.4
053458.52-051228.8
053459.22-050128.0
053459.32-050531.8
053459.47-050827.4
053500.40-050946.6
053500.87-050941.2
053501.37-051107.9
053501.43-050934.6
053501.98-050045.0
053502.16-050440.3
053502.52-050411.8
053502.56-050500.3
053502.79-050305.7
053502.79-050003.7
053502.85-050202.4
053503.12-050919.1
053503.30-050449.9
053503.33-051308.9
053503.41-050542.2
053503.81-050520.9
053503.99-050142.9
053504.23-050844.3
053504.30-050814.7
053504.45-050738.3
053504.55-045831.7
053504.64-050957.9
053505.24-045925.0
053505.65-045855.4
053505.66-050455.2
053505.76-051136.3
053505.95-050839.3
053506.09-051218.8
053506.36-050901.8
053506.49-045842.2
053506.58-050717.8
053506.61-050453.0
255
63
368
403
176
535
549
127
70
773
110
361
1747
579
49
145
534
39
103
359
276
55
98
48
2286
19
558
258
135
28
740
30
213
135
60
670
63
83
11
9
20
1023
20
90
26
159
1771
17
20
54
25402
14
109
437
26
334
7
92
18
43
10
58
15
42
11
0.31
-0.02
0.62
-0.54
-0.25
-0.66
-0.77
-0.97
-0.29
-0.75
0.40
-0.70
-0.73
-0.70
0.10
-0.75
-0.83
0.38
0.24
-0.69
-0.82
0.20
0.47
-0.33
-0.82
0.58
-0.71
-0.64
0.04
0.43
-0.75
0.27
0.13
-0.72
0.33
-0.85
-0.46
-0.76
0.82
1.00
-0.40
-0.57
0.50
-0.76
0.69
0.17
-0.62
0.53
0.70
0.26
-0.55
-0.43
-0.21
-0.78
0.54
-0.83
0.43
-0.28
-0.67
-0.30
0.20
-0.45
0.60
0.43
· · ·
· · ·
· · ·
0.725
0.748
0.658
0.675
0.695
· · ·
0.663
· · ·
0.720
0.751
0.628
0.625
0.644
0.731
· · ·
· · ·
0.639
0.613
· · ·
· · ·
0.733
0.713
· · ·
0.762
0.604
0.552
· · ·
0.767
· · ·
0.567
0.559
0.962
0.668
0.631
0.757
· · ·
· · ·
0.576
1.166
· · ·
0.718
· · ·
· · ·
0.216
· · ·
· · ·
· · ·
· · ·
0.379
0.517
0.223
0.188
0.227
· · ·
0.172
· · ·
0.289
0.451
0.232
0.210
0.247
0.146
· · ·
· · ·
0.213
0.238
· · ·
· · ·
0.301
0.467
· · ·
0.283
0.219
0.236
· · ·
0.313
· · ·
0.310
0.318
0.575
0.180
0.219
0.374
· · ·
· · ·
-0.065
0.453
· · ·
0.335
· · ·
· · ·
0.027
· · ·
· · ·
· · ·
· · ·
0.656
· · ·
1.055
0.647
· · ·
0.649
· · ·
1.394
1.084
-0.077
· · ·
0.556
· · ·
· · ·
0.153
· · ·
0.743
0.257
· · ·
0.236
· · ·
0.658
0.496
-0.023
· · ·
0.301
· · ·
· · ·
· · ·
Table 1 -- Continued
IDb
Namec
X-rayd Hardnesse
Counts
Ratio
J -- H f
(mag)
H -- K f
(mag)
-0.19
-0.76
-0.21
0.14
-0.85
-0.67
-0.83
0.39
-0.56
0.24
0.00
0.54
0.52
0.75
-0.09
-0.46
-0.95
0.48
-0.83
0.38
-0.74
0.71
0.11
0.81
0.01
0.00
-0.65
-0.82
-0.85
-0.76
1.00
0.16
0.58
0.71
0.21
0.50
-0.86
-0.53
-0.31
-0.53
-0.33
-0.30
-0.47
0.06
-0.92
0.20
-0.91
-0.76
-0.79
-0.80
0.50
-0.91
-0.62
-0.54
0.80
-0.66
0.20
-1.00
-0.17
-0.14
-0.40
-1.00
-0.14
0.93
1.439
0.671
1.066
· · ·
0.742
0.615
0.672
· · ·
0.680
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.524
0.590
· · ·
0.558
· · ·
0.683
· · ·
2.364
· · ·
2.401
1.941
1.156
· · ·
0.676
0.572
· · ·
· · ·
· · ·
· · ·
2.294
· · ·
0.631
· · ·
0.677
1.093
0.590
1.899
0.767
· · ·
0.579
· · ·
0.622
0.561
0.605
0.655
· · ·
0.490
0.995
· · ·
· · ·
· · ·
· · ·
0.578
2.217
· · ·
· · ·
0.041
· · ·
· · ·
0.618
0.236
0.584
· · ·
0.356
0.095
0.206
· · ·
0.348
1.612
· · ·
· · ·
3.286
· · ·
· · ·
0.250
0.279
· · ·
0.328
· · ·
0.189
· · ·
1.164
· · ·
1.504
1.264
0.653
· · ·
0.183
0.282
· · ·
1.216
· · ·
· · ·
1.198
· · ·
0.256
· · ·
0.337
0.525
0.277
1.092
0.315
· · ·
0.339
· · ·
0.234
0.329
0.161
0.197
· · ·
0.341
0.540
· · ·
· · ·
· · ·
1.775
0.299
1.373
· · ·
· · ·
0.013
1.263
· · ·
I65
I66
I67
I68
I69
I70
I71
I72
I73
I74
I75
I76
I77
I78
I79
I80
I81
I82
I83
I84
I85
I86
I87
I88
I89
I90
I91
I92
I93
I94
I95
I96
I97
I98
I99
I100
I101
I102
I103
I104
I105
I106
I107
I108
I109
I110
I111
I112
I113
I114
I115
I116
I117
I118
I119
I120
I121
I122
I123
I124
I125
I126
I127
I128
053506.71-051147.2
053506.83-051040.9
053507.54-051116.5
053508.04-050120.0
053508.75-050442.3
053508.96-051027.4
053509.15-050648.9
053509.32-045934.0
053509.42-045943.0
053509.85-045850.9
053510.14-051342.4
053510.32-051112.9
053510.51-045847.3
053511.56-050210.0
053511.81-051401.8
053512.72-051202.6
053512.80-050147.9
053512.92-045557.4
053512.95-050210.3
053513.00-050028.0
053513.34-050921.7
053513.35-050416.3
053513.88-045805.0
053513.99-050740.6
053514.28-051428.1
053514.29-051342.1
053514.63-050627.1
053514.64-050226.5
053514.66-050314.2
053514.67-050854.1
053514.70-050828.0
053514.87-050650.6
053514.94-050120.1
053514.99-050715.0
053515.08-050655.4
053515.18-050758.2
053515.27-050034.4
053515.36-051342.0
053515.49-050114.0
053515.55-050145.3
053515.59-050934.0
053515.61-045929.1
053515.65-045714.5
053515.76-051229.2
053515.80-050327.8
053515.93-051501.3
053516.14-050921.1
053516.35-050438.1
053516.40-045803.7
053516.49-050332.0
053516.66-050856.2
053516.80-050729.2
053516.86-050749.5
053517.12-051240.9
053517.25-050317.7
053517.36-051231.1
053517.41-045958.4
053517.46-050950.9
053517.72-050032.1
053517.92-051535.0
053518.24-051308.9
053518.24-050355.9
053518.26-050806.9
053518.31-050034.8
257
200
1232
14
26
12
1192
56
72
442
86
13
835
16
237
59
79
325
109
29
1088
7
207
21
268
116
34
312
574
33
4
333
24
62
112
4
29
68
26
34
21
151
159
72
177
3013
88
17
264
294
4
45
16
74
10
65
259
44
29
830
561
92
7
60
12
Table 1 -- Continued
IDb
Namec
X-rayd Hardnesse
Counts
Ratio
J -- H f
(mag)
H -- K f
(mag)
-0.73
-0.71
-0.48
0.65
0.85
1.00
0.83
-0.82
0.67
0.67
0.25
0.83
0.03
0.86
0.06
0.05
-0.86
0.43
-0.12
-0.37
0.63
-0.81
-0.63
-0.79
-0.62
0.13
0.84
-0.64
0.08
-0.71
-0.20
0.25
-0.69
-0.32
0.65
0.02
0.21
-0.92
-0.34
-0.69
-0.08
0.04
-0.38
-0.63
0.19
-0.82
0.76
-0.25
0.58
-0.25
-0.81
0.56
0.68
-0.85
0.82
0.00
-0.06
0.94
-0.86
0.50
-0.83
1.00
0.10
-0.64
0.595
0.583
· · ·
· · ·
· · ·
· · ·
· · ·
0.647
· · ·
· · ·
1.610
· · ·
· · ·
2.063
· · ·
1.923
0.765
· · ·
1.177
1.118
· · ·
0.547
0.604
0.686
· · ·
· · ·
0.837
0.784
· · ·
0.950
· · ·
· · ·
0.747
· · ·
· · ·
2.624
1.396
0.583
1.412
0.903
· · ·
1.756
· · ·
· · ·
· · ·
0.628
· · ·
0.235
0.201
· · ·
· · ·
· · ·
· · ·
· · ·
0.307
· · ·
· · ·
1.034
· · ·
· · ·
1.310
· · ·
1.308
0.203
2.444
0.558
0.463
· · ·
0.489
0.176
0.490
· · ·
2.056
0.715
0.528
· · ·
0.558
· · ·
· · ·
0.226
· · ·
· · ·
1.649
0.873
0.124
0.531
0.519
· · ·
0.742
1.935
· · ·
2.256
0.250
· · ·
· · ·
· · ·
· · ·
1.231
0.767
· · ·
2.994
0.629
1.963
· · ·
· · ·
0.672
0.505
1.670
1.855
0.321
1.867
· · ·
· · ·
· · ·
· · ·
0.689
· · ·
0.600
· · ·
· · ·
0.659
· · ·
0.227
· · ·
0.128
· · ·
· · ·
0.888
I129
I130
I131
I132
I133
I134
I135
I136
I137
I138
I139
I140
I141
I142
I143
I144
I145
I146
I147
I148
I149
I150
I151
I152
I153
I154
I155
I156
I157
I158
I159
I160
I161
I162
I163
I164
I165
I166
I167
I168
I169
I170
I171
I172
I173
I174
I175
I176
I177
I178
I179
I180
I181
I182
I183
I184
I185
I186
I187
I188
I189
I190
I191
I192
053518.45-050832.7
053518.61-045944.0
053518.84-051448.1
053518.93-050051.9
053518.94-050638.1
053518.96-050324.9
053518.98-050031.0
053519.65-050230.4
053519.75-050533.4
053519.75-051536.7
053519.86-051510.2
053519.97-050104.1
053519.98-051405.6
053519.99-051252.9
053520.08-051318.8
053520.14-051317.3
053520.37-050228.2
053520.59-050302.0
053520.74-051551.3
053520.76-045835.6
053521.13-050634.2
053521.26-050918.2
053521.31-051214.9
053521.39-050944.2
053521.44-050905.6
053521.50-050155.3
053521.56-050941.0
053521.57-050951.7
053521.86-045650.5
053521.87-050703.7
053521.91-045919.5
053521.92-050352.4
053521.94-051430.1
053522.11-051507.7
053522.34-045954.7
053522.35-050741.0
053522.40-050807.0
053522.45-050913.0
053522.55-050802.6
053522.67-051413.8
053523.13-050037.7
053523.20-051345.6
053523.22-050845.8
053523.33-045722.3
053523.33-050823.5
053523.44-051053.9
053523.51-050715.6
053523.53-051525.9
053524.34-050122.2
053524.58-051131.5
053524.61-051200.7
053524.66-050929.4
053524.87-050623.1
053525.03-050911.4
053525.07-051025.3
053525.17-050510.9
053525.17-051540.6
053525.22-050825.6
053525.24-050929.4
053525.36-050722.1
053525.40-051050.1
053525.44-050654.1
053525.53-045724.6
053525.64-050759.2
15
165
661
23
26
7
24
11
6
359
309
259
202
97
1099
953
73
7
2276
673
109
2202
14667
29
142
97
76
260
93
838
15
8
345
479
23
507
833
415
921
206
13
164
13
182
59
1909
33
96
277
83
363
9
144
26
22
12
162
66
544
4
762
6
145
324
13
Table 1 -- Continued
IDb
Namec
X-rayd Hardnesse
Counts
Ratio
J -- H f
(mag)
H -- K f
(mag)
-1.00
-0.68
0.11
1.00
-0.79
-0.76
1.00
-0.84
-0.06
0.20
1.00
0.13
0.67
-0.08
-0.89
0.16
0.42
0.60
-0.46
0.57
0.42
-0.41
0.41
-0.79
-0.28
0.61
0.03
0.29
0.50
0.82
-0.82
0.67
-0.53
-0.63
-0.78
1.00
-0.61
-0.74
-0.27
-0.10
-0.08
-0.34
0.33
0.25
-0.93
-0.80
-0.41
-0.64
0.91
-0.78
-0.80
0.85
-0.28
0.49
0.39
-0.65
-0.14
-0.53
0.29
-0.45
0.76
-0.71
0.65
-0.39
0.646
0.615
· · ·
· · ·
1.027
0.586
· · ·
0.504
1.455
· · ·
· · ·
1.039
· · ·
0.887
0.768
· · ·
2.435
1.285
· · ·
· · ·
· · ·
0.997
· · ·
0.651
1.059
· · ·
· · ·
2.581
· · ·
· · ·
0.582
· · ·
0.864
0.780
0.646
· · ·
0.705
0.602
1.180
1.612
· · ·
2.291
· · ·
· · ·
0.571
0.660
0.620
0.777
· · ·
0.226
0.688
2.003
1.168
2.628
· · ·
0.812
1.046
0.746
2.176
0.982
· · ·
0.599
· · ·
1.093
0.283
0.335
· · ·
· · ·
· · ·
· · ·
· · ·
0.224
0.764
1.206
· · ·
0.766
· · ·
0.512
0.453
2.101
1.499
0.657
· · ·
· · ·
2.365
0.662
· · ·
0.221
0.515
2.812
2.158
1.501
· · ·
· · ·
0.171
· · ·
0.257
0.311
0.200
· · ·
0.487
0.266
0.607
0.849
· · ·
1.522
· · ·
1.693
0.115
0.231
0.299
0.315
· · ·
0.050
0.477
1.424
0.504
1.639
· · ·
0.414
0.618
0.484
1.141
0.356
· · ·
0.193
· · ·
0.302
I193
I194
I195
I196
I197
I198
I199
I200
I201
I202
I203
I204
I205
I206
I207
I208
I209
I210
I211
I212
I213
I214
I215
I216
I217
I218
I219
I220
I221
I222
I223
I224
I225
I226
I227
I228
I229
I230
I231
I232
I233
I234
I235
I236
I237
I238
I239
I240
I241
I242
I243
I244
I245
I246
I247
I248
I249
I250
I251
I252
I253
I254
I255
I256
053525.72-050705.2
053525.73-050951.5
053525.77-050559.7
053525.85-050247.0
053525.86-050758.3
053526.05-050839.7
053526.27-050742.9
053526.29-050841.9
053526.34-051613.8
053526.47-045953.8
053526.66-050611.9
053526.85-051109.7
053527.07-050623.5
053527.15-051547.1
053527.41-050905.8
053527.45-050243.9
053527.63-050938.6
053527.73-051357.4
053527.79-051704.7
053527.83-050537.9
053528.06-050136.6
053528.14-051015.3
053528.18-050342.7
053528.19-050051.0
053528.21-051139.0
053528.27-045839.8
053528.50-050748.6
053528.60-050546.0
053528.67-050307.9
053528.71-050553.0
053529.02-050605.6
053529.25-050807.1
053529.45-051635.3
053529.81-051609.4
053529.89-051212.4
053529.90-050428.5
053530.00-051229.5
053530.11-050911.3
053530.13-051420.8
053530.32-051353.8
053530.55-050335.7
053530.64-045937.4
053530.72-050647.1
053530.86-045814.2
053531.06-050416.0
053531.22-051230.1
053531.27-051203.6
053531.31-051535.1
053531.46-050547.3
053531.48-051605.0
053531.49-050503.2
053531.57-050548.9
053531.58-051658.2
053531.61-050015.8
053531.91-050551.1
053531.96-050929.7
053532.02-050807.0
053532.24-051200.1
053532.31-051145.8
053532.35-051428.1
053532.35-050831.6
053532.51-050211.7
053532.80-050753.4
053532.86-051606.4
5
1411
9
5
878
83
8
3840
231
88
11
289
12
196
18
270
24
55
711
345
326
37
41
172
50
1228
89
28
8
11
1424
6
1189
476
136
8
157
23
148
213
13
222
6
67
91
311
82
2773
22
660
187
13
181
63
33
2062
7
38
175
150
25
14
23
1143
14
Table 1 -- Continued
IDb
Namec
X-rayd Hardnesse
Counts
Ratio
J -- H f
(mag)
H -- K f
(mag)
0.12
-0.63
-0.14
0.82
-0.11
-0.83
-0.35
0.88
-0.80
0.89
-0.61
0.48
-0.67
0.33
1.00
-0.72
0.93
-0.90
0.91
-0.55
-0.26
-0.61
0.60
-0.46
0.33
-0.92
-0.67
-0.45
0.60
-0.33
-0.88
0.12
-0.64
0.48
0.38
-0.42
-0.31
0.12
-0.72
-0.55
-0.26
-0.84
-0.80
0.14
0.59
0.35
-0.68
0.33
0.58
-0.13
0.11
-0.61
-0.76
-0.58
-0.74
-0.33
-0.32
-0.73
0.32
0.34
-0.79
-0.33
-0.79
-0.08
2.036
0.619
0.783
· · ·
· · ·
0.652
1.451
1.934
0.586
· · ·
0.583
2.309
0.556
· · ·
· · ·
0.577
· · ·
0.628
· · ·
0.517
0.678
1.551
1.349
0.836
· · ·
0.654
0.746
0.796
· · ·
0.508
0.677
· · ·
0.668
· · ·
· · ·
0.750
· · ·
2.532
0.578
1.936
1.110
0.559
0.664
· · ·
1.154
0.271
0.399
· · ·
· · ·
0.200
0.693
1.264
0.264
· · ·
0.161
2.173
0.550
· · ·
· · ·
0.217
· · ·
0.208
· · ·
0.351
0.284
1.042
0.722
0.467
· · ·
0.193
0.315
0.557
· · ·
0.432
0.268
1.505
0.283
· · ·
· · ·
0.223
· · ·
1.365
0.303
1.035
0.661
0.291
0.163
· · ·
· · ·
· · ·
· · ·
1.134
· · ·
· · ·
0.690
0.654
0.903
0.619
1.173
0.612
1.702
1.329
0.559
· · ·
· · ·
0.704
0.644
0.682
0.874
· · ·
0.606
· · ·
· · ·
0.270
0.289
0.305
0.302
0.491
0.198
0.827
0.617
0.288
· · ·
· · ·
0.158
0.274
0.169
0.689
I257
I258
I259
I260
I261
I262
I263
I264
I265
I266
I267
I268
I269
I270
I271
I272
I273
I274
I275
I276
I277
I278
I279
I280
I281
I282
I283
I284
I285
I286
I287
I288
I289
I290
I291
I292
I293
I294
I295
I296
I297
I298
I299
I300
I301
I302
I303
I304
I305
I306
I307
I308
I309
I310
I311
I312
I313
I314
I315
I316
I317
I318
I319
I320
053532.97-051207.7
053533.00-051734.1
053533.16-051412.9
053533.38-050803.8
053533.52-051521.1
053533.61-050043.5
053533.64-050309.5
053533.83-050429.0
053533.86-050907.4
053534.26-045954.5
053534.48-051309.1
053534.52-050053.6
053534.68-051555.1
053535.11-051449.0
053535.12-050600.0
053535.36-051113.6
053535.48-050722.0
053535.55-050700.3
053535.63-050150.3
053535.64-051051.8
053535.71-051545.1
053536.02-051226.9
053536.19-050457.6
053536.40-050117.0
053536.51-050524.1
053536.56-050440.9
053536.69-050415.8
053536.77-051000.9
053536.86-050734.2
053536.95-050527.6
053537.17-051031.4
053537.57-050449.2
053537.70-050633.6
053538.04-051510.1
053538.07-050318.9
053538.30-051420.9
053538.52-045942.0
053538.57-050805.2
053538.65-050958.6
053538.75-050457.1
053538.88-051244.0
053539.04-050705.8
053539.08-050858.2
053539.17-050541.8
053539.39-050509.5
053539.72-045920.2
053539.97-050638.3
053540.19-050429.1
053540.47-050420.5
053540.59-051221.6
053540.66-051111.8
053540.80-050902.9
053541.04-050626.7
053541.38-050354.4
053541.72-050330.1
053541.81-050520.9
053542.04-051013.8
053542.04-051301.8
053542.40-051104.3
053542.64-045957.2
053542.78-051156.9
053542.98-050307.1
053543.27-050918.6
053543.54-050542.0
25
406
974
11
646
128
223
346
20
53
76
27
330
99
13
506
27
210
66
22
204
270
114
943
9
909
2221
11
10
24
312
16
298
122
13
246
518
75
320
40
1047
279
198
7
113
65
75
9
19
85
76
825
119
38
61
15
660
617
56
79
658
42
2262
48
15
Table 1 -- Continued
IDb
Namec
X-rayd Hardnesse
Counts
Ratio
J -- H f
(mag)
H -- K f
(mag)
-0.67
-0.87
0.24
-0.70
0.51
-0.76
0.41
-0.81
0.07
0.55
-0.60
-0.45
0.10
-0.49
0.55
0.39
-0.33
0.27
-0.78
0.46
0.25
-0.73
-0.75
-0.07
0.34
-0.81
-0.03
-0.56
0.24
0.14
-0.27
0.06
-0.27
0.10
0.67
1.00
0.47
1.00
1.00
0.47
1.00
1.00
1.00
0.63
1.00
1.00
0.26
0.57
0.45
0.30
0.00
-0.75
-1.00
-1.00
-1.00
0.00
0.18
-1.00
-0.07
-0.33
-0.50
-0.60
-0.55
-0.29
0.597
0.705
· · ·
0.630
· · ·
0.776
· · ·
0.542
· · ·
· · ·
0.928
0.968
· · ·
0.771
· · ·
· · ·
· · ·
· · ·
0.736
· · ·
· · ·
0.625
0.579
1.037
· · ·
0.867
0.654
0.554
· · ·
· · ·
0.574
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.268
0.200
· · ·
0.247
· · ·
0.407
· · ·
0.212
· · ·
· · ·
0.680
0.271
· · ·
0.355
· · ·
· · ·
· · ·
· · ·
0.372
· · ·
· · ·
0.143
0.373
0.347
· · ·
0.489
0.204
0.301
· · ·
· · ·
0.406
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.410
· · ·
· · ·
0.346
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.608
· · ·
0.754
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0.314
· · ·
0.347
· · ·
I321
I322
I323
I324
I325
I326
I327
I328
I329
I330
I331
I332
I333
I334
I335
I336
I337
I338
I339
I340
I341
I342
I343
I344
I345
I346
I347
I348
I349
I350
I351
I352
I353
I354
I355
I356
I357
I358
I359
I360
I361
I362
I363
I364
I365
I366
I367
I368
I369
I370
I371
I372
I373
I374
I375
I376
I377
I378
I379
I380
I381
I382
I383
I384
053543.82-051001.6
053544.52-050733.2
053544.54-050006.7
053544.88-050718.5
053545.70-050646.5
053546.10-051054.2
053547.25-050251.9
053547.77-051033.0
053548.26-051117.3
053548.39-050522.4
053548.40-050129.7
053548.88-050033.4
053549.00-050544.0
053549.04-050142.4
053549.21-050225.0
053549.96-050426.2
053550.07-051024.0
053550.51-050226.5
053551.12-050710.5
053551.46-050205.5
053551.61-050600.5
053551.66-050810.8
053552.66-050507.0
053553.71-050235.6
053553.86-050656.1
053554.05-050416.3
053554.22-050545.0
053554.67-050630.3
053555.86-050444.3
053557.79-050504.6
053600.05-050434.0
053600.36-050552.3
053600.55-050505.3
053604.60-050402.3
053454.47-050414.3
053457.03-050252.0
053508.43-051226.3
053513.13-050753.3
053518.87-050953.2
053521.33-051248.1
053521.50-050727.8
053522.42-050242.4
053522.67-050651.3
053529.78-045955.7
053530.67-050411.7
053531.19-050917.5
053533.14-051510.5
053547.84-050823.3
053550.91-050106.7
053444.08-050600.1
053452.91-050805.0
053500.97-050541.4
053507.89-050905.3
053510.79-051045.3
053514.97-050444.9
053519.73-051145.3
053529.82-051330.9
053530.76-050337.3
053533.75-045928.7
053535.70-050909.1
053538.44-051010.5
053542.37-050252.1
053544.67-050040.4
053546.94-050700.7
174
288
45
5273
74
695
64
3604
60
67
193
88
31
288
49
88
63
63
1348
167
48
7972
4600
164
258
1398
74
168
87
169
134
68
413
250
18
6
34
4
5
15
4
4
3
27
4
8
268
28
91
57
18
8
3
3
3
34
49
5
15
6
28
15
31
14
16
Table 1 -- Continued
IDb
Namec
X-rayd Hardnesse
Counts
Ratio
J -- H f
(mag)
H -- K f
(mag)
I385
S1
S2
S3
S4
S5
S6
S7
S8
S9
S10
S11
S12
S13
053554.84-050527.6
053554.73-045808.9
053556.14-045657.7
053556.28-045505.9
053558.84-045537.0
053559.31-045846.1
053602.99-045736.2
053604.24-045744.3
053605.75-045110.3
053611.84-050032.2
053612.81-045515.5
053628.11-045709.3
053614.94-045707.7
053620.26-045424.0
23
2987
1681
510
799
743
304
285
832
567
462
185
72
164
-0.30
-0.87
-0.75
0.13
0.03
-0.64
-0.30
-0.25
-0.39
-0.47
-0.29
0.43
0.47
0.34
· · ·
0.483
0.616
· · ·
· · ·
0.695
· · ·
· · ·
0.168
0.189
· · ·
0.808
0.162
· · ·
· · ·
· · ·
· · ·
0.690
· · ·
· · ·
· · ·
· · ·
· · ·
0.168
· · ·
· · ·
· · ·
· · ·
aResults of X-ray imaging analysis and 2MASS correlation are listed
for all the detected X-ray sources.
bI1 -- I354 and S1 -- S11 are detected in the total band image (0.5 -- 8.0 keV)
image of the ACIS-I and the ACIS-S2. I355 -- I369 and S12 -- S13 are de-
tected only in the hard band (2.0 -- 8.0 keV) image of the ACIS-I and the
ACIS-S2.
I370 -- I385 are detected only in the soft band (0.5 -- 2.0 keV)
image of the ACIS-I. No new source was detected in the soft band image
of the ACIS-S2.
cSources are named after their position in the equinox J2000
dX-ray photon counts in 0.5 -- 8.0 keV band.
eDefined as H − S/H + S where H and S are photon counts in the
hard band (2.0 -- 8.0 keV) and in the soft band (0.5 -- 2.0 keV) respectively.
f J -- H and H -- K in CIT color system derived from 2MASS J, H, and
K s-band photometry and Carpenter (2001).
17
Results of Timing and Spectroscopic Analysisa
Table 2
IDb
Upperc
d
NH
Probability
(1022 cm−2)
kB T de
(keV)
Metallicity deg
(solar)
df
LX
(1029 erg s−1)
I4†
I6†
I7
I10
I12
I13†
I14
I17
I20†
I21
I25†
I27
I28
I31
I33
I36
I42†
I47†
I51
I54
I56
I65†
I67†
I71
I74
I77
I82
I85†
I87
I89
I92
I93†
I96
I110†
I113
I114
I121
I124†
I125
I131†
I138†
I139
I140
I141
I143
I144
I147†
I148†
I150†
I151
I156
I158
I161
I162
I164†
I165†
I166
I167
I168
I174†
I177†
I179
I187
I189†
0.005
0.000
0.014
0.008
0.196
0.000
0.079
0.245
0.000
0.226
0.013
0.266
0.015
0.110
0.000
0.124
0.006
0.001
0.000
0.603
0.058
0.665
0.813
0.001
0.691
0.240
0.177
0.026
0.788
0.000
0.193
0.856
0.262
0.849
0.512
0.124
0.981
0.930
0.430
0.461
0.441
0.000
0.400
0.041
0.352
0.363
0.051
0.083
0.000
0.000
0.005
0.256
0.195
0.031
0.248
0.062
0.624
0.295
0.803
0.009
0.035
0.210
0.874
0.002
0.14 (0.12 -- 0.16)
0.07 (0.05 -- 0.08)
0.14 (0.13 -- 0.16)
0.07 (0.05 -- 0.08)
0.14 (0.12 -- 0.16)
0.07 (0.06 -- 0.07)
0.02 (0.01 -- 0.04)
0.05 (0.04 -- 0.07)
0.12 (0.10 -- 0.14)
0.00 (0.00 -- 0.01)
0.05 (0.05 -- 0.06)
0.17 (0.16 -- 0.19)
0.00 (0.00 -- 0.02)
0.32 (0.30 -- 0.34)
0.14 (0.11 -- 0.18)
0.03 (0.02 -- 0.05)
0.34 (0.33 -- 0.36)
0.29 (0.28 -- 0.31)
0.19 (0.18 -- 0.19)
0.10 (0.08 -- 0.12)
0.06 (0.05 -- 0.08)
0.53 (0.48 -- 0.59)
0.88 (0.85 -- 0.91)
0.00 (0.00 -- 0.03)
2.16 (2.07 -- 2.25)
3.31 (3.21 -- 3.41)
3.92 (3.70 -- 4.16)
0.01 (0.00 -- 0.03)
2.57 (2.44 -- 2.72)
1.01 (0.89 -- 1.15)
0.07 (0.05 -- 0.09)
0.07 (0.05 -- 0.08)
1.90 (1.81 -- 2.00)
1.21 (1.18 -- 1.24)
0.00 (0.00 -- 0.03)
0.08 (0.06 -- 0.10)
2.47 (2.36 -- 2.60)
0.74 (0.71 -- 0.79)
0.54 (0.51 -- 0.58)
0.23 (0.21 -- 0.25)
3.60 (3.41 -- 3.82)
0.00 (0.00 -- 0.00)
5.88 (5.56 -- 6.23)
0.33 (0.26 -- 0.42)
1.68 (1.63 -- 1.73)
1.51 (1.46 -- 1.56)
1.07 (1.05 -- 1.10)
0.59 (0.56 -- 0.62)
0.21 (0.20 -- 0.22)
0.32 (0.32 -- 0.32)
0.08 (0.05 -- 0.11)
0.27 (0.25 -- 0.29)
0.12 (0.10 -- 0.15)
0.30 (0.26 -- 0.33)
2.01 (1.94 -- 2.08)
1.25 (1.20 -- 1.31)
0.02 (0.00 -- 0.04)
0.69 (0.66 -- 0.72)
0.26 (0.23 -- 0.29)
0.00 (0.00 -- 0.02)
3.26 (3.09 -- 3.46)
0.14 (0.12 -- 0.16)
0.04 (0.03 -- 0.06)
0.00 (0.00 -- 0.01)
0.10 (0.07 -- 0.12)
0.10 (0.09 -- 0.12)
0.07 (0.06 -- 0.08)
0.13 (0.12 -- 0.14)
0.13 (0.11 -- 0.15)
0.03 (0.02 -- 0.03)
0.08 (0.07 -- 0.10)
0.07 (0.06 -- 0.08)
0.12 (0.10 -- 0.13)
0.14 (0.12 -- 0.16)
0.14 (0.13 -- 0.15)
0.26 (0.24 -- 0.29)
0.16 (0.13 -- 0.19)
0.12 (0.11 -- 0.14)
0.07 (0.05 -- 0.09)
0.36 (0.34 -- 0.39)
0.16 (0.14 -- 0.18)
0.19 (0.17 -- 0.21)
0.38 (0.37 -- 0.39)
0.23 (0.21 -- 0.26)
0.11 (0.10 -- 0.12)
0.00 (-0.16 -- 0.14)
1.60 (1.50 -- 1.71)
0.20 (0.19 -- 0.21)
0.28 (0.18 -- 0.37)
0.74 (0.65 -- 0.85)
4.03 (3.60 -- 4.46)
0.35 (0.32 -- 0.38)
2.44 (2.17 -- 2.71)
0.26 (-0.20 -- 0.72)
0.15 (0.13 -- 0.17)
0.07 (0.07 -- 0.08)
0.01 (-0.09 -- 0.11)
0.11 (0.03 -- 0.19)
0.12 (0.11 -- 0.14)
0.23 (0.21 -- 0.25)
1.06 (0.92 -- 1.21)
0.00 (-0.16 -- 0.04)
0.62 (0.53 -- 0.72)
0.25 (0.17 -- 0.33)
0.00 (-0.25 -- 0.24)
0.32 (-0.40 -- 0.44)
1.20 (0.95 -- 1.45)
0.55 (0.07 -- 1.03)
0.44 (0.37 -- 0.51)
0.00 (-0.06 -- 0.07)
0.26 (0.22 -- 0.30)
0.09 (0.04 -- 0.15)
0.23 (0.22 -- 0.24)
0.33 (0.32 -- 0.34)
1.63 (1.39 -- 1.88)
0.22 (0.20 -- 0.24)
0.14 (0.13 -- 0.16)
0.26 (0.19 -- 0.33)
1.77 (1.64 -- 1.91)
0.36 (0.18 -- 0.53)
0.43 (0.40 -- 0.46)
0.65 (0.58 -- 0.72)
0.17 (0.14 -- 0.20)
0.22 (0.21 -- 0.24)
4.71 (4.22 -- 5.20)
0.10 (0.09 -- 0.11)
0.12 (0.11 -- 0.14)
0.16 (0.14 -- 0.17)
1.40 (1.29 -- 1.60)
0.99 (0.94 -- 1.04)
0.85 (0.81 -- 0.88)
0.98 (0.94 -- 1.02)
1.25 (1.16 -- 1.34)
1.09 (1.06 -- 1.13)
1.11 (1.06 -- 1.34)
1.02 (0.96 -- 1.07)
0.80 (0.76 -- 0.84)
1.06 (1.00 -- 1.12)
1.05 (1.02 -- 1.07)
1.31 (1.25 -- 1.37)
1.26 (1.14 -- 1.38)
1.03 (1.00 -- 1.07)
0.85 (0.76 -- 0.96)
1.21 (1.16 -- 1.26)
1.63 (1.55 -- 1.71)
1.65 (1.59 -- 1.71)
2.56 (2.52 -- 2.59)
1.22 (1.14 -- 1.29)
0.74 (0.69 -- 0.79)
3.35 (2.91 -- 3.83)
2.95 (2.83 -- 3.14)
1.04 (1.02 -- 1.07)
2.31 (2.19 -- 2.46)
2.91 (2.79 -- 3.07)
2.24 (2.01 -- 2.56)
1.63 (1.55 -- 1.70)
1.68 (1.53 -- 1.83)
7.05 (4.97 -- 12.32)
0.98 (0.92 -- 1.05)
0.84 (0.81 -- 0.88)
2.46 (2.31 -- 2.61)
7.61 (7.13 -- 8.37)
0.82 (0.77 -- 0.87)
0.82 (0.78 -- 0.87)
1.93 (1.77 -- 2.10)
3.52 (3.23 -- 3.70)
2.56 (2.33 -- 2.79)
2.83 (2.56 -- 3.18)
5.23 (4.65 -- 5.84)
3.56 (3.45 -- 3.56)
3.22 (2.92 -- 3.52)
5.55 (3.71 -- 9.64)
2.62 (2.50 -- 2.74)
2.73 (2.62 -- 2.84)
2.66 (2.57 -- 2.75)
2.39 (2.23 -- 2.56)
0.95 (0.93 -- 0.97)
1.67 (1.65 -- 1.69)
3.17 (2.74 -- 3.67)
1.33 (1.28 -- 1.39)
0.94 (0.88 -- 1.00)
2.01 (1.73 -- 2.26)
2.00 (1.87 -- 2.12)
5.76 (5.18 -- 6.59)
0.94 (0.90 -- 0.98)
2.48 (2.32 -- 2.63)
1.00 (0.93 -- 1.07)
1.22 (1.19 -- 1.26)
3.49 (3.01 -- 4.11)
0.75 (0.71 -- 0.79)
1.03 (0.99 -- 1.08)
1.21 (1.15 -- 1.27)
18
6.90 (6.50 -- 7.31)
7.64 (7.28 -- 8.01)
9.95 (9.50 -- 10.41)
11.51 (11.07 -- 11.96)
6.62 (6.24 -- 6.99)
27.49 (26.78 -- 28.19)
8.14 (7.78 -- 8.50)
8.28 (7.91 -- 8.66)
5.37 (5.04 -- 5.69)
3.56 (3.34 -- 3.78)
35.16 (34.41 -- 35.92)
10.69 (10.22 -- 11.17)
3.10 (2.88 -- 3.32)
18.72 (18.01 -- 19.43)
1.90 (1.70 -- 2.10)
9.15 (8.79 -- 9.51)
28.50 (27.57 -- 29.43)
43.22 (42.16 -- 44.28)
612.99 (609.09 -- 616.87)
7.16 (6.81 -- 7.51)
4.97 (4.68 -- 5.26)
10.13 (9.44 -- 10.81)
64.39 (62.53 -- 66.26)
14.26 (13.84 -- 14.69)
45.76 (43.47 -- 48.02)
116.03 (111.84 -- 120.22)
49.90 (46.16 -- 53.67)
33.26 (32.23 -- 34.31)
32.89 (30.30 -- 35.48)
11.65 (10.59 -- 12.71)
4.48 (4.22 -- 4.75)
8.54 (8.17 -- 8.90)
28.04 (26.46 -- 29.62)
209.63 (205.71 -- 213.57)
3.28 (3.06 -- 3.50)
4.09 (3.84 -- 4.33)
34.47 (32.28 -- 36.66)
41.27 (39.16 -- 42.37)
20.92 (19.99 -- 21.85)
17.72 (16.98 -- 18.48)
48.11 (45.34 -- 50.89)
2.06 (1.82 -- 2.30)
60.02 (56.14 -- 63.90)
4.78 (4.27 -- 5.29)
89.11 (86.28 -- 91.92)
70.64 (68.28 -- 73.04)
138.00 (134.98 -- 141.05)
26.85 (25.76 -- 27.94)
42.64 (41.71 -- 43.58)
Class h
III+MS
III+MS
III+MS
III+MS
III+MS
other
III+MS
III+MS
III+MS
III+MS
other
III+MS
III+MS
III+MS
III+MS
III+MS
II
III+MS
III+MS
III+MS
III+MS
II
II
III+MS
I
I
· · ·
III+MS
I
I
· · ·
III+MS
I
· · ·
III+MS
III+MS
I
· · ·
· · ·
· · ·
I
I
I
· · ·
· · ·
I
II
II
other
395.08 (391.75 -- 398.37)
III+MS
5.55 (5.18 -- 5.91)
18.63 (17.97 -- 19.28)
5.54 (5.20 -- 5.87)
9.63 (9.03 -- 10.22)
48.96 (46.73 -- 51.20)
51.96 (50.10 -- 53.84)
5.36 (5.09 -- 5.63)
37.13 (35.89 -- 38.38)
4.47 (4.12 -- 4.82)
25.16 (24.56 -- 25.75)
40.79 (38.19 -- 43.41)
7.46 (7.05 -- 7.87)
8.21 (7.85 -- 8.57)
10.02 (9.64 -- 10.40)
other
II
III+MS
· · ·
I
II
III+MS
II
II
III+MS
· · ·
other
III+MS
III+MS
IDb
Upperc
d
NH
Probability
(1022 cm−2)
kB T de
(keV)
Metallicity deg
(solar)
df
LX
(1029 erg s−1)
Table 2 -- Continued
I192
I194†
I197†
I200†
I204
I208
I211
I212†
I213†
I218†
I223†
I225†
I226
I232
I234
I238
I240
I242
I248†
I256
I258
I259†
I261†
I263
I264†
I269
I272†
I274
I277
I278
I280†
I282
I283†
I287
I289†
I292
I293
I295†
I297†
I298
I308
I313
I314
I317
I319†
I322
I324†
I326†
I328
I334
I339†
I342
I343†
I345
I346
I368
S1†
S2
S5
0.038
0.016
0.532
0.000
0.371
0.688
0.166
0.812
0.243
0.138
0.007
0.002
0.198
0.532
0.353
0.267
0.005
0.004
0.000
0.545
0.984
0.147
0.318
0.761
0.160
0.892
0.041
0.671
0.040
0.002
0.918
0.843
0.000
0.160
0.410
0.997
0.894
0.179
0.066
0.121
0.081
0.239
0.061
0.504
0.001
0.566
0.000
0.003
0.000
0.028
0.010
0.000
0.000
0.103
0.000
0.313
0.360
0.005
0.391
0.16 (0.13 -- 0.18)
0.22 (0.21 -- 0.24)
0.07 (0.06 -- 0.08)
0.08 (0.08 -- 0.09)
1.83 (1.74 -- 1.93)
1.74 (1.64 -- 1.86)
0.27 (0.25 -- 0.30)
3.92 (3.77 -- 4.11)
3.04 (2.92 -- 3.19)
3.80 (3.71 -- 3.90)
0.16 (0.15 -- 0.17)
0.19 (0.17 -- 0.21)
0.24 (0.22 -- 0.27)
1.61 (1.51 -- 1.71)
0.93 (0.87 -- 1.01)
0.04 (0.02 -- 0.06)
0.46 (0.45 -- 0.47)
0.23 (0.22 -- 0.25)
0.22 (0.21 -- 0.23)
0.37 (0.35 -- 0.39)
0.14 (0.11 -- 0.16)
0.29 (0.27 -- 0.32)
0.59 (0.55 -- 0.64)
0.96 (0.90 -- 1.03)
5.74 (5.42 -- 6.11)
0.15 (0.12 -- 0.17)
0.08 (0.06 -- 0.10)
0.06 (0.04 -- 0.08)
0.19 (0.15 -- 0.24)
0.12 (0.10 -- 0.15)
0.52 (0.49 -- 0.54)
0.07 (0.06 -- 0.08)
0.10 (0.09 -- 0.11)
0.00 (0.00 -- 0.04)
0.08 (0.06 -- 0.11)
0.02 (0.00 -- 0.05)
0.54 (0.51 -- 0.58)
0.18 (0.16 -- 0.21)
0.59 (0.56 -- 0.62)
0.09 (0.07 -- 0.11)
0.38 (0.36 -- 0.40)
0.61 (0.58 -- 0.64)
0.17 (0.15 -- 0.19)
0.10 (0.09 -- 0.12)
0.06 (0.06 -- 0.07)
0.04 (0.02 -- 0.06)
0.12 (0.11 -- 0.12)
0.10 (0.09 -- 0.11)
0.08 (0.07 -- 0.09)
0.04 (0.02 -- 0.07)
0.08 (0.07 -- 0.09)
0.13 (0.13 -- 0.14)
0.16 (0.15 -- 0.16)
1.47 (1.29 -- 1.68)
0.10 (0.09 -- 0.11)
1.36 (1.24 -- 1.52)
0.08 (0.08 -- 0.09)
0.03 (0.02 -- 0.04)
0.32 (0.30 -- 0.34)
2.02 (1.66 -- 2.26)
1.59 (1.53 -- 1.65)
1.22 (1.16 -- 1.28)
0.96 (0.94 -- 0.97)
2.81 (2.58 -- 3.12)
3.38 (3.02 -- 3.74)
1.99 (1.83 -- 2.15)
2.07 (1.98 -- 2.15)
2.04 (1.93 -- 2.13)
2.71 (2.61 -- 2.81)
1.03 (1.01 -- 1.06)
2.55 (2.37 -- 2.74)
1.30 (1.23 -- 1.36)
1.79 (1.67 -- 1.96)
1.47 (1.36 -- 1.57)
0.79 (0.75 -- 0.84)
1.27 (1.25 -- 1.30)
0.69 (0.67 -- 0.71)
1.60 (1.54 -- 1.66)
2.53 (2.38 -- 2.67)
0.94 (0.88 -- 1.00)
11.50 (9.17 -- 14.33)
5.64 (4.87 -- 6.63)
1.50 (1.40 -- 1.61)
7.54 (6.11 -- 10.43)
1.17 (1.06 -- 1.26)
1.33 (1.26 -- 1.41)
0.95 (0.87 -- 1.04)
3.04 (2.63 -- 4.01)
0.79 (0.74 -- 0.84)
2.50 (2.34 -- 2.66)
0.93 (0.90 -- 0.97)
1.97 (1.88 -- 2.06)
0.82 (0.77 -- 0.86)
2.84 (2.44 -- 3.42)
1.25 (1.14 -- 1.37)
2.84 (2.60 -- 3.14)
0.82 (0.78 -- 0.87)
3.20 (2.98 -- 3.42)
0.77 (0.72 -- 0.82)
1.56 (1.48 -- 1.64)
2.30 (2.18 -- 2.45)
1.00 (0.96 -- 1.04)
1.03 (0.99 -- 1.06)
1.23 (1.19 -- 1.26)
0.92 (0.84 -- 0.98)
1.61 (1.57 -- 1.64)
1.00 (0.96 -- 1.05)
1.17 (1.14 -- 1.20)
1.57 (1.28 -- 1.77)
1.23 (1.18 -- 1.28)
1.33 (1.31 -- 1.35)
1.27 (1.25 -- 1.30)
10.10 (7.56 -- 38.89)
0.98 (0.95 -- 1.01)
9.85 (7.29 -- 23.02)
0.91 (0.89 -- 0.93)
0.95 (0.92 -- 0.98)
0.75 (0.72 -- 0.79)
0.41 (0.33 -- 0.48)
0.33 (0.31 -- 0.36)
0.16 (0.15 -- 0.18)
0.14 (0.14 -- 0.15)
4.67 (4.25 -- 5.08)
0.00 (-0.17 -- 0.16)
0.15 (0.11 -- 0.20)
0.00 (-0.10 -- 0.07)
0.00 (-0.09 -- 0.07)
0.84 (0.76 -- 0.92)
0.18 (0.17 -- 0.20)
0.23 (0.18 -- 0.27)
0.19 (0.16 -- 0.21)
5.00 (4.51 -- 5.50)
0.00 (-0.05 -- 0.03)
0.10 (0.09 -- 0.11)
0.16 (0.15 -- 0.17)
0.18 (0.17 -- 0.19)
0.15 (0.13 -- 0.16)
0.02 (-0.02 -- 0.06)
0.12 (0.10 -- 0.13)
0.28 (0.04 -- 0.52)
0.00 (-0.23 -- 0.12)
0.00 (-0.04 -- 0.04)
1.64 (1.27 -- 2.00)
0.14 (0.12 -- 0.17)
0.13 (0.11 -- 0.15)
0.10 (0.08 -- 0.12)
1.57 (1.30 -- 1.92)
0.04 (0.03 -- 0.05)
1.22 (1.13 -- 1.31)
0.18 (0.17 -- 0.20)
0.46 (0.43 -- 0.49)
0.10 (0.09 -- 0.11)
0.41 (0.28 -- 0.53)
0.24 (0.19 -- 0.28)
0.12 (0.03 -- 0.21)
0.08 (0.07 -- 0.10)
0.45 (0.37 -- 0.53)
0.07 (0.06 -- 0.07)
0.18 (0.15 -- 0.20)
0.04 (-0.01 -- 0.09)
0.14 (0.13 -- 0.15)
0.13 (0.12 -- 0.14)
0.11 (0.10 -- 0.12)
0.17 (0.15 -- 0.19)
0.15 (0.14 -- 0.16)
0.09 (0.08 -- 0.10)
0.12 (0.11 -- 0.13)
0.26 (0.20 -- 0.31)
0.10 (0.09 -- 0.11)
0.09 (0.08 -- 0.09)
0.12 (0.12 -- 0.13)
1.25 (0.74 -- 2.07)
0.13 (0.12 -- 0.13)
0.00 (-0.54 -- 0.32)
0.34 (0.34 -- 0.35)
0.14 (0.13 -- 0.15)
0.19 (0.18 -- 0.20)
6.54 (6.16 -- 6.91)
30.58 (29.74 -- 31.42)
13.41 (12.95 -- 13.87)
58.62 (57.65 -- 59.59)
25.09 (23.53 -- 26.63)
22.07 (20.66 -- 23.49)
25.29 (24.22 -- 26.37)
58.90 (55.25 -- 61.84)
46.47 (43.63 -- 49.06)
212.30 (206.01 -- 218.57)
24.54 (23.87 -- 25.20)
34.11 (33.03 -- 35.19)
10.64 (10.10 -- 11.18)
16.06 (14.74 -- 17.38)
11.87 (10.93 -- 12.66)
4.33 (4.06 -- 4.59)
94.24 (92.37 -- 96.09)
15.22 (14.57 -- 15.87)
44.82 (43.78 -- 45.85)
36.72 (35.52 -- 37.91)
6.46 (6.05 -- 6.86)
37.15 (35.88 -- 38.41)
27.28 (26.06 -- 28.51)
12.87 (11.97 -- 13.76)
62.79 (59.27 -- 66.33)
7.02 (6.58 -- 7.45)
8.43 (8.04 -- 8.83)
3.13 (2.90 -- 3.36)
5.05 (4.61 -- 5.49)
4.30 (4.00 -- 4.60)
36.72 (35.49 -- 37.93)
13.49 (13.04 -- 13.95)
41.61 (40.70 -- 42.53)
4.00 (3.77 -- 4.24)
6.37 (5.99 -- 6.75)
2.55 (2.32 -- 2.77)
20.92 (19.95 -- 21.90)
6.35 (5.97 -- 6.72)
43.61 (42.19 -- 45.01)
4.35 (4.08 -- 4.63)
33.49 (32.28 -- 34.71)
26.76 (25.67 -- 27.85)
11.75 (11.24 -- 12.27)
12.34 (11.83 -- 12.85)
36.51 (35.71 -- 37.30)
4.01 (3.76 -- 4.25)
105.68 (104.17 -- 107.17)
11.73 (11.25 -- 12.20)
63.15 (62.08 -- 64.24)
3.82 (3.54 -- 4.09)
23.32 (22.66 -- 23.98)
170.53 (168.58 -- 172.51)
94.26 (92.82 -- 95.68)
14.26 (13.01 -- 15.51)
23.55 (22.89 -- 24.22)
16.62 (15.30 -- 17.76)
59.24 (58.12 -- 60.38)
21.79 (21.20 -- 22.39)
17.52 (16.75 -- 18.29)
Class h
other
III+MS
other
III+MS
II
I
· · ·
I
I
I
III+MS
II
III+MS
I
I
III+MS
III+MS
III+MS
III+MS
other
III+MS
II
· · ·
II
I
other
III+MS
III+MS
III+MS
I
II
III+MS
III+MS
III+MS
III+MS
III+MS
· · ·
III+MS
II
III+MS
II
II
III+MS
III+MS
III+MS
III+MS
III+MS
III+MS
III+MS
III+MS
III+MS
III+MS
other
· · ·
II
· · ·
III+MS
III+MS
III+MS
aResults of timing and spectroscopic analysis are listed for all the "bright2" sources. Best-fit parameters of thin-thermal
plasma model fitting are tabulated regardless whether the fit is accepted or rejected.
bSources with time-variability are shown with †.
cDefined as R ∞
α χ2
ν (x)dx where the χ2-value and the degree of freedom of the fit is α and ν. Fits with the upper
probability larger than 0.01 are "accepted" while those less than or equal to 0.01 are "rejected". Even the "best-fit"
parameters of rejected fittings represent values of some physical meaning or approximated values. We thus use these values
in the same way with the accepted ones.
dThe lower and upper boundary for 1 σ confidence level are given in parentheses.
eThe best fit value of temperature and metallicity are searched for in the range of less than 20 keV and 5 solar respectively.
19
f Luminosity in 0.5 -- 8.0 keV band (corrected for absorption).
gAbundance relative to solar photospheric (Morrison & McCammon 1983).
hClassification based on the J, H, and K s-band magnitude of 2MASS counterpart. "I" for class I, "II" for class II, and
"III+MS" for class III+MS. "others" are for sources which are out of these groups. " · · · " sources can not be classified
Best-fit Values of Two-Temperature Thin-Thermal Plasma Modela
Table 3
ID
I10
I42
I47
I71
I150
I156
I174
I189
I200
I223
I225
I240
I242
I248
I278
I283
I319
I324
I326
I328
I343
I346
S2
b
NH
(1022 cm−2)
0.42(0.41 -- 0.44)
0.36(0.34 -- 0.38)
0.36(0.35 -- 0.38)
0.00(0.00 -- 0.02)
0.24(0.23 -- 0.25)
0.22(0.19 -- 0.25)
0.00(0.00 -- 0.01)
0.00(0.00 -- 0.01)
0.04(0.03 -- 0.04)
0.13(0.12 -- 0.14)
0.28(0.27 -- 0.30)
0.51(0.50 -- 0.52)
0.15(0.13 -- 0.16)
0.27(0.26 -- 0.28)
0.01(0.00 -- 0.04)
0.11(0.10 -- 0.12)
0.04(0.03 -- 0.05)
0.12(0.12 -- 0.13)
0.07(0.06 -- 0.09)
0.11(0.11 -- 0.12)
0.19(0.18 -- 0.19)
0.06(0.05 -- 0.07)
0.00(0.00 -- 0.01)
bce
kB T 1
(keV)
bde
kB T 2
(keV)
Metallicitybeg
(solar)
bcf
LX 1
(1029 erg s−1)
bdf
LX 2
(1029 erg s−1)
0.22(0.21 -- 0.22)
1.14(1.11 -- 1.23)
0.79(0.75 -- 0.83)
0.80(0.77 -- 0.83)
0.83(0.81 -- 0.85)
0.84(0.72 -- 0.98)
1.06(1.04 -- 1.09)
0.59(0.52 -- 0.65)
0.83(0.81 -- 0.84)
0.79(0.75 -- 0.82)
1.00(0.95 -- 1.06)
0.86(0.83 -- 0.89)
0.70(0.67 -- 0.73)
0.84(0.81 -- 0.87)
0.71(0.65 -- 0.79)
0.89(0.85 -- 0.94)
0.84(0.81 -- 0.86)
1.01(0.99 -- 1.04)
0.81(0.77 -- 0.85)
0.65(0.62 -- 0.68)
0.68(0.66 -- 0.70)
0.78(0.75 -- 0.81)
0.64(0.60 -- 0.68)
1.07(1.02 -- 1.12)
2.41(2.20 -- 2.65)
2.46(2.31 -- 2.62)
2.08(1.92 -- 2.25)
10.51(6.34 -- 26.52)
2.53(2.26 -- 9.32)
3.71(3.01 -- 4.69)
1.70(1.61 -- 1.79)
2.27(2.14 -- 2.43)
1.75(1.67 -- 1.82)
10.09(7.13 -- 16.98)
2.02(1.92 -- 2.11)
2.44(1.95 -- 3.28)
2.93(2.75 -- 3.16)
3.17(2.67 -- 12.51)
2.89(2.75 -- 3.11)
2.12(2.00 -- 2.24)
3.28(3.11 -- 3.45)
2.75(2.27 -- 3.40)
1.36(1.33 -- 1.40)
1.77(1.73 -- 1.81)
1.97(1.81 -- 2.15)
1.64(1.57 -- 1.72)
0.43(0.41 -- 0.45)
0.18(0.16 -- 0.21)
0.47(0.45 -- 0.50)
0.78(0.75 -- 0.82)
0.33(0.32 -- 0.34)
0.61(0.52 -- 0.70)
0.29(0.27 -- 0.30)
0.55(0.51 -- 0.59)
0.51(0.49 -- 0.52)
0.64(0.61 -- 0.67)
0.11(0.10 -- 0.12)
0.32(0.30 -- 0.33)
0.69(0.65 -- 0.72)
0.28(0.27 -- 0.30)
0.40(0.34 -- 0.47)
0.96(0.92 -- 1.00)
0.33(0.31 -- 0.35)
0.20(0.19 -- 0.21)
0.29(0.26 -- 0.31)
0.19(0.18 -- 0.20)
0.31(0.30 -- 0.32)
0.49(0.47 -- 0.52)
0.73(0.70 -- 0.76)
33.08(30.39 -- 35.88)
12.43(11.55 -- 13.29)
15.30(14.28 -- 16.32)
6.64(6.28 -- 7.00)
12.31(11.62 -- 13.00)
17.62(16.61 -- 18.64)
36.59(35.36 -- 37.82)
9.18(8.68 -- 9.69)
39.42(38.43 -- 40.41)
12.26(11.11 -- 13.43)
1.39(1.11 -- 1.67)
17.93(17.37 -- 18.48)
2.51(2.23 -- 2.79)
28.99(28.22 -- 29.75)
9.72(9.17 -- 10.28)
20.03(19.02 -- 21.05)
47.06(45.07 -- 49.08)
8.76(8.28 -- 9.25)
4.62(4.21 -- 5.02)
9.80(8.99 -- 10.62)
8.04(7.64 -- 8.44)
28.45(27.43 -- 29.47)
14.57(13.90 -- 15.23)
22.15(20.68 -- 23.64)
61.20(59.25 -- 63.18)
3.99(3.46 -- 4.52)
20.34(19.38 -- 21.30)
33.04(31.81 -- 34.28)
1.15(0.97 -- 1.32)
7.99(7.39 -- 8.59)
11.27(10.68 -- 11.87)
46.24(44.97 -- 47.49)
5.53(5.14 -- 5.91)
18.43(17.43 -- 19.45)
29.90(28.63 -- 31.18)
10.17(9.65 -- 10.69)
7.42(6.97 -- 7.86)
3.03(2.71 -- 3.36)
37.90(36.84 -- 38.97)
26.54(25.68 -- 27.41)
69.48(67.68 -- 71.25)
7.12(6.55 -- 7.68)
50.93(49.76 -- 52.11)
75.64(74.08 -- 77.22)
13.13(12.45 -- 13.81)
14.14(13.52 -- 14.76)
Class g
III+MS
II
III+MS
III+MS
other
other
III+MS
III+MS
III+MS
III+MS
II
III+MS
III+MS
III+MS
I
III+MS
III+MS
III+MS
III+MS
III+MS
other
II
III+MS
aAccepted fits are listed.
bThe lower and upper boundary for 1 σ confidence level are given in parentheses.
cBest-fit value of the lower temperature component.
dBest-fit value of the higher temperature component.
eThe absorption and the metallicity of the higher temperature component is fixed to be the same with those of the lower temperature
component. The best fit value of temperature and metallicity are searched for in the range of less than 20 keV and 5 solar respectively.
f Luminosity of each component in 0.5 -- 8.0 keV band (corrected for absorption)
gAbundance relative to solar photospheric (Morrison & McCammon 1983).
hClassification based on the J, H, and K s-band magnitude of 2MASS counterpart. "I" for class I, "II" for class II, and "III+MS" for
class III+MS. "others" are for sources which are out of these groups. " · · · " sources can not be classified due to their lack of a NIR
counterpart.
Table 4
X-ray Properties of Each Class
Frequency . . . . . . . . . . . . . . . . . .
Sources with time-variation .
Source with two temperature
Absorption(1022cm−2)a. . . . . . .
Metallicity (solar)ab. . . . . . . . . . .
Temperature (keV)a. . . . . . . . . .
Luminosity (1029erg s−1)a. . . .
Class I
19
6 (32%)
1 ( 5%)
Class II
Class III+MS
18
10 (56%)
3 (17%)
61
21 (34%)
15 (25%)
2.49(2.37 -- 2.62)
0.82(0.61 -- 0.99)
2.99(2.66 -- 3.57)
49.1(46.7 -- 51.5)
0.63(0.60 -- 0.67)
0.60(0.50 -- 0.69)
2.91(2.63 -- 3.26)
36.3(35.1 -- 37.5)
0.11(0.10 -- 0.13)
0.20(0.18 -- 0.22)
1.20(1.13 -- 1.28)
36.4(35.7 -- 37.0)
aThe lower and upper boundary for 1 σ confidence level are given in parentheses.
bAbundance relative to solar photospheric (Morrison & McCammon 1983).
20
|
astro-ph/0511721 | 1 | 0511 | 2005-11-25T14:27:22 | Completing HI observations of galaxies II. The Coma Supercluster | [
"astro-ph"
] | High sensitivity 21-cm HI line observations, with an rms noise level of \sim 0.5 mJy, were made of 35 spiral galaxies in the Coma Supercluster, using the refurbished Arecibo telescope, which resulted in the detection of 25 objects. These data, combined with the measurements available from the literature, provide the set of HI data for 94% of all late-type galaxies in the Coma Supercluster with an apparent photographic magnitude m_p <15.7 mag. We confirm that the typical scale of HI deficiency around the Coma cluster is 2 Mpc, i.e. one virial radius. Comparing the HI mass function (HIMF) of cluster with non-cluster members of the Coma Supercluster we detect a shortage of high HI mass galaxies among cluster members that can be ascribed to the pattern of HI deficiency found in rich clusters. | astro-ph | astro-ph | Astronomy & Astrophysics manuscript no. gavazzi05b
(DOI: will be inserted by hand later)
September 21, 2018
5
0
0
2
v
o
N
5
2
1
v
1
2
7
1
1
5
0
/
h
p
-
o
r
t
s
a
:
v
i
X
r
a
Completing HI observations of galaxies II. The Coma
Supercluster
G. Gavazzi1, K. O'Neil2, A. Boselli3, and W. van Driel4
1 Universit´a degli Studi di Milano-Bicocca, Piazza delle scienze 3, 20126 Milano, Italy
e-mail: [email protected]
2 NRAO, P.O. Box 2, Green Bank, WV 24944, U.S.A.
e-mail: [email protected]
3 Laboratoire d'Astrophysique de Marseille, BP8, Traverse du Siphon, F-13376 Marseille, France
e-mail: [email protected]
4 Observatoire de Paris, Section de Meudon, GEPI, CNRS UMR 8111 and Universit´e Paris 7, 5 place Jules
Janssen, F-92195 Meudon Cedex, France
e-mail: [email protected]
the date of receipt and acceptance should be inserted later
Abstract. High sensitivity 21-cm H i line observations, with an rms noise level of ∼ 0.5 mJy, were made of 35 spiral
galaxies in the Coma Supercluster, using the refurbished Arecibo telescope, which resulted in the detection of 25
objects. These data, combined with the measurements available from the literature, provide the set of H i data
for 94 % of all late-type galaxies in the Coma Supercluster with an apparent photographic magnitude mp ≤ 15.7
mag. We confirm that the typical scale of H i deficiency around the Coma cluster is 2 Mpc, i.e. one virial radius.
Comparing the H i mass function (HIMF) of cluster with non-cluster members of the Coma Supercluster we detect
a shortage of high H i mass galaxies among cluster members that can be ascribed to the pattern of H i deficiency
found in rich clusters.
Key words. Galaxies: distances and redshifts -- Galaxies: general -- Galaxies: ISM -- Galaxies: clusters -- individual
Virgo -- Radio lines: galaxies
1. Introduction
H i line observations of galaxies have provided us with
some of the most powerful diagnostics on the role of the
environment in regulating the evolution of late-type (spi-
ral) galaxies in the local Universe. Spiral galaxies in rich
X-ray luminous clusters display a significant lack of H i
gas with respect to their "undisturbed" counterparts in
the field (Haynes & Giovanelli 1984, Giovanelli & Haynes
1985). This pattern of H i deficiency can be attributed
to various interaction mechanisms: ram-pressure (Gunn
& Gott 1972), viscous stripping (Nulsen 1982), thermal
evaporation (Cowie & Songaila 1977) or tidal interaction
with the cluster potential well (Byrd & Valtonen 1990;
Moore et al. 1996)). Since these mechanisms have a higher
efficiency in or near rich cluster cores, the H i deficiency
parameter (Haynes & Giovanelli 1984 - see Sect. 5.1) is an
environmental indicator that provides a clear signature of
a galaxy's membership of a rich cluster.
The Coma Supercluster, due to its proximity to us (∼
100 Mpc), has received considerable attention in H i stud-
Send offprint requests to: G. Gavazzi
ies. Since the pioneering study by Sullivan et al. (1981),
various works (e.g. Chincarini et al. 1983a; Gavazzi 1987;
Gavazzi 1989: Scodeggio & Gavazzi 1993; Haynes et al.
1997) have provided measurements of the H i content for
most late-type galaxies in the Coma Supercluster. In addi-
tion to these single-dish studies of their global H i proper-
ties, the detailed mapping of galaxies in the Coma and
A1367 clusters with radio synthesis telescopes was ob-
tained by Bravo-Alfaro et al., (2000, 2001) and Dickey
& Gavazzi, (1991).
A high sensitivity, blind H i survey of 7000 square deg. of
sky is planned for 2005-2006 with the ALFA multibeam
system at Arecibo (Giovanelli et al. 2005), and even more
sensitive (1 mJy rms) surveys of parts of these clusters
will be obtained with the ALFA system. They will in-
clude the Coma Supercluster and the Virgo cluster. In
preparation for these surveys, before the installation of
the ALFA system, we used the single-beam Arecibo sys-
tem for the continuation of the pointed observation survey
of late-type galaxies in the Virgo cluster (see Gavazzi et al.
2005a, Paper I) and in the Coma Supercluster area. Here
we report on the results of the Coma Supercluster obser-
2
Gavazzi et al.: H i observations in the Coma Supercluster
Fig. 1. The wedge diagram of the Coma Supercluster galaxies in the declination interval 18◦ < δ < 32◦. Members are
divided among late-type (large circles, filled if observed in HI) and early-type (E-S0a: empty squares). Empty triangles
mark foreground and background galaxies.
vations which, in conjunction with the previously available
H i data-set, enable us to review the properties of galaxies
in this Supercluster, as obtained from optically selected
H i observations.
The selection of the cluster targets for H i observations
is described in Section 2, the observations and the data
reduction are presented in Section 3 and the results are
given in Section 4 and discussed in Section 5. A Hubble
constant of 75 km s−1 Mpc−1 is assumed throughout this
paper.
2. Sample selection
Galaxies in the present study are selected from the CGCG
Catalogue (Zwicky et al. 1961-68) in the region 11h30m <
α < 13h30m; 18◦ < δ < 32◦. There are 1127 CGCG galax-
ies listed in this region with an apparent photographic
magnitude mp ≤15.7. Their wedge-diagram is given in Fig.
1 where the structure of the Coma -- A1367 Supercluster
stands out clearly as the pronounced density enhance-
ment near 7000 km s−1, as part of one of the largest
known coherent structures in the local universe, named
the "Great Wall" (Ramella et al. 1992). Other conspicu-
ous features are: the "Fingers of God" of Coma and A1367,
Gavazzi et al.: H i observations in the Coma Supercluster
3
Fig. 2. The sky projection of Coma Supercluster members (including the "homunculus legs" HL). Spirals are repre-
sented by circles (filled if observed in H i) and E-S0a objects by empty squares.
spanning the interval 4000 < V < 10000 km s−1, mostly
traced by early-type objects, and the large (∼ 7500 Mpc3)
"void" in front of the Supercluster, with a density 150
times lower than the mean galaxy density in the uni-
verse. Other remarkable features, also known as the "legs
of the homunculus" (de Lapparent et al. 1986) are two fil-
aments pointing toward the Coma cluster in the interval
4500 < V < 7000 km s−1. A third "homunculus leg" sur-
rounds the void on the western side, projected near A1367.
Objects in the interval 6000 < V < 8000 km s−1 form a
bridge between the two clusters with a narrow velocity dis-
tribution. Above V > 8000 km s−1 the Coma Supercluster
fades into the background, and setting a boundary be-
tween its members and the projected background objects
is rather arbitrary. In assigning the membership of indi-
vidual galaxies to the various sub-structures we follow the
criteria of Gavazzi et al. (1999).
Out of the 1127 galaxies in Fig. 1, 654 are considered
proper Supercluster members according to these criteria,
and 76 additional galaxies belonging to the "homuncu-
lus legs" (HL) are considered separately (see Table 1).
Their sky projection is given in Fig. 2, revealing a sub-
stantial morphology segregation between late-type (spiral)
and early-type (E-S0a) objects.
The Coma Supercluster has been observed in H i with
remarkable completeness: 65 % of the published data are
found in 6 publications: Chincarini et al. (1983a); Gavazzi
(1987, 1989), Scodeggio & Gavazzi (1993); Haynes et al.
(1997).
After the present high sensitivity (rms ∼ 0.5 mJy) obser-
vations of 33 additional CGCG galaxies (and of two fainter
objects: FOCA 610, 636) 295/315 (94 %) late-type mem-
bers (including HL) are observed (see Table 1), of which
259 were detected and 36 have upper limits. Of the re-
maining 20 unobserved late-type targets 13 are members
of double or multiple systems that could not be resolved
by the Arecibo beam (97-111S, 97-129E, 127-051N, 128-
029E, 127-025N, 159-049S, 128-031N, 128-002, 97-036, 98-
072, 98-073, 98-081, 98-087). The Coma cluster (160-243)
and 6 galaxies (127-121, 129-016, 157-077, 158-046, 160-
180, 161-029) belonging to the HL were not observed due
to scheduling constraints. To fill in a scheduling hole, 13
galaxies in the Virgo cluster were also observed. All but
one were detected. The results of these observations are
given in Table 3 and the HI profiles of the detected galaxies
are shown in Figure 8. These objects will not be consid-
ered further in this paper.
All data on the Coma Supercluster and Virgo galax-
ies are collected and made available worldwide via
4
Gavazzi et al.: H i observations in the Coma Supercluster
Table 1 Sample completeness.
4. Results
Type<Sa Type≥Sa
All H i All H i
Coma members
HL
Background
Foreground
All
391
24
131
46
592
58
10
5
16
89
263
52
118
102
535
249
46
55
92
442
the "Goldmine" website (http://Goldmine.mib.infn.it; see
Gavazzi et al. 2003).
3. Observations
Using the refurbished 305-m Arecibo Gregorian radio tele-
scope we observed 35 galaxies in the Coma Supercluster
(plus 13 in the Virgo cluster) (see Section 2) in February
2004 and January-March 2005. Data were taken with the
L-Band Wide receiver, using nine-level sampling with two
of the 2048 lag subcorrelators set to each polarization
channel. All observations were taken using the position-
switching technique, with each blank sky (or OFF) posi-
tion observed for the same duration, and over the same
portion of the telescope dish as the on-source (ON) ob-
servation. Each 5min+5min ON+OFF pair was followed
by a 10s ON+OFF observation of a well-calibrated noise
diode. The overlaps between both sub-correlators with
the same polarization allowed a contiguous radial velocity
search range of sufficient width from -1000 to 8500 km s−1.
The velocity resolution was 2.6 km s−1, the instrument's
HPBW at 21 cm is 3′.5×3′.1 and the pointing accuracy
is about 15′′. The pointing positions used are the opti-
cal center positions of the target galaxies listed in Table
1. Flux density calibration corrections are good to within
10% (and often much better), see the discussion of the er-
rors involved in O'Neil (2004).
Using standard IDL data reduction software available at
Arecibo, corrections were applied for the variations in the
gain and system temperature with zenith angle and az-
imuth. A baseline of order one to three was fitted to the
data, excluding those velocity ranges with H i line emis-
sion or radio frequency interference (RFI). The velocities
were corrected to the heliocentric system, using the opti-
cal convention, and the polarizations were averaged. All
data were boxcar smoothed to a velocity resolution of 12.9
km s−1 for further analysis. For all spectra the rms noise
level was determined and for the detected objects the cen-
tral line velocity, the line widths at, respectively, the 50%
and 20% level of the peak, and the integrated line flux were
determined. No flux correction depending on the source
size was applied because the optical extent of all detected
targets does not significantly exceed the Arecibo HPBW.
the
rms
In order to identify sources whose H i detections could
have been confused by nearby galaxies, we queried the
NED, HyperLeda and Goldmine databases and inspected
DSS images over a region of 10′ radius surrounding the
central position of each source, given the telescope's
sidelobe pattern. Quoted values are weighted averages
from the HyperLeda database, unless otherwise indicated.
The H i spectra of both the clearly and the marginally
detected galaxies are shown in Figures 7,8 and the
global H i line parameters are listed in Table 3. These
are directly measured values; no corrections have been
applied to them for, e.g., instrumental resolution. Table
3 is organized as follows:
Column 1: Obj. is the galaxy designation;
Column 2-3: (J2000) celestial coordinates;
Column 4: the heliocentric optical recessional velocity (in
km s−1);
Column 5:
(mJy/beam);
Column 6: Sp is the peak flux density of the detected line
(mJy/beam);
Column 7: VHI is the heliocentric central radial velocity
of a line profile (in km s−1), in the optical convention,
with its estimated uncertainty (see below);
Columns 8-9: W50 and W20 are the line widths at 50%
and 20% of peak maximum, respectively, (km s−1);
Column 10: IHI is the integrated line flux (Jy km s−1),
with its estimated uncertainty (see below).
Column 11: A quality flag to the spectra is given, where
Q=1 stands for high signal-to-noise, double horned
profiles, Q=2 for high signal-to-noise, single horned
profiles, and Q=3,4 for low signal-to-noise profiles whose
measured line parameters are not reliable. Q=5 is given
to unpublished profiles.
We estimated the uncertainties σVHI (km s−1) in VHI and
σIHI (Jy km s−1) in IHI following Schneider et al. (1986,
1990), as:
dispersion in the baseline
σVHI = 1.5(W20 − W50)X −1
and
σIHI = 2(1.2W20/R)0.5Rσ = 7.9(W20)0.5σ
(1)
(2)
where IHI is the integrated line flux (Jy km s−1), R is
the instrumental resolution (12.9 km s−1), and X is the
signal-to-noise ratio of a spectrum, i.e. the ratio of the
peak flux density Sp and σ, the rms dispersion in the
baseline (Jy).
The uncertainties in the W20 and W50 line widths
are expected to be 2 and 3 times σVHI , respectively.
5. Discussion
The newly obtained H i data were combined with those
available from the literature for the mp ≤ 15.7 late-type
galaxies in the Coma Supercluster, as listed in Table 4.
Gavazzi et al.: H i observations in the Coma Supercluster
5
Fig. 3. The late-type Coma Supercluster members (including HL) coded according to their H i deficiency parameter:
DefHI ≤ 0.5 (empty circles); DefHI > 0.5 (filled circles).
The sample comprises 315 galaxies, of which 295 were
observed, 259 were detected and 36 remain undetected.
Table 4 is organized as follows:
Column 1: Galaxy designation in the CGCG Catalog;
Column 2: Morphological type;
Column 3: Apparent photographic magnitude from the
CGCG;
Column 4: membership as defined in Gavazzi et al.
(1999). Distances D of 96.0 and 91.3 Mpc are assumed
for Coma and A1367 respectively. Distances of individual
groups and substructures in the Great Wall (including
HL) are taken from Gavazzi et al. (1999) (rescaled to
Ho = 75 km s−1 Mpc−1). Distances from individual
redshifts are assumed for supercluster isolated galaxies
and members of multiplets.
Column 5: recessional velocity, in km s−1;
in solar units:
Column 6: H i mass or mass limit
MHI = 2.36 105D2 IHI . For undetected galaxies we set
IHI =1.5 × rmsHI × W<20−50>, where rmsHI is the rms
of the spectra in mJy and the W<20−50> profile width is
based on the following average line widths of the detected
objects per Hubble type bin: 300 km s−1 for Sa-Sbc, 190
km s−1 for Sc-Scd;
Column 7: Coded reference to the H i measurement (see
the notes to the Table);
Column 8: H i deficiency parameter as defined in Haynes
& Giovanelli (1984) (see Sect. 5.1);
Column 9: Quality flag (see last Column of Table 3).
5.1. The pattern of H i deficiency
For the late-type galaxies in the present study we estimate
the H i deficiency parameter following Haynes & Giovanelli
(1984) as the logarithmic difference between MHI of a
reference sample of isolated galaxies and MHI actually
observed in individual objects: DefHI = LogMHI ref. −
LogMHI obs.. MHI ref is computed from the galaxies opti-
cal linear diameter d as: LogMHI ref = a+bLog(d), where
d is estimated consistently with Haynes & Giovanelli
(1984), a and b are weak functions of the Hubble type, as
listed in Table 3 of Paper I (notice that b ∼ 2 across the
Hubble sequence, i.e. MHI ref increases approximately as
the galaxy linear diameter squared).
Fig.3 shows that H i deficient galaxies segregate around
the center of the Coma cluster, and to a much lesser degree
around A1367. One important issue that can be addressed
with the present data-set, given its high completeness, is
on what scale the phenomenon of H i ablation holds. The
H i deficiency parameter of individual galaxies is given in
6
Gavazzi et al.: H i observations in the Coma Supercluster
Fig. 4. The projected distribution about the X-ray cen-
ter of the Coma cluster of the H i deficiency of Late-
type Supercluster detected (dots) and undetectd (trian-
gles) members (excluding HL). Large circles represent av-
erages in bins of 0.5◦ (1◦) One σ errorbars are given.
Fig. 5. The correlation between log MHI and Bo
T for Sa-
Sb (filled circles), Sbc-Sc (empty circles), Scd-Im-BCD
(filled squares). The direct fit (solid line), the inverse fit
(dashed) (obtained considering all Hubble types) and the
relation used in Paper I (dotted) are given.
Fig.4 as a function of the projected linear separation from
the X-ray center of the Coma cluster, out to 15 Mpc, along
with average values taken in bins of 0.5◦(from 0◦to 2◦) and
in bins of 1◦ further out (see also Table 2). It is apparent
that significant H i deficiency occurs out to approximately
3 Mpc radius. At one virial radius (i.e. at 2.2 Mpc, Girardi
et al. 1998; or 2.9 Mpc Lokas & Mamon 2003; Neumann
et al. 2001), the average H i content of the supercluster
galaxies becames indistinguishable from that of the field,
in agreement with Solanes et al. (2001)1.
The star formation rate, as derived from Hα observations,
show a significant cut-off near 1-2 projected virial radii
from the Coma cluster (Gavazzi et al. 2005b) and from
over-density peaks in the SDSS (Nichol 2004; Miller 2004).
What the H i observations show is the driver of the SFR
decline, i.e. the lack of gas to sustain it.
magnitude range we have included the Virgo galaxies
(Paper I). In doing so we must obviously exclude galaxies
with perturbed H i contents. Conservatively we exclude
galaxies with DefHI ≥ 0.2 (we recall that the DefHI pa-
rameter is determined from diameters (see Section 5.1), in-
dependently of the B luminosity, as in Haynes & Giovanelli
1984). Fig. 5 shows the relation for 465 late-type galax-
ies which were selected accordingly, plotted in three bins
of Hubble type to stress the consistency among them.
The three given linear regressions, obtained combining all
Hubble types, are the direct one (MHI = 3.680 − 0.299 ×
Mp), the inverse one (Mp = 7.090 − 2.795 × MHI) and the
one adopted in Paper I (MHI = 2.9−0.34×Mp). The resid-
ual of the (direct) correlation is σ(logMHI )=0.26, i.e., the
H i mass of disk galaxies can be predicted within a factor
of 1.8 uncertainty from their B luminosity.
5.2. The H i mass -- B o
T relation
5.3. The H i mass function
Taking advantage from the large sample of optically se-
lected galaxies with H i measurements and optical (B-
band) photometry (accurate to within 0.1 mag) we study
the H i mass vs. Bo
T relation for unperturbed galaxies.
However since the Coma Supercluster sample (mp ≤ 15.7)
contains only giant galaxies brighter than Mp = −19.1, for
the purpose of extending this correlation over a broader
1 Galaxies in the HL, (being identified as such outward of
2◦ from Coma and A1367), on average do not show significant
deficiency (they are not included in Fig.4)
Fig. 6 shows the frequency distribution of MHI for the
cluster (Coma + A1367) late-type galaxies (solid his-
togram) and for the isolated galaxies (including HL) (the
latter normalized to the former by the ratio of the num-
ber of galaxies). The H i Mass Function (HIMF) so ob-
tained cannot be meaningfully compared with the one of
Virgo, obtained in Paper I, nor with the one of isolated
galaxies by Zwaan et al. (2003) because of the shallow-
ness of the optical and the 21cm observations available
for the Coma Supercluster. Given the relation MHI =
Table 2 The radial distribution of the mean binned H i deficiency of Late-type Supercluster members (excluding HL) about the Coma cluster.
Gavazzi et al.: H i observations in the Coma Supercluster
7
Θ(Deg.) R (Mpc)
0.0-0.5
0.5-1.0
1.0-1.5
1.5-2.0
2.0-3.0
3.0-4.0
4.0-5.0
5.0-6.0
6.0-7.0
7.0-8.0
0.42
1.25
2.09
2.93
4.19
5.87
7.55
9.24
10.9
12.6
< DefHI >
0.83± 0.19
0.43± 0.43
0.44± 0.45
0.16± 0.43
0.00± 0.25
0.00± 0.19
0.00± 0.32
-0.10± 0.21
-0.01± 0.39
-0.11± 0.33
3.680 − 0.299 × Mp, discussed in the previous section,
the limiting magnitude mp ≤ 15.7 (Mp ≥ −19.1) of
the CGCG implies that on average only galaxies with
logMHI > 9.3M⊙ are targeted. This imposes a com-
pleteness cut-off that is even shallower than the limit-
ing detectable H i mass of logMHI >9.0 M⊙ that de-
rives from the typical noise figure of the 21cm observations
(< σ >∼ 1 mJy). In Paper I we showed that significant
differences between the HIMF of the Virgo cluster and of
the field occur for logMHI <9.0 M⊙, i.e. below both the
present cut-off lines. For 9.5 < logMHI < 10 M⊙ there
is however in Fig. 6 a barely significant excess in the fre-
quency of non-cluster members of the Coma Supercluster
with respect to cluster members. This can be understood
as a signature of the deficiency pattern of cluster galaxies,
as found in Paper I for Virgo.
6. Summary and conclusions
We have observed in the 21-cm H i line, with the re-
furbished Arecibo telescope, 35 galaxies in the Coma
Supercluster and 13 in the Virgo cluster. The high sensi-
tivity of our observations (rms noise ∼ 0.5 mJy) resulted
in the detection of 37 objects and significant upper limits
were obtained for the remaining ones.
Including the present observations the H i survey of the
Coma Supercluster has reached virtually the completion
(94 % among the late-type members).
Combining all data, we determine with high significance
that the typical scale of H i deficiency around the Coma
cluster is 2-3 Mpc, i.e. one virial radius.
With the present data a meaningful determination of the
HIMF can be obtained only for logMHI > 9.0M⊙, insuf-
ficient to compare with the deeper HIMF of Virgo (Paper
I) and of isolated galaxies (Zwaan et al. 2003). Comparing
cluster with non-cluster Supercluster members we detect
however a shortage of high H i mass galaxies among clus-
ter members that can be ascribed to the pattern of H i
deficiency found in rich clusters.
Acknowledgements. The Arecibo Observatory is part of the
National Astronomy and Ionosphere Center, which is operated
Fig. 6. The MHI
function for the members of the
Coma+A1367 clusters (solid) and for the non-cluster
Supercluster members (dotted). The optical (logMHI =9.3
M⊙) and radio (logMHI =9.0 M⊙) completeness limits are
drawn.
by Cornell University under a cooperative agreement with the
National Science Foundation. This research also has made use
of the Lyon-Meudon Extragalactic Database (LEDA), recently
incorporated in HyperLeda and the NASA/IPAC Extragalactic
Database (NED) which is operated by the Jet Propulsion
Laboratory, California Institute of Technology, under contract
with the National Aeronautics and Space Administration, and
the Goldmine database.
References
Beijersbergen, M., 2003, PhD thesis, Groningen University
Binggeli, B., Sandage, A., & Tammann, G. A., 1985, AJ, 90,
1681 (VCC)
8
Gavazzi et al.: H i observations in the Coma Supercluster
Bothun, G., Aaronson, M., Schommer, R., Mould, J., Huchra,
J., & Sullivan, W., 1985, ApJS, 57, 423
Bravo-Alfaro, H., Cayatte, V., van Gorkom, J. H., & Balkowski,
Solanes, J. et al., 2001, ApJ, 548, 97.
Sulentic, J., & Arp H., 1982, AJ, 88, 489
Sullivan III, W.T., Bothun, G.D., Bates, B., & Schommer,
C., 2000, AJ, 119, 580
R.A., 1981, AJ, 86, 919
Bravo-Alfaro, H., Cayatte, V., van Gorkom, J. H., & Balkowski,
van Driel, W., Ragaigne, D., Boselli, A., Donas, J., & Gavazzi,
C., 2001, A&A, 379, 347
G., 2000, A&AS, 144, 463
Vogt, N. P., Haynes, M. P., Herter, T., & Giovanelli, R., 2004,
AJ, 127, 3273
Williams, B., & Kerr, F., 1981, AJ, 86, 953
Zwaan, M. A., et al. 2003, AJ, 125, 2842
Byrd, G., & Valtonen, M., 1990, ApJ, 350, 89
Chincarini, G., Giovanelli, R., & Haynes, M. P., 1983a, ApJ,
269, 13
Chincarini, G., Giovanelli, R., Haynes, M., & Fontanelli, P.,
1983b, ApJ, 267, 511
Cowie, L.L., & Songaila, A., 1977, Nat., 266, 501
de Lapparent, V., Geller, M.J., & Huchra, J.P., 1986, ApJ, 302,
L1
Dell'Antonio, I., Bothun, G., & Geller, M., 1996, AJ, 112, 1759
Dickey, J. M., & Gavazzi, G. 1991, ApJ, 373, 347
Eder, J, Giovanelli, R., & Haynes, M., 1991, AJ 102, 572
Fontanelli, P., 1984 A&A, 138, 85
Gavazzi, G., 1987, ApJ, 320, 96
Gavazzi, G., 1989, ApJ, 346, 59
Gavazzi, G., Carrasco, L., & Galli, R., 1999, A&AS, 136, 227
Gavazzi, G., Boselli, A., Donati, A., Franzetti, P., & Scodeggio,
M., 2003, A&A, 400, 451
Gavazzi, G., Boselli, A., van Driel, W., & O'Neil, K., 2005a,
A&A, 429, 439 (Paper I)
Gavazzi, G., Boselli, Cortese, L., Arosio, I., Gallazzi, A.,
Pedotti, P., & Carrasco, L., 2005b (submitted to A&A)
Giovanelli, R., & Haynes, M., 1985, ApJ, 292, 404
Giovanelli, R., (& 23 co-authors), 2005, (AJ, in press)
Girardi, M., Giuricin, G., Mardirossian, F., Mezzetti, & M.,
Boschin, W., 1998, ApJ, 505, 74
Gunn, J.E., & Gott, J. R.III, 1972, ApJ, 176, 1
Haynes, M. & Giovanelli, R., 1984, AJ, 89, 758
Haynes, M., Giovanelli, R., & Chincarini, G., 1984, ARA&A,
22, 445
Haynes, M., Giovanelli, R., Herter, T. et al., 1997, AJ, 113,
1197
Lewis, B.. Helou, G., & Salpeter, E., 1985, ApJS, 59, 161
Lokas, E.L. & Mamon, G.A., 2003, MNRAS, 343, 401
Lu, N., et al., 1993, ApJS, 88, 383
Magri, C., 1994, AJ, 108, 896
Miller, C. J. 2004 in "Clusters of Galaxies: Probes of
from
Cosmological Structure and Galaxy Evolution,
Symposia.
the Carnegie Observatories Centennial
Published by Cambridge University Press, as part of
the Carnegie Observatories Astrophysics Series. Edited by
J.S. Mulchaey, A. Dressler, and A. Oemler.
Moore, B., Katz, N., Lake, G., Dressler, A., & Oemler, A., Jr.
1996, Nat, 379, 613
Mould, J., et al., 1995, ApJS, 96, 1
Neumann, D. M. et al. 2001, A&A 365, L74
Nichol R.C., 2004 in "Clusters of Galaxies: Probes of
Cosmological Structure and Galaxy Evolution,
from
the Carnegie Observatories Centennial
Symposia.
Published by Cambridge University Press, as part of
the Carnegie Observatories Astrophysics Series. Edited by
J.S. Mulchaey, A. Dressler, and A. Oemler, 2004, p. 24.
Nulsen, P.E.J., 1982, MNRAS, 198, 1007
O'Neil, K., 2004, AJ, 128, 2080
Ramella, M., Geller, M. J., & Huchra, J. P., 1992, ApJ, 384,
396
Salzer, J., Hanson M., & Gavazzi G., 1990, ApJ. 353, 39
Scodeggio, M., & Gavazzi, G., 1993, ApJ, 409, 110
Table 3 Parameters of the newly observed galaxies.
Gavazzi et al.: H i observations in the Coma Supercluster
9
Obj.
RA
Dec
J2000.0
Vopt
km/s
σ
Sp
mJy/beam mJy/beam
VHI
km/s
W50 W20
km/s
km/s
IHI
Qual.
Jy km/s
Coma Supercluster
114316.24
CGCG 97-078
113944.62
CGCG 127-018
114330.88
CGCG 127-039
114456.97
CGCG 97-124
114646.66
CGCG 127-055
115039.40
CGCG 127-067
115123.19
CGCG 157-044
CGCG 97-172
115214.30
CGCG 127-137W 120141.90
120144.40
CGCG 98-023
CGCG 127-136
120147.40
120155.47
CGCG 127-138
120456.20
CGCG 128-015
121755.86
CGCG 99-002
CGCG 128-072
121808.27
121910.00
CGCG 99-013
122052.50
CGCG 128-081
122949.40
CGCG 129-004
124011.35
CGCG 159-048
CGCG 159-071
124543.41
125206.77
CGCG 159-097
125432.95
CGCG 160-009
125756.70
FOCA 636
FOCA 610
125757.70
125947.24
CGCG 130-003
130059.10
CGCG 160-261
130422.57
CGCG 160-128
130635.52
CGCG 160-138
CGCG 160-146
130814.00
131440.60
CGCG 160-169
131730.80
CGCG 130-029
131947.60
CGCG 160-195
CGCG 161-048
132557.14
132643.29
CGCG 161-051
CGCG 161-054
132703.07
Virgo Cluster
VCC 30
VCC 85
VCC 113
VCC 137
VCC 429
VCC 578
VCC 1574
VCC 1623
VCC 1821
VCC 1873
VCC 1898
VCC 2071
CGCG 43066
121054.50
121336.50
121432.90
121508.60
122044.20
122243.60
123432.90
123531.90
124008.90
124118.60
124157.50
124825.40
125515.40
194455.6
224107.7
230043.3
194353.9
211616.9
205426.1
264703.6
183905.7
202417.3
175353.8
210506.7
204452.1
211427.5
182357.8
244118.5
191626.9
252547.2
222219.4
311038.1
292558.5
270134.3
282234.7
275930.0
280342.0
214845.7
275358.6
284838.5
271006.2
273055.3
295951.8
203555.1
304931.9
313703.6
303026.7
305834.3
155654.6
130201.1
120612.0
145819.5
143803.3
183252.1
151052.3
163646.9
065302.1
063127.0
034909.4
091856.8
025348.4
7560±70
-
6908±60
7771±60
6615±52
6349±39
6624±44
7650±56
6794
6905±48
6930
7192±46
6729±16
7640±59
6838±51
7297±16
7112
6847±70
7096±44
6936±44
6573±190
7079±63
4649
8299
7094±32
6896±7
8054±36
7852±60
7337±23
6960±60
6671±71
7297±15
7268±44
7150±60
7668±44
-
-
2155
-
-
-
-
-
-
-
881 ±23
-
2798±23
0.54
0.56
0.62
0.66
0.67
0.43
0.64
0.36
0.49
0.38
0.29
0.60
0.44
0.90
0.63
0.36
0.42
0.57
0.65
1.06
0.65
0.70
0.40
0.38
0.55
0.37
0.74
0.37
0.25
0.69
0.36
0.34
0.66
1.52
0.62
0.92
0.32
0.33
0.56
0.39
0.39
0.17
0.30
0.31
0.26
0.50
0.32
0.47
-
17.0
22.4
-
9.3
3.8
5.1
1.6
3.4
5.8
17.5
-
19.0
2.7
8.3
7.2
5.0
5.5
7.0
17.1
3.2
-
1.8
1.7
4.4
-
25.7
-
-
6.2
4.6
5.0
-
-
7.0
9.5
2.4
17.5
-
4.0
4.0
2.3
1.9
2.5
17.0
-
2.7
21.0
-
6922±2
6911±1
-
6626±3
6400±16
6607±4
7826±16
6871±16
6905±16
6675±16
-
6741±16
(7605)
6795±6
7336±16
7204±16
6736±7
7064±8
6971±1
6424±30
-
4605±30
8125±8
7140±4
-
7920±1
-
-
6850±4
6560±16
7247±16
-
-
-
147
37
-
215
283
240
214
357
262
173
-
102
-
119
235
290
275
309
189
260
-
130
211
335
-
115
-
-
298
317
158
-
-
-
168
58
-
242
304
260
227
398
282
228
-
124
-
174
251
312
317
363
202
357
-
219
236
357
-
137
-
-
318
325
223
-
-
-
2.09±0.06
0.82±0.04
-
1.86±0.08
0.73±0.10
0.84±0.08
0.21±0.10
1.07±0.10
1.01±0.10
2.36±0.10
-
1.64±0.10
-
0.82±0.07
1.41±0.10
1.00±0.10
1.09±0.08
1.60±0.10
2.60±0.10
0.60±0.10
-
0.18±0.05
0.19±0.05
0.51±0.08
-
2.46±0.07
-
-
1.26±0.10
1.30±0.10
0.63±0.10
-
-
6756±3
284
303
1.34±0.09
2084 ±16
1932 ±16
2091 ±16
(4000)
600 ±16
6490 ±16
639 ±16
2108 ±16
1007 ±16
1692 ±16
-
6484 ±16
2802 ±16
116
74
148
-
87
161
127
78
75
112
-
185
382
148
84
171
-
109
178
161
91
87
134
-
196
403
0.83±0.10
0.14±0.10
2.12±0.10
-
0.27±0.10
0.48±0.10
0.22±0.10
0.13±0.10
0.16±0.10
1.37±0.10
-
0.40±0.10
6.09±0.10
-
1
2
-
1
1
1
3
1
1
2
-
2
4
1
1
1
1
1
1
3
-
3
3
3
-
1
-
-
1
1
2
-
-
1
2
2
2
4
2
1
2
2
2
2
-
2
1
10
Gavazzi et al.: H i observations in the Coma Supercluster
Table 4: Basic H i properties of late-type members to the Coma Supercluster (in-
cluding HL).
CGCG
Type mp
mag
097-005
097-026
097-027
097-036
097-062
097-063
097-064
097-068
097-072
097-073
097-076
097-078
097-079
097-082
097-087
097-091
097-092
097-093
097-102N
097-111S
097-114
097-119
097-120
097-121
097-122
097-129E
097-129W
097-130
097-138
097-149
097-151
097-152
097-168
097-169
097-172
098-002
098-007
098-013
098-016
098-017
098-023
098-034
098-041
098-046
098-050
098-051
098-058
098-067
098-071
098-072
098-073
098-074
098-081
098-085
098-087
098-116
099-002
099-013
100-005
100-012
101-033
101-043
101-049
101-054
127-005
127-018
127-025N
127-025S
127-026
127-033
127-035
127-037
127-038
127-039
127-049
127-050
127-051N
127-052
127-053
127-054
Sc
Pec
Sc
S..
Pec
Pec
S..
Sbc
Sa
Pec
Sb
Sa
Pec
Sa
Pec
Sa
Sbc
Pec
Sa
Pec
Pec
Sa
Sa
Sab
Pec
Sbc
Sb
Sa
Pec
S..
Sab
Sa
S..
Sc
S..
Sb
Sbc
Sc
Sc
Sbc
Sb
S..
Sc
Sa
Sc
Sa
Sbc
S..
Pec
S..
Sb
Sa
Sa
Sc
S..
Sc
S..
Sc
Pec
Pec
Sc
Sa
Sbc
Sab
Sbc
Sb
Sc
Sbc
Sbc
Sc
Sa
Pec
Sc
Sbc
Pec
Sbc
Pec
Sa
Sbc
Sb
15.5
13.9
14.6
15.7
15.5
15.7
15.6
14.7
15.0
15.6
15.5
15.2
15.7
15.0
14.3
14.7
15.5
15.5
15.1
16.5
15.4
15.7
14.5
14.6
14.9
15.7
14.0
15.5
15.5
15.6
15.6
15.5
15.7
15.7
15.7
15.6
15.5
15.1
15.3
15.7
15.1
14.8
15.7
14.3
14.1
14.6
14.7
15.7
15.5
15.7
15.7
15.6
15.2
14.7
15.3
14.9
15.5
15.7
14.4
15.3
15.7
15.0
14.9
13.8
15.4
15.0
15.3
14.5
14.8
15.2
15.4
15.4
14.0
15.3
15.5
14.8
15.8
14.0
15.4
14.2
Cloud
Isol.
Pair
Pair
Pair
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
A1367
Isol.
A1367
Isol.
Pair
Pair
Isol.
Isol.
Isol.
Isol.
Isol.
Tripl.
N4065 G
N4065 G
N4065 G
IC202 HL
IC202 HL
Isol.
Isol.
Pair
Pair
Tripl.
Isol.
Pair
Tripl.
Pair
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Pair
Pair
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
A1367
N3937 G
A1367
A1367
Isol.
N3937 G
V
km s−1
log MHI
M ⊙
Ref.
DefHI
Qual
6129
6202
6630
6595
7809
6102
5976
5974
6332
7290
7060
7526
6996
6100
6735
7368
6373
4857
6361
7239
8257
5256
5595
6571
5468
6009
5082
6697
5317
6060
5854
6166
5996
5975
7826
6206
6350
6949
6449
7015
6905
6432
7551
6220
4372
4283
7207
7610
6881
7946
6440
7471
7177
7042
7540
6229
7605
7736
6611
6481
6729
6677
7148
6606
6864
6922
7142
7076
6871
6300
6817
6186
6913
6911
7061
6752
7288
6946
6409
7026
9.72
9.88
9.27
-
9.33
9.09
9.17
9.99
9.14
9.31
< 8.39
< 8.68
9.21
< 8.68
9.83
9.77
9.18
9.03
9.21
-
9.44
8.92
8.80
9.37
9.35
-
10.09
< 8.83
9.69
< 9.25
9.05
9.40
< 8.91
9.45
8.71
9.41
9.86
9.65
9.72
9.66
9.30
9.07
9.28
9.58
9.44
9.6
9.95
< 9.64
9.62
-
-
< 9.10
-
9.69
-
9.91
9.02
9.50
9.37
9.32
9.68
< 8.74
10.06
9.92
9.64
9.63
-
9.87
9.95
9.77
9.49
9.54
10.36
9.21
9.34
9.95
-
9.73
1 -
10.38
126
39
24
-
40
24
126
126
126
40
40
193
40
2
40
126
24
4
40
-
193
4
4
126
126
-
24
1
40
105
88
126
88
6
193
88
88
40
15
40
193
40
24
24
24
88
24
184
184
-
-
88
-
24
-
24
193
193
24
24
43
88
7
24
6
193
-
126
24
24
88
24
2
193
6
2
-
126
24
2
-0.29
-0.45
0.28
-
0.31
0.22
0.03
-0.14
0.50
0.16
> 1.39
> 1.28
0.25
> 1.01
0.19
-0.18
0.31
0.57
0.36
-
-0.17
0.22
0.90
0.28
0.49
-
0.18
> 0.40
-0.22
> 0.07
0.34
0.23
> 0.25
0.05
0.38
0.10
-0.30
-0.12
0.11
0.05
-0.01
0.11
0.65
0.07
-0.04
-0.35
0.00
> -0.71
-0.05
-
-
> 0.36
-
-0.13
-
-0.34
-0.08
-0.40
0.41
-0.12
0.07
> 0.96
-0.24
-0.20
0.04
-0.14
-
0.09
-0.17
0.00
0.15
-0.10
-0.20
0.25
0.32
0.00
-
0.16
-0.11
0.10
1
1
2
-
1
2
3
1
1
2
-
-
1
5
1
1
4
4
1
-
1
1
4
1
1
-
1
5
3
-
1
1
-
2
3
1
1
1
1
2
1
3
2
1
1
4
1
-
1
-
-
-
-
2
-
1
3
1
1
2
1
-
2
3
2
1
-
1
1
1
1
2
1
2
1
1
-
1
1
1
Gavazzi et al.: H i observations in the Coma Supercluster
11
Table 4: Continue
CGCG
Type mp
mag
Pec
127-055
Sb
127-056
Sc
127-061
S..
127-067
Pec
127-071
Sc
127-072
Sb
127-073
Sc
127-082
Sbc
127-083
Sa
127-085
Sbc
127-087
Sc
127-095
Sc
127-099
Sb
127-100
Sbc
127-104
Sb
127-106
Sbc
127-107
Sbc
127-109
Sbc
127-110
Sbc
127-111
Sbc
127-112
127-114E
Pec
127-114W Pec
Sc
127-118
127-120
Sb
Sb
127-121
Sc
127-123
Sb
127-127
Sc
127-133
S..
127-136
S..
127-137W
127-138
S..
Sa
127-139
S..
128-002
Pec
128-003
Sb
128-015
S..
128-016
128-021
Sbc
Sa
128-023
S..
128-029E
S..
128-029W
S..
128-031N
Sbc
128-037
128-042N
S..
S..
128-044
Sc
128-049
Sb
128-052
Sbc
128-053
S..
128-057
S..
128-058
128-059
Sb
Sa
128-063
S..
128-066
Sbc
128-069
Pec
128-072
Sb
128-073
128-075
Sc
Sc
128-079
Sb
128-080
S..
128-081W
Sb
128-082
Sc
128-087
128-089
Sa
Sc
128-090
S..
129-004
Sa
129-009
S..
129-013
S..
129-016
129-018
Sc
Sb
129-020
S..
129-021
Sab
129-022
S..
129-023
Sc
129-025
S..
130-002
130-003
Sb
Sbc
130-005
Sbc
130-006
Sc
130-008
Sbc
130-009
130-012
Sbc
15.1
15.7
15.4
15.5
15.4
14.6
15.1
14.7
15.1
15.5
15.4
14.2
14.5
14.9
15.5
14.5
15.7
15.7
14.4
15.7
14.8
15.0
15.0
15.2
14.1
15.7
14.7
14.6
15.3
15.7
16.0
15.5
15.5
15.7
14.6
15.3
15.2
15.4
14.4
16.7
16.2
16.5
14.8
16.3
15.7
15.0
15.6
15.6
15.6
15.7
15.6
15.3
15.1
15.6
15.4
14.7
15.5
15.6
15.0
16.5
15.7
15.3
14.2
15.5
15.2
15.3
15.7
15.5
13.8
14.8
15.3
14.4
15.7
13.5
15.6
15.4
15.5
15.0
14.9
15.3
15.2
Cloud
Isol.
Isol.
Isol.
N3937 G
N3937 G
N3937 G
N3937 G
N3937 G
N3937 G
N3937 G
N4005 HL
N3937 G
Pair
N3937 G
Isol.
N4005 HL
Isol.
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
N4005 HL
Pair
N4065 G
Isol.
Isol.
N4065 G
Isol.
Pair
Isol.
Isol.
N4065 G
IC762 G
IC762 G
Isol.
IC762 G
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
N4213 G
N4213 G
Pair
N4213 G
N4213 G
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Pair
Isol.
N4615 HL
N4615 HL
Isol.
Isol.
Isol.
Isol.
N4615 HL
Isol.
Isol.
Isol.
Isol.
Isol.
Pair
Pair
V
km s−1
log MHI
M ⊙
Ref.
DefHI
Qual
6626
6814
5954
6400
6388
6438
6439
6654
6743
6595
4941
6199
6458
6854
6814
5027
6355
4731
4495
4617
4828
4771
4771
4551
4470
4131
4479
4048
4667
6675
6871
7210
6713
6780
6435
6741
6619
7064
6719
6634
7158
6936
7194
7319
6922
6445
6692
7307
6998
6778
6228
6747
6526
7188
6795
6948
6682
6630
7349
7204
6910
6671
6841
6776
6736
6415
6962
4979
4716
6579
6697
6972
6746
4380
6663
7140
7058
6521
7266
6335
7131
9.54
9.67
9.85
9.18
9.36
9.89
9.01
9.47
9.39
< 8.99
9.66
10.04
9.71
9.58
9.63
9.88
9.36
9.4
9.81
9.47
9.63
9.84
9.84
9.38
9.32
-
9.84
9.34
9.12
9.68
9.38
< 8.77
9.29
-
9.69
9.51
9.30
9.83
9.92
-
9.39
-
9.53
9.78
10.25
9.21
9.48
10.06
< 9.51
9.52
9.95
9.97
< 9.11
9.16
9.21
9.96
9.93
9.71
9.38
9.33
1 -
9.55
9.23
9.87
9.33
9.27
9.53
-
10.05
9.50
9.85
9.98
9.41
10.02
9.61
9.04
9.30
9.33
9.73
9.93
9.97
193
126
24
193
24
3
1
2
6
88
126
2
126
2
6
24
126
32
24
111
32
24
24
24
24
-
24
111
32
193
193
193
24
-
2
193
24
24
2
-
40
-
2
43
7
24
22
40
184
7
7
24
88
111
193
126
126
40
24
193
111
5
40
7
193
24
184
-
24
15
24
2
184
6
184
193
127
24
2
24
24
-0.45
-0.05
-0.20
-0.31
0.13
-0.05
0.81
0.02
0.06
> 0.46
0.05
0.03
0.10
0.06
0.19
-0.32
0.02
-0.04
0.22
-0.41
0.25
-0.02
-0.13
0.13
0.22
-
-0.03
0.14
-0.08
-1.02
-0.32
> -0.09
0.34
-
-0.11
-0.11
-0.31
0.02
-0.30
-
-0.10
-
0.11
-0.30
-0.36
0.49
0.07
0.01
> -0.35
-0.25
-0.28
0.06
> 0.34
0.28
0.02
0.01
0.10
-0.25
0.05
-0.29
-0.05
0.13
0.35
-0.35
-0.08
0.11
-0.03
-
-0.23
0.14
-1.15
-0.25
-0.42
0.12
-0.64
0.68
0.21
0.04
-0.53
-0.03
0.19
1
1
1
1
2
1
4
2
3
-
1
1
1
3
1
1
3
2
1
1
1
3
3
1
3
-
1
1
1
2
1
-
3
-
2
2
1
1
1
-
3
-
2
3
1
4
2
1
-
2
1
1
-
2
1
1
3
2
3
1
1
1
3
1
1
1
1
-
1
3
2
3
4
1
1
3
1
2
2
1
1
12
Gavazzi et al.: H i observations in the Coma Supercluster
Table 4: Continue
CGCG
Type mp
mag
130-014
130-021
130-025
130-026
130-027
130-029
131-008
131-009
157-012
157-032
157-035
157-044
157-051
157-062
157-064
157-075
157-077
158-009
158-010
158-029
158-030
158-031
158-036
158-038
158-042
158-046
158-047
158-053N
158-054
158-055
158-056
158-061
158-070
158-081
158-091
158-102
158-105
158-112
159-004
159-005
159-008
159-009
159-010
159-019
159-031
159-033
159-037
159-040
159-048
159-049S
159-054
159-055
159-058
159-059
159-060
159-061
159-064
159-068
159-071
159-072N
159-072S
159-076
159-080
159-081
159-082
159-090
159-091
159-092
159-093
159-095
159-096
159-097
159-101
159-102
160-001
160-005
160-007
160-009
160-012
160-015
160-018
Sbc
Sa
Sa
Sc
Sbc
Sc
Sbc
Sc
Sbc
Sa
Sb
Pec
Sc
Pec
Sb
Sc
S..
Sb
Sbc
S..
Sab
Sb
Sb
Sab
S..
S..
Sb
Sa
Pec
Sb
Sa
Sa
Sbc
Pec
Sab
S..
Sbc
Sbc
S..
Sbc
Sb
Sab
Sb
Sbc
Sa
Sa
Sab
Sa
S..
S..
Sc
Sbc
Sa
Sab
Pec
Sbc
S..
Sa
Sc
Pec
Pec
Sbc
Sb
Sbc
Sc
Sc
S..
Sc
Sc
Sbc
Sc
Pec
Pec
Sab
Sb
Sb
S..
S..
S..
S..
S..
15.1
15.4
15.5
15.5
15.6
15.4
15.6
15.3
15.1
15.2
13.7
15.4
15.3
15.5
14.8
15.7
15.4
14.0
15.2
14.1
14.6
13.8
13.8
15.3
14.8
15.0
13.5
14.7
14.6
15.3
15.5
13.7
15.3
14.5
15.7
15.7
15.1
14.4
15.7
14.7
14.6
14.1
15.7
14.9
15.3
15.0
14.6
15.2
15.5
15.7
15.5
15.6
15.5
14.5
15.5
14.8
15.6
15.7
15.5
14.8
14.8
14.5
15.7
15.5
14.8
15.5
15.1
14.9
15.3
14.9
15.1
15.4
15.3
14.5
15.6
14.8
15.4
15.5
15.7
15.5
15.3
Cloud
Isol.
Isol.
Isol.
Quadr.
Quadr.
Pair
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
N4005 HL
Isol.
Isol.
Isol.
N4005 HL
Pair
Pair
N4169 HL
N4169 HL
N4169 HL
Isol.
Isol.
N4169 HL
N4169 HL
N4169 HL
Pair
Isol.
Isol.
Tripl.
N4169 HL
Quadr.
Isol.
IC3165 G
N4615 HL
Isol.
Pair
Pair
Pair
Isol.
N4615 HL
Isol.
N4615 HL
Pair
Isol.
Isol.
Isol.
Isol.
Pair
N4615 HL
Pair
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Pair
Pair
Isol.
Isol.
Pair
Pair
Tripl.
Isol.
N4615 HL
N4615 HL
Isol.
Isol.
Isol.
Coma
Coma
Coma
Isol.
Coma
Coma
Isol.
Coma
Coma
V
km s−1
log MHI
M ⊙
Ref.
DefHI
Qual
7096
7163
7001
6870
6834
6560
5972
7522
6814
6811
6281
6607
5151
6882
6407
6694
4100
7494
7930
3836
3970
3825
6532
6725
3868
3848
3903
6599
7685
7650
8102
3876
7634
6734
7607
4515
6824
7165
7004
6996
7393
4551
7009
4573
7511
7674
7291
7019
7064
6330
4759
7737
6797
7528
7182
6966
7265
6313
6971
6631
6590
6743
6859
8116
8078
8315
6443
4754
5446
6837
6186
6424
7745
7061
7945
6319
6462
7132
6348
7443
7092
9.57
9.33
9.68
9.85
9.94
9.39
9.76
9.64
9.73
9.16
10.18
9.19
9.65
9.68
9.84
9.54
-
9.07
9.46
8.98
9.51
10.03
9.96
9.36
8.97
-
9.79
9.79
9.76
9.53
9.53
<8.24
9.58
9.26
9.54
9.49
9.98
9.73
9.11
9.61
10.03
10.14
9.62
9.62
9.91
9.34
9.73
9.93
9.53
-
9.41
10.01
< 9.50
9.76
9.67
9.49
< 9.64
9.25
9.72
10.09
10.03
9.51
9.54
9.85
10.01
10.16
9.16
9.94
< 8.58
9.64
9.96
9.02
9.12
9.97
9.44
10.06
< 8.82
< 8.84
< 9.36
< 8.90
< 8.82
126
127
15
24
40
193
88
24
40
24
24
193
24
5
40
7
-
24
24
22
22
39
2
40
88
-
39
2
2
24
24
168
5
2
2
7
24
2
88
2
126
24
5
6
127
126
126
24
193
-
88
24
184
2
5
2
184
184
193
2
2
2
126
40
126
126
126
126
40
2
126
193
6
126
88
126
88
193
184
40
88
0.09
0.18
0.02
-0.03
-0.13
0.04
-0.04
-0.04
-0.19
0.64
-0.05
0.22
-0.04
0.01
0.02
0.06
-
0.65
0.17
0.46
-0.35
-0.13
-0.20
0.07
0.11
-
0.07
-0.02
-0.13
0.46
0.16
0.84
0.23
-0.03
-0.17
0.03
-0.16
0.41
0.15
0.27
0.06
-0.52
0.24
-0.09
-0.16
0.54
-0.28
-0.34
-0.25
-
-0.01
0.01
> -0.07
-0.31
0.04
0.28
> -0.12
0.03
-0.03
-0.03
-0.12
0.39
0.14
-0.20
-0.18
-0.56
0.00
0.02
> 0.63
-0.24
-0.01
0.21
0.07
-0.25
0.09
-0.01
> 0.74
> 0.55
> -0.16
> 0.63
> 0.63
1
1
1
1
1
1
1
1
1
4
1
1
2
2
2
2
-
4
2
2
2
1
1
1
2
-
1
1
3
4
1
-
1
2
2
1
1
4
2
2
1
1
1
1
1
1
1
1
1
-
1
1
-
2
3
3
-
4
1
4
4
1
1
1
1
1
1
1
5
1
1
3
1
3
1
1
-
-
-
5
-
Gavazzi et al.: H i observations in the Coma Supercluster
13
Table 4: Continue
CGCG
Type mp
mag
160-020
160-025
160-026
160-031
160-032
160-055
160-058
160-064
160-067
160-073
160-076
160-081
160-086
160-088
160-095
160-096N
160-098
160-102
160-106
160-108
160-114
160-121
160-127
160-128
160-137
160-138
160-139
160-141
160-146
160-148
160-151
160-152W
160-155
160-156
160-163N
160-164
160-166
160-167
160-168
160-169
160-173
160-175
160-180
160-181
160-182
160-183
160-192
160-195
160-206
160-207
160-209
160-212
160-213
160-243
160-252
160-257
160-260
160-261
161-029
161-031
161-040
161-041
161-043
161-044
161-048
161-051
161-052
161-054
161-063
161-066
161-069
161-071
161-073
Pec
Sa
Sc
S..
Sb
Sab
Sbc
Pec
Pec
Pec
Sc
Sb
Pec
Sb
Sb
Pec
Pec
Sab
Pec
Pec
S..
Sb
Sc
Pec
Sa
S..
Pec
Pec
Sa
Sa
Pec
Sb
Sb
Sa
S..
Sc
Sb
Sb
Sc
S..
Sc
S..
S..
Sc
Sab
Pec
Sb
S..
S..
Sc
Pec
Sa
Pec
S..
Pec
Sa
Sa
S..
Sb
Sbc
Sc
S..
Sa
Sc
Sa
S..
Pec
Sa
Sbc
S..
Sb
Pec
Sb
15.5
14.0
15.5
15.7
14.9
14.2
15.5
15.4
15.4
15.1
15.6
14.7
15.4
14.6
13.7
15.2
15.3
14.8
15.1
15.5
15.6
15.5
15.5
15.3
13.9
15.7
15.0
15.5
15.4
14.3
15.1
14.0
15.3
15.3
15.7
15.2
13.6
15.0
14.2
15.6
13.6
15.1
15.4
14.3
15.0
14.7
14.3
15.7
15.6
15.3
15.4
14.9
15.5
15.6
15.1
14.6
13.7
15.6
15.7
14.9
15.6
15.5
14.4
15.3
15.1
15.6
15.1
15.5
15.5
15.7
14.6
14.9
14.2
Cloud
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Coma
Pair
N5056 HL
N5056 HL
N5056 HL
N5056 HL
Pair
Isol.
N5056 HL
N5056 HL
N5056 HL
Isol.
Isol.
N5056 HL
N5056 HL
N5056 HL
N5056 HL
Isol.
N5056 HL
Isol.
Tripl.
N5056 HL
N5056 HL
Pair
Coma
Coma
Coma
Coma
Coma
Coma
Coma
N5056 HL
Tripl.
Isol.
N5056 HL
Isol.
N5056 HL
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
Isol.
N5056 L
Isol.
V
km s−1
log MHI
M ⊙
Ref.
DefHI
Qual
4968
6702
7545
6852
7747
7164
7616
7368
7655
5425
5390
5898
7499
7287
5482
6892
8762
7095
6876
8323
7454
6676
5500
7920
7050
7861
4749
7292
7385
5988
6258
5610
6366
7262
6877
6074
6408
6039
7476
6850
5592
5661
5581
5550
6994
5605
6649
7247
5053
5081
7168
7549
9386
5121
7718
5821
7985
6917
4930
7270
7260
4979
6638
4983
7280
7150
7072
6756
7300
7380
7172
4827
7320
8.93
< 8.75
9.29
< 8.84
< 8.83
9.34
9.49
< 8.32
9.34
8.57
9.64
< 8.86
8.74
9.35
9.33
8.42
9.05
9.97
9.05
< 8.29
< 8.74
9.92
9.71
9.73
9.79
< 8.56
9.96
8.95
< 8.41
9.75
8.97
9.95
9.60
9.59
< 8.92
9.69
9.80
9.75
10.36
9.41
10.12
9.25
-
9.87
9.62
9.57
10.25
9.14
8.99
9.71
9.28
8.78
< 8.17
-
9.01
8.66
9.18
< 8.56
-
9.75
9.43
8.81
9.62
9.21
< 8.82
< 9.17
9.46
9.41
9.97
10.12
9.88
9.84
10.38
24
27
40
6
2
40
126
132
181
132
132
1
126
40
126
132
1
126
132
6
88
126
126
193
40
193
126
1
193
126
24
2
126
24
24
126
2
126
2
193
126
24
-
126
5
40
126
193
40
5
24
132
1
-
40
132
126
193
-
40
40
88
24
40
193
193
88
193
88
40
24
5
24
0.27
> 0.92
0.23
> 0.62
> 0.76
0.49
0.40
> 0.93
-0.01
0.96
-0.35
> 1.29
0.76
0.42
0.96
0.95
0.41
-0.01
0.54
> 0.92
> 0.81
0.02
-0.10
-0.33
-0.05
> 0.71
-0.19
0.32
> 1.12
-0.07
0.31
0.05
0.04
0.28
> 0.28
-0.19
0.24
0.02
-0.40
0.08
-0.12
0.18
-
-0.11
0.07
-0.14
-0.09
-0.04
0.25
0.05
0.28
0.90
> 1.11
-
0.56
0.97
0.81
> 0.81
-
0.08
-0.02
0.32
0.33
0.48
> 0.72
> -0.12
-0.50
-0.07
-0.06
-0.56
-0.27
-0.26
-0.21
4
-
2
5
5
2
1
-
2
5
5
5
1
1
1
5
2
1
5
5
-
1
1
1
1
-
2
1
-
1
2
1
1
1
-
1
1
1
1
1
1
1
-
1
1
3
1
2
1
1
2
5
5
-
1
5
1
-
-
2
1
1
2
1
-
-
2
1
1
3
1
1
1
1 : Giovanelli & Haynes (1985); 2 : Chincarini et al. (1983a); 3 : Sullivan et al. (1981); 4 : Chincarini et al. (1983b); 5 : Fontanelli (1984); 6 :
Bothun et al. (1985); 7 : Williams & Kerr (1981); 15 : Haynes & Giovanelli (1984); 22 : Sulentic & Arp H (1982); 24 : Gavazzi (1987) 27 : Eder et
al. (1991); 39 : Lewis et al. (1985); 40 : Gavazzi (1989); 43 : Salzer et al. (1990); 88 : Scodeggio & Gavazzi (1993); 105: Lu et al. (1993); 111: Mould
et al. (1995); 126: Haynes et al. (1997); 127: Dell'Antonio et al. (1996); 132: Bravo Alfaro (2001); 168: Magri et al. (1994); 181: Beijsbergen (2003);
184: van Driel et al. (2000); 185: Vogt et al. (2004); 193: This work
14
Gavazzi et al.: H i observations in the Coma Supercluster
7. Appendix A: Notes on individual galaxies
CGCG 127-018: Our H i detection (VHI =6922±2 km s−1, W50=147 km s−1 and IHI =2.1±0.06 Jy km s−1) is consistent with our previous (van Driel
et al. 2000) Nan¸cay detection (VHI =6936±16 km s−1, W50=131 km s−1 and IHI =2.7±0.3 Jy km s−1) and with the Arecibo detection of Gavazzi
(1987) at VHI =6935 km s−1.
CGCG 127-039: The integrated line intensity of our H i detection (VHI =6911±1 km s−1, W50=37 km s−1 and IHI =0.82±0.04 Jy km s−1) is
3.4 times lower than that of our previous Nan¸cay detection (van Driel et al. 2000), with VHI =6919±9 km s−1, W50=40 km s−1 and IHI =2.8±0.2
Jy km s−1. The latter was made with an elongated 3′.6×21′ (α×δ)HPBW, which is considerably larger than the 3′.6 round Arecibo beam. This
difference can be due to the 13.6 B mag SBbc spiral NGC 3832, 17′.2 due south of the target galaxy, whose mean H i line parameters (VHI =6909±6
km s−1, W50=171 km s−1 and IHI =10.4 Jy km s−1) are based on 5 spectra, all obtained at Arecibo (Chincarini et al. 1983a; Giovanardi & Salpeter
1985; Lewis 1985; Lewis et al. 1985; Sullivan et al. 1981).
CGCG 127-055: Our H i detection (VHI =6626±3 km s−1, W50=215 km s−1 and IHI =1.9±0.08 Jy km s−1) is consistent with our previous
measurement (van Driel et al. 2000) made at Nan¸cay, with VHI =6656±21 km s−1, W50=183 km s−1 and IHI =2.1±0.3 Jy km s−1. Our new data
have a 3.6 times better rms noise level.
CGCG 128-072: Given the uncertainties, the global parameters of our H i detection (VHI =6795±6 km s−1, W50=119 km s−1 and IHI =0.82±0.07
Jy km s−1) are consistent with those of our previous Nan¸cay detection (van Driel et al. 2000), with VHI =6848±81 km s−1, W50=213 km s−1 and
IHI =1.4±0.6 Jy km s−1. Our new data have a 5.4 times better rms noise level.
CGCG 129-004: discrepant values for the optical redshift have been published, 4847 and 6729 km s−1 (Gavazzi et al. 1999a, and the compilation
by Falco et al. 1999); we find an H i value of 6736±7 km s−1, consistent with the latter value.
CGCG 130-003: there are two discrepant literature values for its optical redshift, 7094 and 22,425 km s−1 (Straus et al. 1992; Gregory et al.
1988); we find an H i value of 7140±4 km s−1, consistent with the former value.
CGCG 157-044: our H i detection (VHI =6607±4 km s−1, W50=240 km s−1 and IHI =0.84±0.08 Jy km s−1) has a 1.7 times lower line intensity
than our previous Nan¸cay detection (van Driel et al. 2000), with VHI =6628±36 km s−1, W50=309 km s−1 and IHI =1.5±0.4 Jy km s−1. The
difference is less than two times the uncertainty in the latter value, however, and therefore not significant.
CGCG 159-071: the global parameters of our H i detection (VHI =6971±1 km s−1 and W50=189 km s−1 and IHI =2.6±0.1 Jy km s−1) are
consistent with those of our previous Nan¸cay detection (van Driel et al. 2000), with VHI =6985±6 km s−1, W50=164 km s−1 and IHI =2.4±0.4
Jy km s−1.
CGCG 159-097: Its optical redshift, 6573±190 km s−1, is not well determined. We measured an H i value of 6424±30 km s−1, consistent with 3
of the published optical values -- only the optical velocity of 6883±75 km s−1 measured by van Haarlem et al. (1993) is in disagreement with all
other values.
CGCG 160-128: Our detection (VHI =7920±1 km s−1, W50=115 km s−1 and IHI =2.5±0.07 Jy km s−1) is consistent with our previous Nan¸cay
detection (VHI =7940±5 km s−1, W50=100 km s−1 and IHI =2.2±0.3 Jy km s−1), which was based on data with a 4 times higher rms of 2.8 mJy
(van Driel et al. 2000).
CGCG 161-051: We did not confirm our previous, quite tentative Nan¸cay detection (van Driel et al. 2000), with VHI =6993: km s−1, W50=235:
km s−1 and IHI =1.5: Jy km s−1. Our new spectrum has a 2.4 times better rms, of 1.5 mJy.
CGCG 161-054: Our detection (VHI =6756±3 km s−1, W50=284 km s−1 and IHI =1.3±0.1 Jy km s−1) is consistent with our previous, tentative
Nan¸cay detection, with VHI =6760: km s−1, W50=335: km s−1 and IHI =1.8: Jy km s−1, which was based on data with a 4 times higher rms of 3.2
mJy (van Driel et al. 2000).
Gavazzi et al.: H i observations in the Coma Supercluster
15
Fig. 7. H i spectra of the tentatively detected galaxies in the Coma Supercluster.
16
Gavazzi et al.: H i observations in the Coma Supercluster
Fig. 8. H i spectra of the tentatively detected galaxies in the Virgo cluster.
|
astro-ph/0208532 | 1 | 0208 | 2002-08-29T17:30:01 | The Multi-Band Magnification Bias for Gravitational Lenses | [
"astro-ph"
] | We present a generalization of the concept of magnification bias for gravitationally-lensed quasars, in which the quasars are selected by flux in more than one wavelength band. To illustrate the principle, we consider the case of two-band selection, in which the fluxes in the two bands are uncorrelated, perfectly correlated, or correlated with scatter. For uncorrelated fluxes, we show that the previously-held result - that the bias is the product of the single-band biases - is generally false. We demonstrate some important properties of the multi-band magnification bias using model luminosity functions inspired by observed correlations among X-ray, optical, infrared and radio fluxes of quasars. In particular, the bias need not be an increasing function of each flux, and the bias can be extremely large for non-linear correlations. The latter fact may account for the high lensing rates found in some X-ray/optical and infrared/radio selected samples. | astro-ph | astro-ph |
The Multi-Band Magnification Bias for Gravitational Lenses
J. Stuart B. Wyithe1, Joshua N. Winn2, David Rusin
Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138
[email protected]; [email protected]; [email protected]
ABSTRACT
the concept of magnification bias
for
We present a generalization of
gravitationally-lensed quasars, in which the quasars are selected by flux in more
than one wavelength band. To illustrate the principle, we consider the case of
two-band selection, in which the fluxes in the two bands are uncorrelated, per-
fectly correlated, or correlated with scatter. For uncorrelated fluxes, we show
that the previously-held result -- that the bias is the product of the single-band
biases -- is generally false. We demonstrate some important properties of the
multi-band magnification bias using model luminosity functions inspired by ob-
served correlations among X-ray, optical, infrared and radio fluxes of quasars. In
particular, the bias need not be an increasing function of each flux, and the bias
can be extremely large for non-linear correlations. The latter fact may account
for the high lensing rates found in some X-ray/optical and infrared/radio selected
samples.
Subject headings: Cosmology: gravitational lensing
1.
Introduction
If a massive galaxy lies along the line of sight to a background quasar, the galaxy may
act as a gravitational lens, magnifying and forming multiple images of the quasar. Beginning
with the pioneering work of Turner, Ostriker, & Gott (1984), many authors have computed
the number of lenses that should appear in well-defined samples of quasars, with particular
attention given to the dependence of this statistic on the vacuum energy density (Turner
1990; Kochanek 1996; Helbig et al. 1999; Sarbu, Rusin, & Ma 2001; Li & Ostriker 2002).
1Hubble Fellow
2NSF Astronomy & Astrophysics Postdoctoral Fellow
-- 2 --
These calculations must take into account not only the probability that a massive galaxy
will be aligned closely enough with a background quasar (the lensing cross-section), but
also the enhancement of the quasar flux due to lensing (the magnification bias). This is
because quasar samples are usually defined by observed flux in some wavelength band, and
gravitational lensing boosts the observed flux, thereby sampling a fainter portion of the
quasar luminosity function. For example, if intrinsically faint quasars are sufficiently more
numerous than bright quasars, then a quasar with a given observed flux is more likely to be
lensed than the cross-section alone would imply.
More recently attention has turned toward the information about galaxy mass profiles
that can be gleaned from lens statistics. These statistics include lensing rates (see, e.g.,
Keeton & Madau 2001; Wyithe, Turner, & Spergel 2001; Li & Ostriker 2002), the ratio of
four-image to two-image lenses (see, e.g., Rusin & Tegmark 2001; Finch et al. 2002), the image
separation distribution (see, e.g., Kochanek & White 2001) and the brightness distribution
of central images (see, e.g., Rusin & Ma 2001; Keeton 2001, 2002; Evans & Hunter 2002;
Oguri 2002). All of these applications of lens statistics require a good understanding of
magnification bias.
Borgeest, von Linde, & Refsdal (1991) noted that quasar samples selected by both radio
and optical flux measurements are subject to what they called a "double magnification bias."
If the radio and optical fluxes from a given quasar are nearly independent, then quasars
bright in both bands are especially likely to be lensed3. By assuming that gravitational
lensing produces only one possible value of magnification, and using power-law luminosity
functions for the optical and radio bands, Borgeest, von Linde, & Refsdal (1991) showed
that the resulting two-band magnification bias is the product of the bias factors computed
separately for each band.
It is timely to revisit the issue of multi-band magnification bias with a more general
approach. With the advent of large-area sky surveys at many wavelengths, it has become
possible to define samples of thousands of quasars by their observed fluxes in X-ray, optical,
infrared, and radio bands. Quasars appear in large numbers in, for example, the RASS
(ROSAT All-Sky Survey: Truemper 1982; Voges et al. 1999) and eventually ChaMP (Chandra
Multi-wavelength Project: Wilkes et al. 2001; Silverman et al. 2002) at X-ray wavelengths;
NVSS (NRAO-VLA Sky Survey: Condon et al. 1998) and FIRST (Faint Images of the Radio
Sky at Twenty centimeters: Becker, White, & Helfand 1995; White et al. 1997) at radio
3Note that the important property of the two bands is independence, not a large separation in wavelength
(as has since been stated in the literature; Bade et al. (1997)), although of course these two properties are
related.
-- 3 --
wavelengths; 2MASS (Two Micron All Sky Survey: Kleinmann et al. 1994) at near-infrared
wavelengths; and SDSS (Sloan Digital Sky Survey: York et al. 2000; Schneider et al. 2002)
at optical wavelengths. Cross-correlation of these catalogs (see, e.g., McMahon et al. 2001;
Ivezic et al. 2002) will become an increasingly important source of information about quasars
in general, and gravitational lens statistics in particular.
A few lenses have already been discovered using multi-band selection criteria, at lensing
rates that are larger than the 0.2 -- 1% typical of single-band lens surveys. Bade et al. (1997)
discovered the gravitational lens RX J0911.4+0551 by matching RASS sources with optical
sources from Schmidt plates. Of the ∼ 40 radio-quiet X-ray -- luminous high-redshift quasars
known, two are lensed (Wu, Bade, & Beckmann 1999). A search for very red quasars through
the matching of FIRST and 2MASS has identified two gravitational lenses out of thirteen
sources (Gregg et al. 2002; Lacy et al. 2002). None of these projects were designed explicitly
to discover gravitational lenses, although this is a realistic possibility for the future.
In this paper we investigate the magnification bias for quasar samples defined by mea-
surements in multiple wavelength bands. After presenting the basic formalism for N bands
(§2), we specialize to the case of two bands and consider some illustrative examples. We
consider the cases in which the two fluxes are uncorrelated (§2.1), perfectly correlated (§2.2),
and correlated with non-zero scatter (§2.3). We then use a realistic model of the optical lu-
minosity function for quasars to demonstrate a few interesting properties of the multi-band
magnification bias (§3); in particular, the bias does not necessarily increase with flux in
each band, and there is a profound difference between the case of a linear correlation and
a non-linear correlation with flux in another band. Finally, in §4 we summarize our results,
and discuss possible applications of this formalism to real quasar samples.
2. Magnification Bias and the Multiple Imaging Rate
We begin by reviewing the case of single-band magnification bias (Turner 1980; Turner,
Ostriker, & Gott 1984). In a sample of quasars at redshift z with (apparent4) luminosity L1,
the fraction of multiple-image lensed quasars is
F (L1, z) ≈
,
(1)
B1τmult
B1τmult + (1 − τmult)
where τmult is the cross-section for multiple imaging, and B1(L1) is the magnification bias.
For τmult ≪ 1 and B1τmult ≪ 1, this reduces to the usual expression F (L1, z) = B1τmult. The
4By "apparent," we mean that L1 is the luminosity inferred from the observed flux and the luminosity
distance, without taking into account the possible magnification due to lensing.
-- 4 --
magnification bias is evaluated as
B1(L1, z) = R ∞
0
dµ
µ
dP
dµ Φ1(L1/µ, z)
Φ1(L1, z)
,
(2)
where Φ1(L1, z) is the quasar luminosity function, µ is the sum of the unsigned magnifications
of the multiple images, and dP
dµ is the probability distribution for µ, taken for a singular
isothermal sphere throughout the paper (dP = 8µ−3dµ for µ ≥ 2). This expression can be
understood as a likelihood ratio. The denominator is the likelihood that the quasar is drawn
from the sample of unlensed quasars with luminosity L1 (within dL1). The numerator is the
likelihood that the quasar is drawn from the fainter sample of quasars with luminosity L1/µ
(within dL1/µ), summed over all possible values of µ.
Understood this way, the generalization to N bands is straightforward. We require
knowledge of the multivariate luminosity function, ΦN(L1, L2, L3, ..., LN, z). For a point
source, the magnification is the same for all bands, because gravitational lensing is achro-
matic5. The multi-band magnification bias is therefore
B1...N (L1, L2, L3, ..., LN , z) = R ∞
0 dµ 1
µN
dP
dµ Φ1...N (L1/µ, L2/µ, L3/µ, ..., LN /µ, z)
Φ1...N (L1, L2, L3, ..., LN , z)
.
(3)
The dependence of B on the apparent luminosities of the quasars depends on the corre-
lations, if any, between the intrinsic luminosities of the quasars in those bands. To illustrate
the interesting properties that can result, in the following sections we concentrate on the
simplest non-trivial case, the two-band magnification bias. All of the results are easily gen-
eralized to N bands.
2.1. Two-Band Magnification Bias: No Correlation
Borgeest, von Linde, & Refsdal (1991) considered quasars observed at both optical
and radio wavelengths, and assumed a power-law luminosity function for each band. They
5Gravitational lensing magnification is sensitive to source size. Therefore if the emission regions for
the two bands differ greatly in their spatial extent, then there is the possibility of wavelength-dependent
magnification. For example, the small optical emission region of a quasar may be microlensed by stars in the
lensing galaxy, whereas the more extended emission regions at infrared or radio wavelengths is not generally
microlensed (see Wyithe & Turner (2002) for a recent discussion of microlensing and magnification bias). We
ignore the possibility of microlensing in this paper, but note that since the mean magnification of microlensed
sources equals the magnification of un-microlensed sources, the results presented will be qualitatively correct,
even for correlations involving bands that are subject to microlensing.
-- 5 --
showed that if the fluxes in these bands are statistically independent, and if there is only
one possible value of the lensing magnification, then the two-band magnification bias is
equal to the product of the biases that would be computed separately for the optical and
radio bands. This result is not true in general. As we show below, even if the two bands are
independent, the result does not hold because real gravitational lenses produce a distribution
of magnifications.
First, we reproduce the result of Borgeest, von Linde, & Refsdal (1991) using our for-
malism. For N = 2, Eq. (3) is
B12(L1, L2, z) = R ∞
0
dµ
µ2
dP
dµ Φ12(L1/µ, L2/µ, z)
Φ12(L1, L2, z)
.
(4)
If the bands are independent, then Φ12(L1, L2, z) = Φ1(L1, z)Φ2(L2, z). The lens model used
by Borgeest, von Linde, & Refsdal (1991) can be described by dP
dµ = δ(µ− µ0), in which case
B12(L1, L2, z) =
1
µ2
0
Φ12(L1/µ0, L2/µ0, z)
Φ12(L1, L2, z)
=
1
µ2
0
Φ1(L1/µ0, z)Φ2(L2/µ0, z)
Φ1(L1, z)Φ2(L2, z)
= B1(L1, z)B2(L2, z).
(5)
This results fails for the more realistic case in which there is a range of possible mag-
nifications, because dP
dµ appears once in the numerator of the multi-band magnification bias,
but appears separately in each numerator in the product of the single-band biases. For
example, following Borgeest, von Linde, & Refsdal (1991), suppose Φ1(L1, z) = Φ1,∗Lα1
1 and
Φ2(L2, z) = Φ2,∗Lα2
2 . If we adopt the magnification distribution appropriate for an isothermal
sphere ( dP
dµ = 8
(6)
(7)
µ3 for all µ ≥ 2), then
B12(L1, L2, z) =Z ∞
2
dµ
µ2
8
µ3
1
µα1µα2
=
8
4 + α1 + α2
2−(4+α1+α2),
for α1 + α2 > −4. Analogous calculations of the single-band bias factors give
B1(L1, z)B2(L2, z) =
16
(3 + α1)(3 + α2)
2−(4+α1+α2),
which can be either larger or smaller than Eq. (6).
2.2. Two-Band Magnification Bias: Perfect Correlation
If the two bands are perfectly correlated, with L2 = f (L1), one might expect that
no new information is provided by the observation in the second band, and therefore that
-- 6 --
the two-band bias is equal to the single-band bias for either band. This is not quite true.
Gravitational magnification multiplies both fluxes by the same factor. If the unmagnified
fluxes are linearly correlated, then the magnified fluxes also obey the correlation. However,
if the correlation is non-linear, then the magnified fluxes do not obey the correlation, and the
source must be gravitationally lensed. In the the appendix, we derive this result formally,
by calculating the magnification bias for general correlations (see the next section) in the
limit of zero scatter.
2.3. Two-Band Magnification Bias: Imperfect Correlation
More generally, L1 and L2 are correlated with some intrinsic scatter. On physical
grounds we expect the magnitude of the scatter to scale with the luminosity [i.e. ∆L2/L2 ∼
g(∆L1/L1)], which makes it convenient to use logarithmic variables l = log L. Suppose that
the correlation between L1 and L2 is a power-law, L1 = Lγ
2, or l1 = γl2. Because of the
correlation, it is convenient to express the luminosity function in terms of the new variables
u1 ≡ 1
γ l1 + l2 and u2 ≡ −γl1 + l2, which describe the location parallel and perpendicular
In these variables, the luminosity function
to the correlation, respectively (see Fig. 1).
(expressed in density per square logarithmic interval) is Ψ12(u1, u2, z) = (γ + 1
γ )Φ12(l1, l2, z).
The luminosity function can also be written Ψ12(u1, u2, z) = Ψ1(u1, z)p(u2u1, z), where
Ψ1(u1, z) is the luminosity function in the new variable u1 and p(u2u1, z) is the conditional
probability of u2 given u1. Because we expect the scatter to be symmetric in reflection about
the correlation6 , we assume p(u2u1, z) to be Gaussian with variance σ, hence
Ψ12(u1, u2, z) = Ψ1(u1, z)
1
√2πσ
exp(cid:18)−
u2
2
2σ2(cid:19).
Defining M = log µ, the magnification bias is
B12(l1, l2, z) = Z ∞
= Z ∞
0
0
dM
dM
dP
dM
dP
dM
Φ12(l1, l2, z)
Φ12(l1 − M, l2 − M, z)
γ )Ψ12hu1 − (1 + 1
(γ + 1
γ )M, u2 + (γ − 1)M, zi
(γ + 1
γ )Ψ12(u1, u2, z)
(8)
.
(9)
6As an example of why we expect the scatter to be symmetric in reflection about the correlation, consider
a sample of quasars with flux measured in two optical wave bands, say r and i. We would expect to find a
variation of bias with i-band at fixed r-band that is qualitatively similar to the variation with r-band at fixed
i-band. This symmetry in the magnification bias requires symmetry of the scatter in reflection about the
correlation. We therefore choose to model the scatter as a symmetric function in logarithms of luminosity;
that is, defined normal to the correlation (i.e. along the u2 axis).
-- 7 --
dM ′ dP
dM ′
Ψ1 [u1 − M ′, z]
Ψ1(u1, z)
exp −
Inserting Eq. (8),
B12(l1, l2, z) =Z ∞
0
1
2σ2 "(cid:18)u2 +
γ(γ − 1)
1 + γ
M ′(cid:19)2
2#!,
− u2
(10)
where we have defined M ′ = (1 + 1
γ )M. This expression illustrates many important points.
First, if the correlation is linear (γ = 1) then B12 is independent of u2, and the contours of
constant bias run normal to the correlation (i.e. along lines of constant l2 + l1). Furthermore,
if Ψ1(u1, z) is a monotonically decreasing function of u1, then we find that B12(l1, l2, z) is an
increasing function of both l1 and l2 (since u2 increases monotonically with both l1 and l2).
This example will be further explored in Case 1 of §3. Eq. (10) also demonstrates the behavior
arising from non-linear correlations (γ 6= 1). Here the exponential plays an important role;
it introduces an asymmetry in the bias across the correlation. If, for example, γ > 1, then
large biases can result from negative values of u2 (i.e. below the correlation), because the
exponent becomes positive. On the other hand, the exponent is negative for all u2 > 0, and
hence the bias above the correlation is small. This example will be further explored in Case
3 of §3.
In the following section, we evaluate Eq. 10 numerically, with more realistic assumptions,
in order to illustrate these and other interesting and potentially observable properties of the
multi-band magnification bias.
3. Magnification Bias for Illustrative Bi-Variate Luminosity Functions
We consider measurements made in two bands, and a power-law correlation between
the two bands: L2 = Lγ
1. As in §2.3, the scatter (normal to the correlation) is assumed to be
Gaussian in logarithms with half-width σ. We consider 4 examples of Eq. (8). The first two
examples involve linear correlations (γ = 1), and the second two examples involve non-linear
correlations.
For the first of the two correlated bands, we use a luminosity function Φ1(L1, z) that is
appropriate for optical wavelengths. A good representation of the observed optical quasar
luminosity function at redshifts z . 3 is provided by the following double power-law form
(Boyle, Shanks, & Peterson 1988; Pei 1995):
Φo(L, z) =
Φ∗/L∗(z)
[L/L∗(z)]βl + [L/L∗(z)]βh
.
(11)
At the faint end, the logarithmic slope of this function is −βl = −1.58, while at the bright
end the slope is −βh = −3.43 (Boyle et al. 2000). Moreover, all dependence on redshift (for
-- 8 --
z . 3) is in the break luminosity L∗(z). We therefore show the luminosity function in units
of L∗ throughout the remainder of this paper. Setting Φ1(L1) = Φo(L1), we find from Eq. 8
that Φo(L1) is related to Ψ1(u1, u2) through
Φo(L1) =Z ∞
0
dL2Ψ1(u1)
1
√2πσ
exp(cid:18)−
u2
2
2σ2(cid:19).
(12)
This equation defines the functions Ψ1(u1) used in this section.
1. γ = 1.0, σ = 0.15: A linear correlation with constant scatter. This is the situation that
might be expected between two different optical bands. The contours of Φ12 (gray),
and of B12 (black), are shown in the top left panel of Fig. 2. As was derived in the
previous section (and might be expected intuitively), the bias increases monotonically
with both luminosities, and is constant along lines normal to the correlation.
The top right panel shows the corresponding single-band luminosity function (gray
line), and magnification bias (dotted black line). These functions are the same for
both bands because of the linear correlation. In addition we plot B12(L1, L2) for two
paths through (L1, L2)-space: one at fixed L1 (thick dashed line, plotted as a function
of L2) and the other below but parallel to the correlation (thin dashed line, plotted as
a function of L1).
2. γ = 1.0, σ = 0.15 − 0.02 log u1: Same as the previous example, except in this case we
allow the logarithmic scatter to depend on luminosity. The results are shown in the
lower panels of Fig. 2, in the same format as the previous example.
The contours of magnification bias wrap around the contours of Φ12, increasing rapidly
as one moves normal to the correlation (along the u2 axis). This can be understood as
follows. Magnification draws quasars from regions of lower intrinsic luminosity, where
the probability density of quasars (as given by Φ12) is larger. This is especially true
when the observed luminosities fall at some distance from the correlation, because the
scatter in the correlation is larger at lower luminosities.
The dependence of B12 on L2, for fixed L1, is again shown by the thick dark dashed line.
Interestingly, the dependence is not monotonic. For small values of L2 the magnification
bias is large. The bias decreases as L2 rises through the expected intrinsic value,
and then increases again. The bias along the path denoted by the thin dashed line
demonstrates that the bias can become very large for sources below the correlation.
3. γ = 1.5, σ = 0.2: A non-linear correlation with a constant scatter. This situation
approximates the correlation that has been observed between the X-ray (L2) and optical
-- 9 --
(L1) bands for quasars (Brinkmann et al. 2000). Results for this case are shown in the
top two panels of Fig. 3, in the same format as the previous examples.
In this case, Φ2(L2) (thin gray line) is a flatter function than Φ1(L1) (thick gray line),
and therefore the single-band bias B2 (thin dotted line) is smaller than B1 (thick dotted
line). Although B12 is an increasing function of L1, it is actually a decreasing function
of L2 (for fixed L1). This runs counter to the naive expectation that the brighter the
quasar is (regardless of band), the more likely it is to be lensed. The reason is that
when L2 is smaller than expected from the correlation, reducing both luminosities by
the same factor µ (along a line of unit slope, in the top left panel) brings one to a
region of much higher probability.
4. γ = −1.0, σ = 0.3: An anti-correlation with constant scatter. This is a somewhat
artificial example, but we might imagine there are two ways for a quasar with a fixed
energy source to radiate its energy, and one of these ways can be blocked by a variable
amount. For example, the optical and far-infrared luminosities might be expected to
exhibit some degree of anti-correlation due to dust obscuration.
The results are plotted in the lower panels of Fig. 3. The contours of magnification
bias are parallel to the anti-correlation. For small luminosities the bias is smaller than
unity. Quasars in this region are less likely to be lensed than the cross-section alone
would imply, because lensed quasars would be drawn from a population with very small
density. Conversely, for large luminosities, the bias becomes arbitrarily large.
4. Discussion
The multi-band magnification bias is an a posteriori statistic. It is used to estimate the
probability that the apparent luminosities of a given quasar, as measured in several bands,
are due to gravitational magnification, rather than being intrinsic to the quasar. When a
sample of quasars is selected through the matching of sources in two different catalogs, both
fluxes must be used to perform this calculation. One must also have some knowledge of the
intrinsic correlation (if any) of the fluxes, and the distribution of magnifications produced
by lensing.
One might expect that the multi-band bias is maximized when the bands are uncor-
related (an example of which is shown for the radio-optical correlation of SDSS early data
release quasars in Fig. 4), since in that case there is no redundant information in the flux
measurements. Upon further reflection, or using the mathematics developed in this paper,
one realizes that this is not true -- the relevant information is how discrepant the observed
-- 10 --
fluxes are from the correlation, and whether the discrepancy can be made smaller if the
observed fluxes are all reduced by a constant factor.
Many of the illustrative examples presented in this paper approximate certain correla-
tions that have been observed for real quasars. In particular, the multi-band magnification
bias may result in very high lens fractions for certain quasar samples. First, we consider the
case of a quasar sample selected by optical colors. The top left panel of Fig. 4 is a logarithmic
plot of SDSS i-band vs. r-band fluxes for the SDSS early data release quasars (Schneider et
al. 2002). The data show a linear correlation with scatter, the magnification bias for which
is illustrated in the top panels of Fig. 2. The magnification bias must be computed using
both optical measurements, unless the sample is 100% complete in one filter (i.e., unless
after selecting quasars in i, the r-band magnitude was measured in every single case). As
an example consider the sample of SDSS z > 5.8 quasars (Fan et al. 2001). Since the z-
band selection is at ∼ 1100A in the rest-frame, quasars with fixed absolute B magnitude
(∼ 4400A) are more likely to be selected if they are bluer than average. Thus a sample of
quasars selected in this manner will be bluer than average and lie blueward of the correlation
on a plot of the intrinsic correlation between MV and MB. The magnification bias for these
sources may be significantly smaller than that computed using only extrapolations of the
B-band luminosity function (Wyithe & Loeb 2002; Commerford, Haiman, & Schaye 2002).
r
Next, we consider examples of non-linear luminosity correlations. Brinkmann et al.
(2000) measured the correlation between ROSAT X-ray and FIRST radio fluxes for matched
quasar samples. They find that while radio-quiet quasars show a linear relationship between
X-ray and radio luminosity, radio-loud quasars have an X-ray flux Lx that varies with the
radio luminosity Lr as Lx ∝ L0.48±0.05
with an intrinsic scatter of ∼ 0.2 dex. Furthermore,
Brinkmann et al. (2000) showed that the X-ray luminosity correlates non-linearly with optical
luminosity Lo, following Lx ∝ L1.42±0.09
. The second of these correlations (in flux) is plotted
in Fig. 4 for quasars in the SDSS Early Data Release (Schneider et al. 2002), but can be
seen more clearly in Fig. 14 of Brinkmann et al. (2000). The multi-band magnification
bias corresponding to the second correlation7 is illustrated in the top panels of Fig. 3. The
magnification bias can be extremely large for sources that are luminous at both optical
and X-ray wavelengths. This may be the explanation for the apparently high probability
of lensing in bright X-ray selected quasar catalogs (Bade et al. 1997). The location of the
gravitational lens RX J0911.4+0551, which was selected from cross-correlation of optical and
X-ray catalogs, is shown on this plot by the large dot8. The fluxes place the quasar below
o
7Note that while we have presented results for a non-linear correlation with index γ > 1, the result for
γ < 1 is simply obtained through reversal of the axes.
8We used the integrated R from Bade et al. (1997) and color transformations from Fukugita, Shimasaku
-- 11 --
the correlation, in the region where we expect the magnification bias to be large (see the
upper panels of Fig. 3). The lens HE 1104 -- 1805 is also X-ray loud (Wisotzki et al. 1993;
Reimers et al. 1995). While this lensed quasar was not discovered through cross-correlation
between catalogs, it is interesting to note its location on this plot, shown by the open square
in Fig. 4. The quasar is found to be very bright in both bands, and is again in the region of
high magnification bias. It is suggestive that the two X-ray loud gravitational lenses both
appear to lie in the region of high magnification bias, as expected.
Finally, the multi-band magnification bias may also provide an explanation for the large
gravitational lens fraction (2 out of 13) found through the matching of FIRST and 2MASS
sources (Gregg et al. 2002; Lacy et al. 2002). Fig. 4 shows the correlation for near infrared
luminosities verses radio luminosities compiled from table 1 of Barkhouse & Hall (2001). The
radio/near-IR correlation appears to be steeper than linear. If true we might expect very
large biases for luminous near-IR sources. The top panel of Fig. 3 demonstrates magnification
bias for a non-linear correlation, and shows that the bias of around 100 necessary to achieve
a lens fraction of 2/13 is possible.
5. Summary
This paper has discussed the multi-band magnification bias for gravitational lensing
with arbitrary luminosity functions in several bands. Previous discussion of the multi-band
magnification bias (Borgeest, von Linde, & Refsdal 1991) focused on the case where the fluxes
in the two bands are independent. If a single value for the lens magnification is considered,
they showed that this assumption leads to a multiple magnification bias that is equal to the
product of the single-band biases. However, we have shown that this equality breaks down
in the more realistic case when there is a distribution of possible magnifications.
We also discussed the multi-band magnification bias when the fluxes in the various bands
are correlated. In the case of a perfect (i.e. zero scatter) linear correlation, the information
from the second band does not change the magnification bias. However, if the correlation
is non-linear, then sources with fluxes that obey the correlation cannot be lensed. On the
other hand, sources with fluxes that do not obey the correlation must be lensed.
Of course, real correlations have intrinsic scatter. We have calculated the multi-band
magnification bias for bi-variate luminosity functions with finite scatter about both linear
and non-linear correlations. For a linear correlation (as expected for a quasar sample selected
& Ichikawa (1995).
-- 12 --
by optical colors) we find that the magnification bias is an increasing function of either flux.
Calculations of lens statistics from incomplete color-selected quasar samples should therefore
account for the multi-band magnification bias.
Non-linear correlations (and anti-correlations) with finite scatter were also explored. If
the fluxes in two bands are correlated through a relation that is steeper than linear, then
sources that lie below the correlation can be subject to a very large bias. The observed
correlation between X-ray and optical flux (and possibly between infrared and radio flux)
for quasars is steeper than linear. Suggestively, the two known X-ray loud gravitationally
lensed quasars lie below the X-ray/optical correlation in the region of large magnification
bias. Thus the multiple magnification bias may provide an explanation for the large lensing
rates found in X-ray/optical and infrared/radio selected samples.
The authors wish to acknowledge discussions with Chris Kochanek that led us to work
on this topic. We also thank Wai-Hong Tham and Lara Winn for enduring conversations.
J.S.B.W. is supported by a Hubble Fellowship grant from the Space Telescope Science Insti-
tute, which is operated by the Association of Universities for Research in Astronomy, Inc.,
under NASA contract NAS 5-26555. J.N.W. is supported by an Astronomy & Astrophysics
Postdoctoral Fellowship, under NSF grant AST-0104347.
Bade, N., Siebert, J., Lopez, S., Voges, W., & Reimers, D. 1997, A&A, 317, L13
REFERENCES
Barkhouse, W. A. & Hall, P. B. 2001, AJ, 121, 2843
Becker, R. H., White, R. L., & Helfand, D. J. 1995, ApJ, 450, 559
Borgeest, U., von Linde, J., & Refsdal, S. 1991, A&A, 251, L35
Boyle, B.J., Shanks, T., & Peterson, B.A. 1988, MNRAS, 235, 935
Boyle, B. J., Shanks, T., Croom, S. M., Smith, R. J., Miller, L., Loaring, N., & Heymans,
C. 2000, MNRAS, 317, 1014
Brinkmann, W., Laurent-Muehleisen, S. A., Voges, W., Siebert, J., Becker, R. H., Brother-
ton, M. S., White, R. L., & Gregg, M. D. 2000, A&A, 356, 445
Commerford, J., Haiman, Z. & Schaye, J., 2002, astro-ph/0206441
Condon, J.J., et al. 1998, AJ, 115, 1693
-- 13 --
Evans, N.W. & Hunter, C., 2002, astro-ph/0204206
Fan, X. et al. 2001, AJ, 122, 2833
Finch, T.K., Carlivati, L.P., Winn, J.N. & Schechter, P.L., 2002, astro-ph/0205489
Fukugita, M., Shimasaku, K., & Ichikawa, T. 1995, PASP, 107, 945
Gregg, M. D., Lacy, M., White, R. L., Glikman, E., Helfand, D., Becker, R. H., & Brotherton,
M. S. 2002, ApJ, 564, 133
Helbig, P., Marlow, D., Quast, R., Wilkinson, P. N., Browne, I. W. A., & Koopmans, L. V. E.
1999, A&AS, 136, 297
Ivezic, Z., et al., 2002, astro-ph/0202408
Keeton, C. R. 2001, ApJ, 561, 46
Keeton, C. R. 2002, astro-ph/0206243
Keeton, C. R. & Madau, P. 2001, ApJ, 549, L25
Kleinmann, S. G. et al. 1994, Ap&SS, 217, 11
Kochanek, C. S. 1996, ApJ, 466, 638
Kochanek, C. S. & White, M. 2001, ApJ, 559, 531
Koopmans, L.V.E. & Treu, T., 2002, ApJ, 568, L5
Kundic, T., et al., 1997, ApJ, 482, 75
Lacy, M., Gregg, M., Becker, R. H., White, R. L., Glikman, E., Helfand, D., & Winn, J. N.
2002, AJ, 123, 2925
Li, L. & Ostriker, J. P. 2002, ApJ, 566, 65
McMahon, R.G., Irwin, M.J., 1992, in Digitised Optical Sky Surveys, eds. H.T. MacGillivray
& E.B. Thomson (Dordrecht: Kluwer), p417
McMahon, R.G., White, R.L., Helfand, D.J. & Becker, R.H., 2001, astro-ph/0110437
Oguri, M. 2002, astro-ph/0207520
Pei, Y. C. 1995, ApJ, 438, 623
-- 14 --
Reimers, D., Bade, N., Schartel, N., Hagen, H.-J., Engels, D., & Toussaint, F. 1995, A&A,
296, L49
Rusin, D. & Ma, C.-P. 2001, ApJ, 549, L33
Rusin, D. & Tegmark, M. 2001, ApJ, 553, 709
Sarbu, N., Rusin, D., & Ma, C.-P. 2001, ApJ, 561, L147
Schechter, P.L., et al. 1997, ApJ, 475, L85
Schneider, D. P. et al. 2002, AJ, 123, 567
Silverman, J.D., et al. 2002, AJ, 569, L1
Truemper, J. 1982, Advances in Space Research, 2, 241
Turner, E. L., Ostriker, J. P., & Gott, J. R. III 1984, ApJ, 284, 1
Turner, E. L. 1990, ApJ, 365, L43
Turner, E. L. 1980, ApJ, 242, L135
Voges, W. et al. 1999, A&A, 349, 389
Walsh, D., Carswell, R. F., & Weymann, R. J. 1979, Nature, 279, 381
White, R. L., Becker, R. H., Helfand, D. J., & Gregg, M. D. 1997, ApJ, 475, 479
Wilkes, B.J., et al. 2001, in ASP Conf. Ser. 232, The New Era of Wide-Field Astronomy, ed.
R.G. Clowes, A.J. Adamson & G.E. Bromage (San Francisco: ASP), 47
Wisotzki, L., Koehler, T., Kayser, R., & Reimers, D. 1993, A&A, 278, L15
Wu, X., Bade, N., & Beckmann, V. 1999, A&A, 347, 63
Wyithe, J. S. B., Turner, E. L., & Spergel, D. N. 2001, ApJ, 555, 504
Wyithe, J. S. B., & Turner, E. L., 2002, ApJ, in press
Wyithe, J. S. B., & Loeb, A. 2002, ApJ, in press
York, D. et al., 2000, AJ, 120, 1579
This preprint was prepared with the AAS LATEX macros v5.0.
-- 15 --
A. Two-Band Magnification Bias for Perfect Correlations
In this appendix we derive the results mentioned in § 2.2 for the two-band magnification
It is
bias in the case of perfect correlations by taking the limit of small σ in Eq. (10).
convenient to rewrite the exponential as follows
Ψ1 [u1 − M ′, z]
1
lim
σ→0
exp(cid:18)−
2σ2 M ′ γ(γ − 1)
γ + 1 (cid:20) γ(γ − 1)
γ + 1
M ′ + 2u2(cid:21)(cid:19).
lim
σ→0
B12(l1, l2, z) =Z ∞
0
dM ′ dP
dM ′
Ψ1(u1, z)
(A1)
First, consider the case of a linear correlation (γ = 1). In this case we find the expo-
nential function is unity for all σ > 0, and Eq. (A1) reduces to
lim
σ→0
B12(l1, l2, z) =Z ∞
0
dM ′ dP
dM ′
Ψ1 [u1 − M ′, z]
Ψ1(u1, z)
=Z ∞
0
dM
dP
dM
Ψ1 [2(l1 − M), z]
Ψ1(2l1, z)
.
(A2)
Note that in this case we must have l1 = l2. We can relate Ψ1(u1, z) to Φ1(l1, z) in the limit
of small σ:
lim
σ→0
Φ1(l1, z) = Z ∞
2Z ∞
0
1
=
0
dl2 lim
σ→0
Φ12(l1, l2) =Z ∞
0
Ψ12(u1, u2)
2
=
1
2Z ∞
0
dl2Ψ1(l1 + l2)δ(l1 + l2) =
Ψ1(2l1).
dl2 lim
σ→0
1
2
dl2Ψ1(u1)δ(u2)
(A3)
Thus, for a perfect linear correlation, we find that
lim
σ→0
B12(l1, l2, z) =Z ∞
0
dM
dP
dM
limσ→0 Φ1 [l1 − M, z]
limσ→0 Φ1(l1, z)
= B1(l1, z),
(A4)
which is the single band magnification bias. Therefore, if two luminosities obey a perfect
linear correlation, the magnification bias is simply equal to the single-band magnification
bias computed from either luminosity.
and 0 < M ′
Next consider the case of a non-linear correlation (γ 6= 1). Specifically, we assume γ > 1
exp(cid:18)−
min < M ′ < ∞. Then
2σ2 M ′ γ(γ − 1)
M ′ + 2u2(cid:21)(cid:19) = 0
γ + 1 (cid:20)γ(γ − 1)
u2 > −
γ + 1
if
1
lim
σ→0
γ(γ − 1)
γ + 1
γ(γ − 1)
γ + 1
γ(γ − 1)
γ + 1
M ′
2
M ′
2
M ′
.(A5)
2
= 1
if
u2 = −
= ∞
if
u2 < −
Note that only sources with u2 = 0 or u2 < − γ(γ−1)
that limσ→0 B12(l1, l2, z) = ∞ if u2 < − γ(γ−1)
γ+1 M ′
γ+1 M ′
min are allowed. As a result we find
min. Sources that lie on the correlation have
-- 16 --
u2 = 0, and therefore limσ→0 B12(l1, l2, z) = 0. Thus if the correlation is non-linear, sources
that lie on the correlation cannot be lensed, while sources that lie off the correlation must
be lensed.
-- 17 --
u
2
u 1
Fig. 1. -- Schematic of variables defining the luminosity function for non-zero scatter about
a power-law correlation. The grey lines are contours of a bi-variate luminosity function and
are shown for context.
-- 18 --
Fig. 2. -- Bi-variate magnification biases for linear correlations with scatter. Top: Linear
correlation, with a scatter that is insensitive to luminosity (γ = 1.0, σ = 0.15). Bottom:
Linear correlation with a scatter that decreases with luminosity (γ = 1.0, σ = 0.15 −
0.02 log u1). The left hand panels show contours of the bi-variate luminosity function (grey
lines). The solid lines are contours of magnification bias. The right hand panels show the
corresponding single band luminosity functions (grey lines). Also shown are the single band
magnification biases (dotted lines), and the magnification bias along the paths denoted by
the dashed lines in the left hand figure. The bias for the path denoted by the thin dashed
line is plotted as a function of L1, while the bias along the thick dashed line is plotted as a
function of L2. Because of the linear correlation, the single band luminosity functions and
magnification biases are identical for the two bands.
-- 19 --
Fig. 3. -- Bi-variate magnification biases for non-linear correlations with scatter. Top: Non-
linear correlation, with a logarithmic scatter that is insensitive to luminosity (γ = 1.5,
σ = 0.2). Bottom: Non-linear anti-correlation, with a logarithmic scatter that is insensitive
to luminosity (γ = −1.0, σ = 0.3). The left hand panels show contours of the bi-variate
luminosity function (grey lines). The solid lines are contours of magnification bias. The
right hand panels show the corresponding single band luminosity functions (grey lines) and
the single band magnification biases (dotted lines). Thick and thin lines denote quantities
in L1 and L2 respectively. Also shown are the magnification biases along the paths denoted
by the dashed lines in the left hand figure. The bias for the path denoted by the thin dashed
line is plotted as a function of L1, while the bias along the thick dashed line is plotted as a
function of L2.
-- 20 --
xL Lr1.5
Fig. 4. -- Correlations in different bands. Top Left: SDSS i-band vs. SDSS r-band flux.
(Schneider et al. 2002) Top Right: FIRST radio flux vs. SDSS r-band flux (Schneider
et al. 2002). Lower Left: ROSAT X-ray counts vs. SDSS r-band flux (Schneider et
al. 2002) for quasars with redshifts larger than 0.5. The large dot in this panel repre-
sents RX J0911.4+0551, while the open square shows the location of HE 1104-1805. Bottom
Right: Radio vs. Near IR luminosity (Barkhouse & Hall 2001). In the left hand panels the
observed correlation lines are drawn to guide the eye.
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.